You are on page 1of 46

 

 
Different pools of black carbon in sediments from the Gulf of Cádiz (SW
Spain): Method comparison and spatial distribution

Laura Sánchez-Garcı́a, José R. de Andrés, Yves Gélinas, Michael W.I.


Schmidt, Patrick Louchouarn

PII: S0304-4203(13)00036-4
DOI: doi: 10.1016/j.marchem.2013.02.006
Reference: MARCHE 2983

To appear in: Marine Chemistry

Received date: 23 July 2012


Revised date: 10 February 2013
Accepted date: 11 February 2013

Please cite this article as: Sánchez-Garcı́a, Laura, de Andrés, José R., Gélinas, Yves,
Schmidt, Michael W.I., Louchouarn, Patrick, Different pools of black carbon in sediments
from the Gulf of Cádiz (SW Spain): Method comparison and spatial distribution, Marine
Chemistry (2013), doi: 10.1016/j.marchem.2013.02.006

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Different pools of black carbon in sediments from

PT
the Gulf of Cádiz (SW Spain): method comparison

RI
SC
and spatial distribution

NU
MA
D
P TE
CE
AC

1
ACCEPTED MANUSCRIPT

PT
Black carbon (BC) accounts for a quantitatively important fraction of the carbon sink in marine

sediments. Yet, the determination of its role in the environment remains controversial largely due

RI
to the large variability of BC estimates yielded by the diversity of analytical techniques presently

SC
used. In this study, we combine the application of a thermal oxidation technique (graphitic BC,

NU
or GBC) with a molecular method (benzene polycarboxylic acids, or BPCA) for determining the

presence of different types of BC in 15 marine sediments from the Gulf of Cádiz (GoC).
MA
Accounting for the two condensation end-members of the BC set (chars and graphitic carbon)

constitutes the first approach to a comprehensive estimate of the BC burial in the GoC
D

Continental Shelf. Measurements of BC in the GoC sediments ranged from 0.01 to 0.23% dw,
TE

with different ranges obtained with each technique (0.01-0.11 for GBC versus 0.01-0.23 for
P

BPCA). BC resulting from the use of the thermal GBC method (measuring soot and graphitic
CE

carbon) or the BPCA as molecular markers (detecting mostly char) showed different spatial

distributions in the GoC and varied from 13·103 to 26·103 t of BC/yr depending on the method
AC

used. Mostly petrogenic in origin GBC was less abundant and more homogeneously distributed

along the Guadiana River plume, whereas largely pyrogenic BPCA-BC was generally more

abundant but specifically enriched in the mid-region of the plume. The Guadiana River plume

combined with hydrodynamic sorting are argued to be the main distribution factors for graphitic

carbon and charcoals, whereas airborne soot seems to compose an undetermined and potentially

overlapping fraction of both analytical measurements (GBC and BPCA). The structural analysis

of the thermal residues and the molecular markers provided valuable information on the BC

nature in the environmental matrices. The results of this study illustrate the importance of

2
ACCEPTED MANUSCRIPT

considering different analytical windows of the BC catchall for constraining regional burial sinks

of heterogeneous BC sources.

PT
RI
SC
NU
MA
D
P TE
CE
AC

3
ACCEPTED MANUSCRIPT

1. Introduction

Refractory organic residues largely derived from incomplete combustion of biomass or fossil

fuels (black carbon, BC) are receiving increasing interest from the scientific community due to

PT
their implication in important biogeochemical processes (e.g. Earth´s radiative heat balance,

RI
Earth´s fire history, pollutant carrier, etc.) and potential role in global warming (IPCC, 2007).

SC
Terrestrial in origin, BC occurs ubiquitously in aquatic systems; water column (e.g. Dittmar et

al., 2012; Stubbins et al., 2012) and sediments (e.g. Goldberg, 1985; Gustafsson and Gschwend,

NU
1998; Masiello and Druffel, 2003; Middelburg et al., 1999; Suman et al., 1997; Verardo and
MA
Ruddiman, 1996), where it builds up and represents a significant sink in the global carbon cycle

(Kuhlbusch, 1998; Masiello and Druffel, 1998). BC is a quantitatively important component of


D

the sedimentary carbon burial flux, with the continental margins receiving the major (~90%)
TE

contribution (Suman et al., 1997). Quantifying one of the main sinks of carbon in the world (i.e.

ocean), is essential to better constrain the global BC budget, yet the global inventory and burial
P
CE

flux of BC in marine sediments remains unclear (e.g. Dickens et al. 2004a; Flores-Cervantes et

al., 2009; Gustafsson and Gschwend, 1998; Lohmann et al., 2009; Sánchez-García et al., 2012;
AC

Suman et al., 1997).

One of the main limitations for constraining global BC inventories is the analytical challenge of

measuring BC in natural matrices. Considered as a catchall of carbonaceous moieties with

aromatic structure, BC offers analytical complexity to be assessed depending on its different

origin and properties (aromaticity, condensation degree and recalcitrance; Hammes et al., 2007;

Harvey et al, 2012a; Masiello, 2004). The often overlapping boundaries between the constituents

in the BC set (Seiler and Crutzen, 1980) have resulted into a variety of methods and techniques

(e.g. optical, microscopic, thermal, chemical, spectroscopic and molecular markers) that

4
ACCEPTED MANUSCRIPT

determine different parts of the spectrum. However, beyond a purely analytical aspect, what is

really interesting about BC is the tool it offers to reconstruct environmental processes.

The term of BC includes combustion-derived materials (pyrogenic) such as char, charcoal and

PT
soot. Other highly aromatic materials that can also contribute to the pool of carbon measured by

RI
most BC measurement methods result from sedimentary metamorphism (petrogenic) and include

SC
coal, graphitized carbon and pure graphite (Veilleux et al., 2009). Within this series of

carbonaceous material, the least (char) and most (soot plus graphitic carbon) condensed forms of

NU
BC represent two end-members of the BC set, where the larger particle size (≥mm) and higher
MA
reactivity of the former contrast with the smaller particle size (submicron) and lower reactivity of

the latter (Harvey et al., 2012a-b). Pyrogenic in origin, char and soot are formed under different
D

conditions, with chars largely forming at lower temperature flaming and smoldering phases, and
TE

more dense soot resulting from the recondensation of small particles at high temperatures, high

pressure and temperature metamorphism. Petrogenic in origin, highly condensed graphitic


P
CE

carbon derives from the erosion of geologic material and is tightly associated with rocks and thus

tends to distribute following density gradients (Brandes et al., 2008; Dickens et al., 2006;
AC

Veilleux et al., 2009). Their differentiated origins and structures make the different BC forms

useful tracers of different sources and/or dynamics in the environment (Dickens et al., 2004;

Harvey et al., 2012a; Kuo et al., 2008a; Louchouarn et al., 2007; Masiello, 2004; Veilleux et al.,

2009).

BC in sedimentary environments constitutes an organic integrator of historical inputs from

biomass (vegetation fires or biofuel combustion) and industrial (transportation, factories, urban

activities) combustion in the drainage basin (Elmquist et al., 2007; Hunsinger et al., 2008; Kuo et

al., 2011). Solid residues (char) and aromatic condensates (soot) largely formed during biomass

5
ACCEPTED MANUSCRIPT

and fossil fuel combustion on continents, respectively, may eventually become buried in adjacent

continental shelves (Hunsinger et al., 2008; Kuo et al., 2011), together with petrogenic moieties

from the catchment area. BC has been speculated to be an important part of the marine cycle due

PT
to its relatively recalcitrant nature (Hedges et al., 2000; Hwang and Druffel, 2003), largely due to

RI
its condensed structure and in some cases physical protection (e.g. mineral coating) (Brodowski

et al., 2005a; Forbes et al., 2006; Knicker and Hatcher, 1997). Alternatively, some studies have

SC
shown quantitative (Middelburg et al., 1999; Stubbins et al., 2012; Worwoods et al., 2012;

NU
Zimmermann et al., 2012) and qualitative (Harvey et al., 2012b; Hockaday et al., 2006;

Worwoods et al., 2012) evidences of degradation (oxidation and dissolution) of BC under


MA
centennial or sub-centennial timescales. Still, its larger resistance to alteration relative to other

forms of organic matter (OM) (particularly under anoxic conditions) has made it a useful tracer
D
TE

for reconstructing combustion history and source inputs at different scales (e.g. De la Rosa et al.,

2011; Dickens et al., 2004a-b; Kuo et al., 2011; Hunsinger et al., 2008; Sánchez-García et al.,
P

2010a; Wolf et al., 2012).


CE

Mediterranean landmasses are interesting areas for studying coastal export of BC due to a
AC

combination of abundant vegetation fires and pollution burdens from Southern Europe. In the

SW of Spain, the Gulf of Cádiz (GoC) represents a sedimentary system under the influence of

heterogeneous sources of terrestrial pollution. Rivers that flow into the GoC through large

extensions of dry Mediterranean grasslands and a number of metropolitan areas receive

pyrogenic inputs from wild fires, agricultural practices, transportation, domestic heating, urban

and industrial activities, as well as erosional inputs from the basin drainage with presence of

Carbonaceous geology. Some of these rivers (i.e. Tinto and Odiel) are considered among the

most polluted fluvial systems in Western Europe (e.g. Elbaz-Poulichet et al., 1999). The local

6
ACCEPTED MANUSCRIPT

concurrence of diverse sources of BC (pyrogenic and petrogenic) makes this well-studied

sedimentary system an excellent environment to investigate the distribution of BC, in relation

with different polluting activities in the area.

PT
The goal of this study is to assess recent terrestrial inputs of BC to the GoC Continental Shelf

RI
exploiting bulk (thermal) and molecular (biomarkers) level techniques for discussing the relative

SC
contribution of different BC species. Measuring different, operationally-defined components of

the BC set provides a more comprehensive characterization of the sedimentary BC in the GoC.

NU
The two measurements of BC are combined with existing information on other organic moieties
MA
for discussing the main terrestrial sources to the GoC surface sediments. BC concentrations are

then combined with sediment properties to derive regional burial fluxes of the two pools of BC.
D

This qualitative and semi-quantitative discussion provides new insights in the structural
TE

composition of BC in marine sediments.


P

2. Materials and Methods


CE

2.1. Study area


AC

The study is conducted in the inner GoC Continental Shelf, between the estuaries of the

Guadiana and Odiel-Tinto Rivers (Fig. 1). This region is highly influenced by water run-off and

sediment load derived from the Guadiana, one of the major rivers in the Iberian Peninsula (742

km length and 67,000 km2 of drainage basin). Sediments deposited on the shelf are mainly

supplied and redistributed by this river (~58·104 m3/yr of suspended load and ~44·104 m3/yr of

bed load; Morales, 1995, 1997), the intense south-eastward littoral drift and the advection of the

North Atlantic Surficial Current (González et al., 2004). Owing to the dominant eastward littoral

drift (Fig. 1), sediments are also provided by the polluted Tinto and Odiel Rivers, although to a

7
ACCEPTED MANUSCRIPT

lower extent (e.g. Elbaz-Poulichet et al., 1999; Leblanc et al., 2000; Sánchez-García et al.,

2010b). Local rivers go across geologic formations from the Pleistocene-Pliocene Epoch

(conglomerates, gravel, sands, sandstone, silts and clays), and Carboniferous-Devonian Periods

PT
(volcanites, volcaniclastic rocks, shale, quartz, etc.), draining one of the largest sulphur deposits

RI
in the world (Iberian Pyritic Belt). The Tinto and Odiel Rivers receive important volumes of

point and non-point inputs from a number of industries established along the banks of the estuary

SC
since the mid-1900s (e.g. Grande et al., 2000) and discharge large concentrations of heavy metals

NU
and other inorganic pollutants (Borrego et al., 2002).
MA
The adjacent landmass stretches across large extensions of agricultural fields and Mediterranean

grasslands, where natural wildfires are common (mostly in summer) owing to specific climate
D

and vegetation characteristics. Fire department statistics of the neighboring urban centers
TE

estimate that ~5,000 ha of forest burn annually in >200 fires in the area (Zamora et al., 2010;

IGN, 2010). Scarce precipitations (200-800 mm) and high temperatures (annual mean of 14-18
P
CE

ºC; IGN, 2010; summer mean of 21-22 ºC for 1997-2010; INE, 2010), characteristic of the

temperate Mediterranean climate governing the Southern Iberian Peninsula, are key conditions
AC

favoring regular wildfire incidence. Sources of fossil fuel combustion in the study area include

transportation, domestic heating (mostly oil, natural gas and coal; European Commission, 2011),

and industrial activities. Located ~10 km inland from the Tinto-Odiel Estuary (Fig. 1), the city of

Huelva is the major metropolitan center in the region (~150,000 inhabitants; INE, 2010). The

important industrial complex developed along the banks of the estuary of the Tinto and Odiel

Rivers includes petroleum refineries, metal smelters, chemical plants, paper industry, fertilizer

factory, paper mills, etc., and is the major industrial center of the region. The input and

distribution of pyrogenic residues from biomass burning and fossil combustion to the area of

8
ACCEPTED MANUSCRIPT

study are thus controlled by a combination of diverse sources of emission followed by

atmospheric and riverborne transport. In addition, erosion of shale and other geological

Carboniferous materials could contribute to the petrogenic input of BC. The predominantly

PT
northerly winds in the region (Fig. 1) contribute to the dispersion of the continentally-produced

RI
pollutants into the GoC Continental Shelf.

SC
2.2. Sampling and geochemical description of the GoC sediments

NU
In 2001, sediment samples were collected from 15 stations between the Guadiana and Tinto-

Odiel Estuary in the GoC inner shelf (Fig. 1). The sediments distribution in the GoC illustrates
MA
the terrestrial influence in the coastal system along the Guadiana River plume, and allow for

discussing the relative influence of this river relative to the neighboring Tinto and Odiel Rivers.
D

The geochemical composition (both organic and inorganic) of the 15 marine sediments has been
TE

well described in previous studies (Sánchez-García et al., 2008, 2009, 2010b). The samples were
P

collected from the top 20 cm of sediments (accumulation of ~50 years; Table 2) using of a
CE

Shipeck dredge and were kept frozen until analysis. The sediments were dried at ≤40ºC, to

prevent the alteration of the organic fraction, crushed and homogenized at <0.25 mm prior to the
AC

chemical analysis.

The majority of the 15 sediments are composed of silt-clay clastic materials (fine fractions

accounting for 42-99% of the dry sediment mass), with minor sand contributions, typical of

sediment mixtures of continental-marine transition areas (Sánchez-García et al., 2010b).

Elemental, isotopic and molecular analyses of sedimentary OM indicated a NW-to-SE decline in

the proportion of terrestrial inputs following the dispersion of the Guadiana River plume into the

marine system (Sánchez-García et al., 2008; 2009). The total organic carbon (TOC) content in

the sediments ranged from 0.5 to 1.4% dry weight (dw) (Table 1). Atomic ratios of TOC to total

9
ACCEPTED MANUSCRIPT

nitrogen (C/N) varied between 6 and 23, and stable carbon isotope ratios (δ13C) ranged from -

22.6 to -25.9‰, suggesting a mixture of terrestrial and marine sources varying from 26-37% to

64-98% of terrestrial OM fraction, respectively (Sánchez-García et al., 2009).

PT
2.3. Isolation, quantification and characterization of black carbon

RI
Most of the methods developed for BC quantification in sediments rely on its resistance to

SC
degradation (thermal or chemical oxidation). While the use of benzene polycarboxylic acids

NU
(BPCAs) as molecular markers of BC are traditionally employed to measure BC in soils (e.g.

Brodowski et al., 2005b; Glaser et al., 1998; Glaser and Amelung, 2003) and in a few instances
MA
in dissolved OM (Dittmar, 2008; Ziolkowski and Druffel, 2010; Ziolkowski et al., 2011), a key

challenge remains for applying it to marine sediments (e.g. Ziolkowski and Druffel, 2009). The
D

main reason is because in this type of environmental matrix there are a number of non-BC
TE

moieties that may act as interfering material in different ways and extent. For instance, the
P

presence of certain sedimentary moieties (e.g. clays) has been discussed to probably cause the
CE

nitric acid oxidation of BC to proceed less efficiently (Ziolkowski and Druffel, 2009), thus

resulting in BPCA underestimation. In contrast, thermal methods quantifying oxidation resistant


AC

carbon may produce false positives, although is described to be more prone to underestimation

owing to mineral catalyzed oxidation of soot, or losses through sorption of the hydrophobic soot

moieties (Elmquist et al., 2004; Hammes et al., 2007). Inherent to their respective fundaments

(BPCA quantifying aromatic structures and thermal oxidation accounting for thermally-resistant

residues), both methods make analytical assumptions that may inevitably derive into either false

positives or underestimation. However, the two techniques contribute with different points of

view of the BC composition, with the BPCA method informing on BC quality through its

molecular marker pattern, and the thermal residue indicating the proportion of sedimentary

10
ACCEPTED MANUSCRIPT

carbon being resistant to thermal degradation. All together, combining the two analytical

approaches offers a great opportunity to achieve a more comprehensive characterization of the

BC fraction in marine sediments.

PT
2.3.1. Thermal oxidation: the GBC method

RI
This technique relies on the differential thermal resistance of BC and non-BC organic

SC
components in the sample, which is heated in an O2-rich atmosphere to volatilize-oxidize the

NU
most labile organic moieties, and measure the remaining residue as BC. A detailed description of

the thermal method applied here is presented elsewhere (Gélinas et al., 2001). Between 50-100 g
MA
of dry and ground sediments were weighed and subjected to a chemical step prior to the thermal

oxidation, to remove the most labile OM and minimize mineral-catalyzed condensation


D

reactions. During the chemical pre-treatment, a combination of inorganic and organic acids (HCl,
TE

HF and trifluoroacetic acid - TFA) is applied sequentially at different concentrations and


P

temperature conditions to de-mineralize the sediments matrix and remove hydrolysable OM


CE

(mostly proteins and carbohydrates). The OC fraction remaining after the subsequent thermal
AC

oxidation (here 375ºC/24h) is then quantified as Graphitic BC (GBC–including highly

condensed soot and graphitic carbon; Gélinas et al., 2001; Louchouarn et al., 2007) with an

elemental analyzer (Carlo Erba model 2400). Due to sample size limitations, the GBC approach

included only single measurements (n=1). Typical precision of this method for sediments (GBC

range of 0.05 to ~8%) is better than 2% for the thermal oxidation step alone and better than 10%

for the entire method including the chemical pre-treatment (Gélinas et al., 2001; Louchouarn et

al., 2007).

2.3.2. Use of molecular markers: the BPCA method

11
ACCEPTED MANUSCRIPT

The molecular markers technique cleaves the highly aromatic core of BC to produce BPCAs

from polycyclic or substituted aromatic centers during oxidative degradation (Hayatsu et al.,

1982). It yields structural information of the BC fraction by means of its molecular marker

PT
pattern (Hammes et al., 2008). The use of BPCAs as specific markers of BC in soils was initially

RI
proposed by Glaser et al. (1998) and lately revised by Brodowski et al. (2005b). Recently, the

method has been further revised and improved in terms of analytical handling and reproducibility

SC
(Schneider et al., 2011). Its application on marine sediments has been largely avoided due to the

NU
complex interferences derived from their heterogeneous composition and, in particular, the

presence of interfering organic moieties such as lignin, humic or proteinaceous compounds. To


MA
our knowledge, the use of BPCAs to determine BC in marine sediments has been uniquely tested

on the reference material SRM 1941b (Hammes et al., 2007; Ziolkowski and Druffel, 2009) and
D
TE

on a very small set of environmental samples (de la Rosa et al., 2011). This is the first time that

the BPCA method is applied to a larger set (n=15) of coastal sediments.


P
CE

The BPCA method consists of chemically treating the sample to form oxidation products that are

molecular markers of the aromatic core of BC. The samples are first digested in a microwave
AC

oven with TFA (4M, 170ºC, 30 min) to minimize the risk of acid-catalyzed BPCA formation

from polyvalent cations (Brodowski et al., 2005b). The different digestion temperature and times

applied here respect to the revised method by Brodowski et al. (2005b) (105 ºC and 4 hours),

respond to the different digestion system used here. Using microwave instead of heating bombs

allows for using shorter times of digestion, relative to Brodowski et al. The residue was filtrated

(glass fibre filter), rinsed with deionized water and dried at 30–40 ⁰C for at least 2 h.

Subsequently, the dry residue was oxidized with HNO3 (65%, 170ºC, 8h). The resulting solution

was diluted with deionized water to reduce the acid concentration and cleaned up by means of

12
ACCEPTED MANUSCRIPT

cation exchange (Dowex 50W X 8, 200-400 mesh). After freeze-drying and derivatization with

N,O-bis(trimethylsilyl)-trifluoroacetamide (Brodowski et al., 2005b; Glaser et al., 1998), the

aromatic acids were analyzed by capillary gas chromatography (GC) coupled to a mass

PT
spectrometer (MS) on an Agilent instrument (6890 N with MSD 5973 detector) equipped with a

RI
DB5ms capillary column (50 m x 0.2 mm i.d., 0.33 µm film thickness). Helium was used as

carrier gas at a flow rate of 1.2 mL/min. The injector and GC/MS interface were maintained at

SC
290 °C. The GC temperature was programmed from 70 °C (with a 2 min initial delay) to a first

NU
plateau of 200 °C (held for 5 min) at a ramp rate of 15 °C/min, then raised by 5 °C/min to a final

temperature of 300 °C (held 15 min).


MA
A series of external standards (hemimellitic, trimellitic, trimesic, pyromellitic, benzene
D

pentacarboxylic and mellitic acids) was used for calibration purposes. Phthalic acid was used as
TE

internal standard for recovery calculations. It was added to the samples at the beginning of the

chemical pretreatment (i.e. TFA, HCl and HF) and analyzed in parallel to the phthalic acid added
P
CE

to a preparation of external standards. The recovery of the BPCA production was calculated to be

between 80% and 102%. Salicylic acid and bipheny-2,2´-dicarboxylic acid were used as internal
AC

standards for correcting from losses during the cleaning procedure; the former was added prior to

the sample cleanup and the latter prior to derivatization. All methylated BPCAs were quantified

relative to the biphyenl-2,2´-dicarboxylic acid internal standard. No additional response factors

were applied, as all methylated BPCAs exhibited an equal response to detection. The sum of the

yields of BPCAs after nitric acid oxidation is a relative measure of the BC content in the sample

(BPCA-BC), after applying a correction factor of 2.27 (Glaser et al., 1998). This factor,

empirically determined by referencing to commercial charcoal, is used to account for the

incomplete conversion of BC to BPCAs, which depends on the degree of aromatic condensation

13
ACCEPTED MANUSCRIPT

and may differ considerably from sample to sample. Samples were analyzed in replicates (n=3)

and are reported as mean±SD.

2.4. Estimation of burial and sink fluxes of the two types of black carbon

PT
To estimate burial and sink fluxes of the different types of BC measured in the GoC sediments,

RI
several sediment properties must be taken into account. The BC burial flux (μg/cm2·yr) was

SC
estimated by:

NU
(Equation 1)

MA
ϕ
D

μg/cm2∙yr)
P TE

burial fluxes
CE
AC

The GoC Fburial is then applied to the corresponding areal extension (cm2) to calculate the BC

sink flux (Fsink) in metric tons per year (t/yr).

14
ACCEPTED MANUSCRIPT

The uncertainties reported for the BPCA burial fluxes represent the propagated standard

PT
deviation of the three replicates used to measure the BPCA-BC. The uncertainty for the two GoC

RI
BC sinks is estimated from the spatial variability of GBC and BPCA in the GoC (i.e. standard

deviation of the mean burial flux of GBC and BPCA for the 15 marine sediments).

SC
3. Results and discussion

NU
3.1. Measurements of black carbon by thermal oxidation (GBC)
MA
GBC was detected in all GoC sediments at concentrations ranging from 0.01-0.11% dw (mean of

0.04±0.03% dw; Table 1). These values are in the low range of BC concentrations estimated with
D
TE

different versions of the thermal method in coastal sediments from the Mexican margin (0.09-

0.26% dw; GBC method; Gélinas et al., 2001), Pan-Artic (~0.02-0.15% dw; CTO method;
P

Elmquist et al., 2008), North America (~0.01-0.69% dw; CTO method; Gustafsson and
CE

Gschwend, 1998), Northern Europe (0.04-1.77% dw; CTO method; Persson et al., 2002;
AC

Sánchez-García et al., 2010a), South Atlantic (0.04-0.17% dw; CTO method; Lohmann et al.,

2009), or Atlantic-Iberian margin (~0.05-0.16% dw; CTO method; Middelburg et al., 1999).

Differences relative to the present results may in some extent be due to the different thermal

method applied, since the CTO method (chemical-thermal oxidation; Gustafsson et al., 2001)

used in most of those studies directly subjected the sample to thermal oxidation without previous

elimination of potentially interfering organic material. Here, applying a multiacid treatment (i.e.

HCl, HF and TFA) prior to the thermal oxidation resulted in lower concentrations than only

thermal heating by factors of 1-48 (data not shown). Gélinas and coworkers (2001) reported

15
ACCEPTED MANUSCRIPT

similar factors (GBC-BC being 5-72 times lower than CTO-BC) in seven marine sediments from

the Mexican margin and Canada. Similarly, Schmidt et al. (2001) reported GBC contents about

50 times lower than those directly measured through thermal oxidation in eight Australian soils.

PT
Because environmental samples such as sediments and soil contain a milliard of organic moieties

RI
(proteins, carbohydrates, lignin, humic substances, etc.), the charring of the non-BC

carbonaceous material present in the samples (e.g. Gélinas et al., 2001; Hammes et al., 2007)

SC
may cause considerable differences by applying one or another thermal version.

NU
In the GoC Continental Shelf, GBC showed a relatively homogeneous spatial distribution, with
MA
only four samples containing concentrations higher than the mean value of 0.04% dw (Table 1).

The larger content of GBC in S1-S3 may be interpreted by their proximity to the Guadiana River
D

mouth (Fig. 2), implying that this could be the major source of GBC in the area. Since GBC is
TE

mostly petrogenic in nature (e.g. Dickens et al., 2004a-b), tightly associated with rocks (Brandes

et al., 2008), and follows a distribution of dense mineral grains, it is expected to find higher
P
CE

concentrations in the proximity of the river mouth. The granulometry of the GoC sediments may

help to understand this spatial pattern, with the largest particles (sand-sized or large silt-sized)
AC

generally containing higher values of GBC near the Guadiana mouth (Table 1). Recently, the use

of a two end-member isotopic mass balance revealed that the abundance of terrigenous organic

carbon (terrOC) in the 15 GoC sediments decreased as a function of distance from the Guadiana

mouth (Sánchez-García et al., 2009). This spatial pattern was explained by hydrodynamic

sorting, where the terrOC-rich coarse particles, characterized by a fresher OM fraction, were

deposited in the proximity of the river mouth, and the terrOC-depleted fine-soil particulates were

selectively transported over longer distances. Elemental, isotopic and molecular signatures

demonstrated the higher degradation state of the terrOC-poor particles selectively transported out

16
ACCEPTED MANUSCRIPT

on the GoC shelf (Sánchez-García et al., 2008, 2009). This spatial pattern generally agrees with

the offshore decrease of GBC observed within the first 10 km along the SE-ward Guadiana

plume. Further out though, the amount of GBC drops to lower values (Fig. 2), and remain more

PT
or less constant even at distances ~49 km from the Guadiana mouth. The nearly constant

RI
BC/TOC percentages (~2-4%) observed in all stations except S-1 to S-3 and S-9 suggest the

presence of an extra input of BC other than the Guadiana River contributing with a steady

SC
concentration of BC in the sediments. Contribution of light, airborne particles of BC such as

NU
soot, less affected by the hydrodynamic sorting, could explain the homogeneous distribution

observed over the direct discharge of the Guadiana. Apparently, the GBC fraction in the GoC
MA
sediments is composed by a mixture of petrogenic BC mainly delivered by the Guadiana River

and pyrogenic soot most likely supplied by local northerly winds.


D
TE

3.2. Measurements of black carbon using molecular markers (BPCAs)


P

The use of BPCAs yielded sedimentary BC concentrations in the GoC ranging from 0.01 to
CE

0.23% dw (mean of 0.09±0.08% dw) (Table 1). These results are based on the use of the

empirical correction factor of 2.27 (Glaser et al., 1998) as the best estimate of BPCA-BC for the
AC

GoC sediments, as this factor has been previously cautioned to possibly underestimate the real

amount of BC on soil samples (Brodowski et al., 2005b). Whereas one has to be aware that this

calculation is only an approximation, the use of 2.27 is recommended for determining absolute

estimates of BC on environmental samples, since it is the minimum conversion factor assessed

so far (Brodowski et al., 2005b). We interpret these concentrations as conservative due to the

inefficient oxidation of BC with the nitric acid that may occur in presence of certain components

such as clay (Ziolkowski and Druffel, 2009). Although this was not specifically quantified here,

17
ACCEPTED MANUSCRIPT

the lower concentration of BPCA-BC observed on samples with larger content of clay (Table 1)

may be indicative of such matrix effect.

The only BPCA measurements on marine sediments available for comparison are those of the

PT
reference sediment NIST 1941b (0.44±0.04% dw, Ziolkowski and Druffel, 2009; and 0.05-

RI
0.30% dw, Hammes et al., 2007), and four samples collected from the same study region (0.03-

SC
0.15% dw; de la Rosa et al., 2011). In our study, higher concentrations were generally observed

near shore (except for S1, S3 and S4), within ~16 km off the Guadiana River mouth, where six

NU
out of nine samples showed BPCA-BC values >0.13% dw (Fig. 2). Interestingly, the near-shore
MA
samples showing low BPCA-BC values are those of highest GBC content (S1 and S3),

documenting the different nature of both materials. Overall, the BPCA method is considered to
D

be better suited for detecting pyrogenic chars (Brodowski et al., 2005b; Hammes et al., 2007),
TE

solid BC residues still retaining some structural information and functionalities of the original

vegetation (e.g. Goldberg, 1985). Because of their larger size (relative to soot or graphitic
P
CE

carbon), char particles are prone to settle down closer to their delivery (Kuo et al., 2008a-b), so

the generally higher values of BPCA-BC found in the vicinity of the Guadiana estuary suggests
AC

this river as the main supplier of this type of BC in the GoC Continental Shelf, as for GBC. In

contrast to the latter, BPCA-BC is found in high concentrations as far as ~16 km from the

Guadiana mouth.

The relative distribution of BPCAs provides information on structural composition of the

original BC mainly through the estimate of the proportion of condensed aromatic structures.

Although some BC materials can form BPCAs with two carboxylic acids, only the compounds

with three to six acid groups were quantified in this work to avoid the potential artifact of

incorporating non-BC derived BPCA (e.g. from lignin or other biologic materials) into the total

18
ACCEPTED MANUSCRIPT

BC estimate (Ziolkowski and Druffel, 2009). The individual BPCA distribution largely varied

between the samples (Fig. 3). The tri-, tetra- and penta-substituted acids (B3CA, B4CA and

B5CA, respectively) were generally present in all samples (Fig. 3a-c). In contrast, the hexa-

PT
substituted acid (B6CA) was found in only 2/3 of the samples and was most prominent in the

RI
central zone of the plume (Fig. 3d). The number of carboxylic acid groups on each BPCA is a

function of the number of aromatic carbons attached to it prior oxidation. A higher proportion of

SC
B6CA suggests larger content of the most condensed BC, where a dense structure of aromatic

NU
rings leaves little space for substituted groups (Ziolkowski and Druffel, 2009, 2010). This

produces lower average acid number “N” (Ziolkowski and Druffel, 2010) in samples S1, S3, S4,
MA
S7, and S8 (Table 1), pointing to more abundant proportion of char-BC in these samples (Fig. 4),

since charred wood-derived BPCAs typically show average N values <4.5 (Ziolkowski, 2009).
D
TE

The virtual absence of B6CA in nearshore samples such as S1, S3 and S4 is consistent with the

larger size of the less condensed particles that tend to settle closer to the river mouth (Sánchez-
P

García et al., 2009). In contrast, average acid numbers between 4.5 and 5.5 in other samples
CE

suggest a mixture of charred and more condensed BPCAs (Ziolkowski, 2009; Schneider et al.,
AC

2011), indicating the presence of some soot particles.

3.3. Differences between molecular markers and thermally derived black carbon

Reported environmental measurements of BC show a large range of concentrations in marine

sediments. For example in ocean sediments, BC has been described to represent an important

fraction of the TOC content that varies from 2 to ~60% of TOC in continental shelf sediments

(e.g. Cornelissen et al., 2005; Gustafsson and Gschwend, 1997; Lim and Cachier, 1996;

Middelburg et al., 1999; Sánchez García et al., 2010a) and from 15 to >50% of TOC in deep

ocean sediments (e.g. Masiello and Druffel, 1998; Verardo and Ruddiman, 1996). However, the

19
ACCEPTED MANUSCRIPT

choice of different techniques to determine BC in environmental matrices may result in BC

concentrations differing by large factors (Schmidt et al., 2001). In the GoC, the analysis of BC

upon use of the BPCA versus GBC method produced differences by a factor of up to 9 (Table 1).

PT
Overall, the molecular markers produced higher BC values (BPCA-BC: 1-34% of TOC, mean of

RI
11±10%) than the thermal method (GBC: 2-15% of TOC; mean of 5±3%). In a comparative

analysis of BC in soils, Schmidt et al. (2001) similarly reported larger BPCA-BC concentrations

SC
than those derived from using the GBC thermal method. In agreement with the variations

NU
observed here, the differences between the soil samples were also larger when the BPCA method

was used. In the recent multilab intercomparison by Hammes et al. (2007), BC measurements on
MA
the reference marine sediment NIST 1941b also varied considerably (from 2.0-8.6% of TOC to

12.8-22.6% of TOC) when the BPCA or thermal oxidation (CTO method) was applied,
D
TE

respectively. In the Hammes et al. study, the highly condensed composition of BC was argued to

explain the higher concentration of thermal residue relative to BPCA-BC. It was also indicated
P

that the use of a washing step to remove labile OM (as we did here) consistently gave lower
CE

results than when the samples were directly heated. In both studies BC structure and
AC

condensation were considered key factors in the observed methodological differences, with the

thermal oxidation (as CTO method) providing better differentiation between soot and char.

Different factors may have contributed to the distributional differences observed in the GoC

sediments, as well as the different analytical windows covered by the techniques used (Masiello,

2004). Although both methods (GBC and BPCA) isolate carbonaceous material of aromatic

nature, they rely on analytical techniques inherently designed to detect different forms of BC.

The GBC method is best suited for quantifying the most condensed forms of BC (i.e. soot and

graphitic carbon) (Gélinas et al., 2001), irrespective of their source (pyrogenic or petrogenic). In

20
ACCEPTED MANUSCRIPT

contrast, the BPCA method targets less condensed BC (i.e. char and charcoal) (Brodowski et al.,

2005a-b), although uncertainties remain regarding its specificity and efficiency (Hammes et al.,

2007). Despites its potential to measure aromatic structures independently of their origin, BPCA-

PT
BC has been considered to be more effective in measuring pyrogenic structures ranging from

RI
slightly charred biomass to slightly condensed particles of soot (Masiello, 2004). Obtaining

different results by using GBC or BPCA is thus expected since both methods largely target

SC
different sources and carbonaceous structures. In the GoC, the structural diversity of BC may

NU
explain the differential distribution along the Guadiana River plume, where coarser grain-

associated, petrogenic BC and larger-sized pyrogenic char-BC remain close to the river mouth
MA
while smaller-sized pyrogenic particles (condensed soot and lighter char) travel further out on

the shelf. The degree of similarity between both estimates in the out-region of the plume may be
D
TE

interpreted in the light of the common soot measured by both methods. While BPCA basically

measures pyrogenic BC (both soot and char), GBC detects highly condensed BC (both
P

petrogenic graphitic carbon and pyrogenic soot). The differential specificity of the methods to
CE

detect chars (BPCA) or graphitic carbon (GBC) contrasts to the common ability of both to
AC

measure soot, irrespective of the efficiency they attain. In fact, part of the BPCA-BC in the GoC

sediments seem to be composed by slightly condensed particles of soot, as suggested by the

observed N values (Fig. 4). This measurement may overlap with the soot detected by the GBC

technique, explaining in some extent the similar distribution of GBC and BPCA-BC observed

between distances 20-50 km from the Guadiana mouth. Atmospheric deposition emerges as a

likely way of supply of the soot from the constant distribution observed for both measurements.

All in all, the differences observed in the spatial distributions of BPCA-BC and GBC in the GoC

may be explained by a combination of the different sources (pyrogenic versus petrogenic),

21
ACCEPTED MANUSCRIPT

distribution pathways (riverine versus aerial) and carbonaceous structure (more or less

condensed) of the two BC measurements.

However, beyond a simple estimation of its concentration, the combination of both results

PT
provides useful insights on the origin and composition of sedimentary BC. Assessing the

RI
presence of different types of BC in the sediments allows for some reconstruction of the regional

SC
sources of BC in the GoC shelf, where the Guadiana River acts as a major source of pyrogenic

char and petrogenic BC, and atmospheric deposition plays a role in delivering soot.

NU
3.4. Burial flux of black carbon on the GoC inner shelf
MA
Estimates of burial fluxes of BC at regional to local scales is essential to better constrain the

global BC budget in one of the most important interfaces of BC in the biosphere (i.e. coastal
D

sediments). Since BC in the environment is understood as a catchall of combustion (Hedges et


TE

al., 2000) and petrogenic (Dickens et al., 2004a-b) products, and because no method detects the
P

entire set of BC (Masiello, 2004), combining methods to assess complementary portions of the
CE

set is important to achieve a comprehensive estimate of the sedimentary BC budget.


AC

Using the measured BC and density values, deduced sedimentation rates (see Sánchez-García et

al., 2010a for calculation details), and available information on porosity (Jönsson et al., 2003),

we estimated the burial flux of the two BC measurements in the GoC (Table 2). The burial fluxes

of BPCA-BC in the GoC sediments were found to vary between 21 and 633 μg/cm2∙yr, with

generally smaller fluxes for GBC (42-285 μg/cm2∙yr) (Table 2). Overall, these burial fluxes are

comparable to those obtained with similar thermal methods for other continental shelf regimes.

For instance, BC burial flux in the Gulf of Maine Proper (inner shelf) have been reported to

range 86-194 μg/cm2∙yr (CTO method; Gustafsson and Gschwend, 1998), in the Washington

22
ACCEPTED MANUSCRIPT

coasts (inner and outer shelf) 1.9-3.1 μg/cm2∙yr (GBC method; Dickens et al., 2004a), near

metropolitan areas in the NE USA (inner shelf) 470-2300 μg/cm2∙yr (CTO method; Gustafsson

and Gschwend, 1998), and in the Gulf of Panama (inner shelf) 100-300 μg/cm2∙yr (KOH/H2O2

PT
digestion method; Suman, 1986). The GoC burial fluxes are in the lower range of those

RI
measured close to urban nuclei. Condensed fractions of BC from pyrogenic and petrogenic

sources accumulate in the GoC in similar fluxes as those measured in inner shelves from North

SC
and Central America. Unfortunately, there is no report of BC burial fluxes from BPCA

NU
measurements.
MA
The burial fluxes of BC on the inner shelf of the GoC, comprised between the San Vicente Cape

and the Gibraltar Rock and down to 200 m of bathymetry (area of 10,788 km2, ARGIS-

calculated), amounts to annual sink fluxes of ~26±23·103 t of BPCA-BC and ~13±8.6·103 t of


D
TE

GBC (245±205 and 120±76 μg/cm2·yr, respectively). Extrapolating from 15 samples largely

from the Guadiana River plume to the entire GoC inner shelf made us to consider these fluxes as
P
CE

maximum values due to the direct burden of the river. Still, we consider the estimates as valid

because of the steady distribution of BC observed over the ~15 km out of the Guadiana Mouth,
AC

which indicates that the high values potentially biasing the resulting fluxes toward some

overestimation are just restricted to certain samples close the Guadiana Mouth. The Fsink derived

from the GBC method is in the range of those estimated by the CTO method in coastal regimes

of comparable extension in Scandinavia, such as the Bothnian Bay (~8·103 t of BC/yr, n=9; 415

μg/cm2·yr), Borholm Sea (~20·103 t of BC/yr, n=8; 1,544 μg/cm2·yr) or Gotland Sea (~178·103 t

of BC/yr, n=19; 2,136 μg/cm2·yr). The higher thermally derived-burial fluxes in those regions

relative to the present study corresponded to direct BC accumulation zones only (Sánchez-García

et al., 2012). Sink fluxes of relatively similar magnitude were also estimated by the CTO method

23
ACCEPTED MANUSCRIPT

for the New England Shelf (400·103 to 800·103 t of BC/yr, 100-200 μg/cm2·yr; Gustafsson and

Gschwend, 1998), or South Atlantic shelves and basin (480-700·103 t of BC/yr, 0.5-7.8

μg/cm2·yr; Lohmann et al., 2009), whereas larger fluxes were derived from a finer analysis based

PT
on the effective accumulation area in the Northern European Shelf (~1,100·103 t of BC/yr, 1,100

μg/cm2·yr; Sánchez-García et al., 2012). Although a straight comparison between those fluxes

RI
(CTO-based) and the present one (GBC-based) would not be fair because of the different thermal

SC
versions employed, it is the same order of magnitude obtained by both measurements what really

NU
matters here.
MA
Aware of the technical limitations and uncertainties of each of the two methods used, we do not

intend the numbers presented here to represent two clear-cut pools of BC in the GoC, but rather
D

an indication of the BC ranges in which both burial fluxes may be. Assuming these limitations
TE

and the fact that some overestimation from the soot overlapping may occur, we could attempt to

make a first estimation of the total sink flux of BC for the ARGIS-defined region. Considering
P
CE

that the two analytical methods measures two different types of BC (from slightly to highly

condensed pyrogenic char and soot versus highly condensed pyrogenic and petrogenic GBC), we
AC

estimate the total sink flux of BC (i.e. the sum of BPCA-BC and GBC) to be ~39±24·103 t of

BC/yr. This approach may be regarded as a conservative value of the total sink flux of BC in the

GoC, as the lighter and most reactive moieties of the BC spectrum may have been

underestimated (Masiello, 2004). Alternatively, soot may have been doubly counted by both the

GBC and BPCA methods, so partially offsetting that underestimation. Regardless of

improvement requirement, this estimate provides a more realistic number of the sedimentary sink

of BC in the GoC, and illustrates the benefits of accounting for different parts of the BC set to

obtain more accurate, regional BC budgets.

24
ACCEPTED MANUSCRIPT

4. Conclusion

This study provides a comprehensive characterization of the BC in sediments from the GoC

inner shelf, on the basis of two analytical techniques. The use of molecular markers (i.e. BPCAs)

PT
is applied to a set of 15 marine sediments to determine their BC content, in comparison to a more

RI
traditional approach for this type of environmental samples (i.e. thermal oxidation). The

SC
analytical differences (factors of 0.1-9.2 between BPCA and GBC estimates) derived from

inherent limitations and the different part of the BC spectrum detected by each technique were

NU
similar to those reported in previous inter-comparisons in marine sediment standards and soils.
MA
Measurements of BPCA-BC and GBC, which respectively trace char and soot plus petrogenic

graphitic carbon, showed distribution patterns consistent to those described for other terrestrial
D

components in the light of hydrodynamic sorting and two major terrestrial sources in the region.
TE

The Guadiana River integrates a pyrogenic-petrogenic signal of multiple activities developed

along its 742 km–course and delivers BC of dominant charcoal (i.e. fires) and, in minor extent,
P
CE

petrogenic signature, whereas wind emerges as the most likely transport way for soot. While a

one-method determination would have partially informed about the amount of one specific type
AC

of BC contained in the GoC sediments, the application of two complementary analytical

approaches provide useful insights on the origin and nature of the sedimentary BC. Whilst none

of the analytical measurements (GBC versus BPCA-BC) should be independently used as

absolute quantifications of BC in the GoC sediments, they combined provide a more realistic

range of regional sink fluxes. The ultimate value of the two BC quantification methods is not

how they compare to one another, but what useful information on the BC nature they may

provide. Since defining a clear-cut boundary between the two types of BC is difficult to attain

with the present operational definitions, combining different techniques seems crucial for a more

25
ACCEPTED MANUSCRIPT

comprehensive constraint of the environmental pools of BC in coastal systems. This combining

exercise constitutes a valuable first approach to estimate more realistically how much BC

accumulates in the continental shelves, and contributes to filling the gaps in the terrestrial

PT
compartment of the global BC budget.

RI
SC
Acknowledgments

NU
MA
D
P TE
CE
AC

26
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
TE
P
CE
AC

27
ACCEPTED MANUSCRIPT

References

Borrego, J., Morales, J.A., de la Torre, M.L., Grande, J.A. Geochemical characteristic of heavy

metal pollution in surface sediments of the Tinto and Odiel river estuary (southwestern Spain).

PT
Environ. Geol. 2002; 41: 785-796.

RI
Boudreau, B.P. The diffusive tortuosity of fine-grained unlithified sediments. Geochim.

SC
Cosmochim. Acta 1996; 60, 3139-3142.

Brandes, J.A., Cody, G.D., Rumble, D., Haberstroh, P., Wirick, S., Gélinas, Y. Carbon K-edges

NU
XANES spectroscopy of natural graphite, Carbon 2008; 46, 1424-1434.
MA
Brodowski, S., Amelung, W., Haumaier, L., Abetz, C., Zech, W. Morphological and chemical

properties of black carbon in physical soil fractions as revealed by scanning electron microscopy
D

and energy dispersive X-ray spectroscopy. Geoderma 2005a; 128: 116–129.


TE

Brodowski, S., Rodionov, A., Haumaier, L., Glaser, B., Amelung, W. Revised black carbon
P

assessment using benzene polycarboxylic acids. Org. Geochem. 2005b; 36: 1299-1310.
CE

Cornelissen, G., Gustafsson, Ö., Bucheli, T.D., Jonker, M.T.O., Koelmans, A.A., Van Noort,
AC

P.C.M. Critical review: extensive sorption of organic compounds to black carbon, coal and

kerogen in sediments and soils: mechanisms and consequences for distribution, bioaccumulation

and biodegradation. Environ. Sci. Technol. 2005; 39: 6881–6895.

De la Rosa, J.M., Sánchez-García, L., de Andrés, J.R., González-Vila, F.J., Knicker, H.

Contribution of Black Carbon in recent sediments of the Gulf of Cádiz. Applicability of different

quantification methodologies. Quatern. Int. 2011; 243, 264-272.

Dickens, A.F., Gélinas, Y., Masiello, C.A., Wakeham, S., Hedges J.I. Reburial of fossil organic

carbon in marine sediments. Nature 2004a; 427: 33-339.

28
ACCEPTED MANUSCRIPT

Dickens, A.F., Gélinas, Y., Hedges J.I. Physical separation of combustion and rock sources of

graphitic black carbon in sediments. Mar. Chem. 2004b; 92: 215-223.

Dickens, A.F., Baldock, J.A., Smernik, R.J., Wakeham, S.G., Arnason, T.S., Gélinas, Y.,

PT
13
Hedges, J.I. Solid-State C NMR analysis of size and density fractions of marine sediments:

RI
Insight into organic carbon sources and preservation mechanisms. Geochim. Cosmochim. Acta

SC
2006; 70: 666-686.

Dittmar, T. The molecular-level determination of black carbon in marine dissolved organic

NU
matter. Org. Geochem. 2008; 39, 396-407.MA
Dittmar, T., B. P. Koch, N. Hertkorn, G. Kattner. A simple and efficient method for the solid-

phase extraction of dissolved organic matter (SPE-DOM) from seawater. Limnol. Oceanogr.-
D

Meth. 2008; 6, 230-235.


P TE
CE

Elbaz-Poulichet, F.,Morley,N.H., Cruzado, A.,Velásquez, Z., Achterberg, E.P., Braungardt, C.B.


AC

Tracemetal and nutrient distribution in anextremely low pH (2.5) river-estuarine system, the Ría

of Huelva (south-west Spain). Sci. Total Environ. 1999; 227: 73-83.

Elmquist, M., Gustafsson, Ö., Andersson, P. Quantification of sedimentary black carbon using

the chemothermal oxidation method: An evaluation of ex situ pretreatments and standard

additions approaches. Limnol. Oceanogr.-Meth. 2004; 2; 417-427.

Elmquist, M., Z. Zencak, Ö. Gustafsson. A 700 year sediment record of black carbon and

polycyclic aromatic hydrocarbons near the EMEP air monitoring station in Aspvreten, Sweden.

Environ. Sci. Technol. 2007; 41, 6926-6932.

29
ACCEPTED MANUSCRIPT

Elmquist, M., I. Semiletov, L.D. Guo, Ö. Gustafsson. Pan-Arctic patterns in black carbon

sources and fluvial discharges deduced from radiocarbon and PAH source apportionment

markers in estuarine surface sediments. Glob. Biogeochem. Cy., 2008; 22: GB2018, doi:10.1029/

PT
8692007GB002994.

RI
SC
European Commission, 2011.

NU
Flores-Cervantes, D.X., Plata, D. L., MacFarlane, J.K., Reddy, C.M., Gschwend, P.M. Black
MA
carbon in marine particulate organic carbon: inputs and cycling of highly recalcitrant organic

carbon in the Gulf of Maine. Mar. Chem. 2009; 113: 172-181.


D

Forbes, M.S., Raison, R.J., Skjemstad, J.O. Formation, transformation and transport of black
TE

carbon (charcoal) in terrestrial and aquatic ecosystems. Sci. Total Environ. 2006; 370: 190-206.
P

Gélinas, Y., Prentice, K.M., Baldock, J.A., Hedges, J.I. An improved thermal oxidation method
CE

for the quantification of soot/graphitic black carbon in sediments and soils. Environ. Sci.
AC

Technol. 2001; 35: 3519-3525.

Glaser, B., Haumaier, L., Guggenberger, G., Zech, W. Black carbon in soils: the use of

bezenecarboxylic acids as specific markers. Org. Geochem. 1998; 29: 811-819.

Glaser, B., Amelung, W. Pyrogenic carbon in native grassland soils along a climosequence in

North America. Global Biogeochem. Cycles 2003; 17, doi:10.1029/2002GB002019.

Goldberg, E.D. Black carbon in the environment. New York: John Wiley; 1985.

30
ACCEPTED MANUSCRIPT

González, R., Dias, J.M.A., Lobo, F., Mendes, I. Sedimentological and paleoenvironmental

characterization of transgressive sediments on the Guadiana Shelf (Northern Gulf of Cádiz, SW

Iberia). Quatern. Int. 2004; 120: 133-144.

PT
Grande, J.A., Borrego, J., Morales, J.A. A study of heavy metal pollution in the Tinto–Odiel

RI
estuary in southwestern Spain using factor analysis. Environ. Geol..2000; 39:1095-1101

SC
Gustafsson, Ö., Gschwend. P.M. Soot as a strong partition medium for polycyclic aromatic

hydrocarbons. In: Eganhouse, R.P., editors. Molecular Markers in the Environmental

NU
Geochemistry. Washington, D.C.: Am. Chem. Soc.; 1997. p. 365-381.
MA
Gustafsson, Ö., Gschwend. P.M. The flux of black carbon to surface sediments on the New

England continental shelf. Geochim. Cosmochim. Acta 1998; 62: 465-472.


D

Gustafsson, Ö ., Bucheli, T. D., Kukulska, Z., Andersson, M., Largeau, C., Rouzaud, J.-N., et al.
TE

Evaluation of a protocol for the quantification of black carbon in sediments. Glob. Biogeochem.
P

Cy., 2001; 15: 881-890.


CE

Hammes, K., Schmidt, M.W.I., Smernik, R.J., Currie, L.A., Ball, W.P., Nguyen, T.H., et al.
AC

Comparative analyses of black carbon reference materials from soil, water, sediment and

atmosphere and environmental implications. Global Biogeochem. Cy. 2007; 21, GB3016,

doi:10.1029/2006GB002914.

Hammes, K., Torn, M.S., Lapenas, A.G., Schmidt, M.W.I. Centennial black carbon turnover

observed in a Russian steppe soil. Biogeosciences 2008; 5, 1339-1350.

Harvey, O.R., Kuo, L-J., Zimmerman, A.R., Louchouarn, P., Amonette, J.E., Herbert, B.E. An

index-based approach to assessing recalcitrance and soil carbon sequestration potential of

engineered black carbons (biochars). Environ. Sci. Technol. 2012a; 46: 1415-1421.

31
ACCEPTED MANUSCRIPT

Harvey, O.R., Herbert, B.E., Kuo, L-J., Louchouarn, P. Two-dimensional perturbation-based

correlation infrared spectroscopy (2D-PCIS) reveals mechanisms by which biochars develop

surface charge and resistance to environmental degradation. Environ. Sci. Technol. 2012b (In

PT
Review).

RI
Hayatsu, R., Scott, R.G., Winans, R.E. Oxidation of coal. In: Trahanovsky, W.S., editors.

SC
Oxidation in Organic Gemistry, Part D. New York: Academic Press; 1982. p. 279-354.

Hedges, J.I., Eglinton, G., Hatcher, P.G., Kirchman, D.L., Arnosti, C., Derenne, S., et al. The

NU
moleculary-uncharaterized component of nonliving organic matter in natural environments. Org.
MA
Geochem. 2000; 31: 945-958.
D
P TE
CE

Hunsinger, G.B., Mitra, S., Warrick, J.A., Alexander, C.R. Oceanic loading of wildfire-derived
AC

organic compounds from a small mountainous river. J. Geophys. Res. 2008; 113.

doi:10.1029/2007JG000476.

Instituto Geográfico Nacional (Spanish Ministry of Public Works). Maps of climate and

environmental risks. Available from: http://www.ign.es/espmap/mambiente_bach.htm. 2010.

Instituto Nacional de Estadística de España. Climatology. Available from: http://www.ine.es.

2010.

32
ACCEPTED MANUSCRIPT

IPCC, Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to

the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge:

Cambridge Univ. Press; 2007.

PT
RI
SC
Jönsson, A., Gustafsson, Ö., Axelman, J., Sundberg, H. Global accounting of PCBs in the

Continental Shelf Sediments. Environ. Sci. Technol. 2003; 37, 245-255.

NU
MA
Kuhlbusch, T.A.J. Black carbon and the carbon cycle. Science 1998; 280: 1903-1904.
D
TE

Kuo, L-J, Louchouarn, P., Herbert, B.E. Fate of CuO-derived lignin oxidation products during

plant combustion: Application to the evaluation of char inputs to soil organic matter. Org.
P

Geochem. 2008a; 39, 1522-1536.


CE

Kuo, L-J, Herbert, B.E., Louchouarn, P. Can levoglucosan be used to characterize and quantify
AC

char/charcoal black carbon in environmental media? Org. Geochem. 2008b; 39, 1466-1478.

Kuo, L-J, Louchouarn, P., Herbert, B.E. Influence of combustion conditions on yields of solvent-

extractable anhydrosugars and lignin phenols in chars: Implications for characterizations of

biomass combustion residues. Chemosphere 2011, 85: 797-805.

Leblanc, M., Morales, J.M., Borrego, J., Elbaz-Poulichert, F. 4,500-Years-old mining pollution

in southwestern Spain: long-term implications for modern mining pollution. Econ. Geol. 2000;

95:655-662.

33
ACCEPTED MANUSCRIPT

Lim, B., H. Cachier. Determination of black carbon by chemical oxidation and thermal treatment

in recent marine and lake sediments and Cretaceous-Tertiary clays. Chem. Geol. 1996; 131: 143-

154.

PT
Lohmann, R., Bollinger, K., Cantwell, M., Feichter, J., Fischer-Bruns, I., Zabel, M. Fluxes of

RI
soot black carbon to South Atlantic sediments. Global Biogeochem. Cy. 2009; 23, GB1015,

SC
doi:10.1029/2008GB003253.

Louchouarn P., Chillrud, S., Houel, S., Yan, B., Chaky, D., Rumpel, C. et al. Elemental and

NU
isotopic evidence of soot- and char-derived black carbon inputs to New York City’s atmosphere
MA
during the 20th Century. Environ. Sci. Technol. 2007; 41, 82-87.

Masiello, C.A., Druffel, E.R.M. Black carbon in deep-sea sediments. Science, 1998; 280: 1911-
D

1913.
TE

Masiello, C.A., Druffel, E.R.M. Organic and black carbon 13C and 14C through the Santa Monica
P

Basin oxic-anoxic transition. Geophys. Res. Lett. 2003; 30: 4, doi:10.1029/2002GL015050.


CE

Masiello, C.A. New directions in black carbon organic geochemistry. Mar. Chem. 2004; 92: 201-
AC

213.

Middelburg, J.J., Soetaert, K., Herman, P.M.J. Empirical relationships for use in global

diagenetic models. Deep Sea Res. 1997; Part I, 44: 327-344.

Middelburg, J.J., Nieuwenhuize, J., vanBruegel, P. Black carbon in marine sediments. Mar.

Chem. 1999; 65: 245-252.

Morales, J.A. Sedimentology of the Guadiana Estuary (S.W. Spain-Portugal). Ph.D. Thesis,

University of Sevilla, Sevilla. Serv. Publ. Univ. Huelva, 1995. 274 pp.

34
ACCEPTED MANUSCRIPT

Morales, J.A. Evolution and facies architecture of the mesotidal Guadiana River delta (S.W.

Spain – Portugal). Mar. Geol. 1997; 138: 127-148.

Persson, N.J., Gustafsson, Ö., Bucheli, T.D., Ishaq, R., Naes, K., Broman, D. Soot-carbon

PT
influenced distribution of PCDD/Fs in the marine environment of the Grenlandsfjords, Norway.

RI
Environ. Sci. Technol. 2002; 36, 4968-4974.

SC
Sánchez-García, L., de Andrés, J.R., Martín-Rubí, J.A., Gónzalez-Vila, F.J., Polvillo, O. Use of

lipid biomarker patterns as proxies of environmental variability in coastal sedimentary record

NU
from the Gulf of Cádiz (SW Spain). Org. Geochem. 2008; 39: 958-964.
MA
Sánchez-García, L., de Andrés, J.R., Martín-Rubí, J.A., Louchouarn, P. Diagenetic state and

source characterization of marine sediments from the inner continental shelf of the Gulf of Cádiz
D

(SW Spain), constrained by terrigenous biomarkers. Org. Geochem. 2009; 40: 184-194.
TE

Sánchez-García, L., Cato, I., Gustafsson, Ö. A large-scale evaluation of the black carbon
P

influence on the PAH distribution in recent sediments from the Swedish continental shelf. Mar.
CE

Chem. 2010a; 119: 44-51.


AC

Sánchez-García, L., de Andrés, J.R., Martín-Rubí, J.A. Geochemical signature in off-shore

sediments from the Gulf of Cádiz inner shelf. Provenance and spatial variability. J. Mar. Sci.

2010b; 80:191-202.

Sánchez-García, Cato, I., Gustafsson, Ö. The sequestration sink of black carbon in the Northern

European Shelf sediments. Glob. Biogeoch. Cy. 2012; 26: GB1001, 12 pp.,

doi:10.1029/2010GB003956

Schmidt, M.W.I., Skjemstad, J.O., Czimczik, C.I., Glaser, B., Prentice, K.M., Gélinas, Y.

Comparative analysis of black carbon in soils. Global Biogeochem. Cy. 2001; 15: 163-167.

35
ACCEPTED MANUSCRIPT

Schneider, M.P.W., Smittenberg, R.H., Dittmar, T., Schmidt, M.W.I. Comparison of gas and

liquid chromatography for the determination of benzenepolycarboxylic acids as molecular tracers

of black carbon. Org. Geochem. 2011; 42: 275-282.

PT
Seiler, W., Crutzen, P.J. Estimates of gross and net fluxes of carbon between the biosphere and

RI
the atmosphere from biomass burning. Clim. Change 1980; 2: 207-247.

SC
NU
Suman, D.O. Charcoal production from agricultural burning in central Panama and its deposition
MA
in the sediments of the Gulf of Panama. Environ. Conserv. 1986; 13: 51-60.

Suman, D.O., Kuhlbusch, T.A.J., Lim, B. Marine sediments: A reservoir for black carbon and
D

their use as spatial and temporal records of combustion. In: J.S. Clark et al., editors. NATO ASI
TE

Series I: Global Environmental Change, Berlin: Springer-Verlag; 1997 p. 271-293.


P

Veilleux, M.-H., Dickens, A., Brandes, J., Gélinas, Y. Density separation of combustion-derived
CE

soot and petrogenic graphitic black carbon: Quantification and isotopic characterization. IOP
AC

Conf. Series: Earth and Environmental Science 2009; 5: doi:10.1088/1755-1307/5/1/012010

Verardo, D.J., Ruddiman, W.F. Late Pleistocene charcoal in tropical Atlantic deep-sea

sediments: climatic and geochemical significance. Geology, 1996; 24: 855-857.

Wolf, M., Lehndorff, E., Wiesenberg, G.L.B., Stockhausen, M., Schwark, L., Amelung, W.

Towards reconstruction of past fire regimes from geochemical analysis of charcoal. Org.

Geochem. 2012; http://dx.doi.org/10.1016/j.orggeochem.2012.11.002.

36
ACCEPTED MANUSCRIPT

Worwoods, M.J., Louchouarn, P., Kuo, L.-J., Harvey, O.R. Characterization and biodegradation

of water-soluble biomarkers and organic carbon extracted from low temperature chars. Org.

Geochem. 2012; in press.

PT
Zamora, R., Molina-Martínez, J.R., Herrera, M.A., Rodríguez Silva, F. A model for wildfire

RI
prevention planning in game resources. Ecol. Model. 2010; 221: 19-26.

SC
Zimmermann, M., Bird, M.I., Wurster, C., Saiz, G., Goodrick, I., Barta, J., Capek, P.,

Santruckova, H., Smernik, R. Rapid degradation of pyrogenic carbon. Glob. Change Biol. 2012;

NU
18: 3306-3316.Ziolkowski, L. A. Radiocarbon of black carbon in marine dissolved organic
MA
carbon, Ph.D. thesis, Earth Syst. Sci., Univ. of Calif., Irvine; 2009.

Ziolkowski, L.A., Druffel, E.R.M. The feasibility of isolation and detection of fullerenes and
D

carbon nanotubes using the benzene polycarboxylic acid method. Mar. Poll. Bull. 2009; 59: 213-
TE

218.
P

Ziolkowski, L.A., Druffel, E.R.M. Aged black carbon identified in marine dissolved organic
CE

carbon. Geophys. Res. Lett. 2010; 37, L16601, doi:10.1029/2010GL043963.


AC

Ziolkowski, L.A., Chamberlin, A.R., Greaves, J., Druffel, E.R.M. Quantification of black carbon

in marine systems using the benzene polycarboxylic acid method: a mechanistic and yield study.

Limnol. Oceanogr.-Meth. 2011; 9, 140-149.

37
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
TE
P

Figure 1
CE
AC

38
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D

Figure 2
TE
P
CE
AC

39
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
Figure 3
D
TE
P
CE
AC

40
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
TE
P

Figure 4
CE
AC

41
ACCEPTED MANUSCRIPT

Table 1. Geochemical characterization of the sediments (granulometry, OC%, C/N and δ13C data
from Sánchez-García et al., 2008). Black carbon (BC) concentrations in the GoC sediments
obtained with the two analytical methods and the average number of acids (N) for the molecular
marker-based BC.
Guadia Sand+S Cla OC C/ δ13C BPC

PT
Samp na ilt y N BC (% A Nc
BC (% dw)
le distanc TOC)
e

RI
(%) (%) (% (‰) GB GB BPC
(km) BPCAb
) Ca C A

SC
83 1.0 9.8 - 0.01±0. 1.0±0.
S-1 0.09 8.7
5.4 15 1 25.9 00 4 3.3

NU
66 0.8 7.0 - 0.04±0. 4.5±4.
S-3 0.08 9.5
6.4 34 4 25.5 03 1 4.4
71 0.7 8.2 - 0.15±0.
MA
S-2 0.11 15 19±19
7.0 29 7 25.3 10 5.0
66 0.7 7.3 - 0.01±0. 1.0±0.
S-4 0.03 4.1
7.2 34 5 25.5 00 5 3.9
D

63 0.8 8.0 - 0.14±0. 17±8.


S-5 0.03 3.3
TE

9.6 37 2 25.7 07 5 4.8


80 0.5 7.3 - 0.17±0.
S-6 0.02 3.6 34±25
11.1 20 0 25.1 12 4.8
P

56 1.3 15 - 0.17±0. 13±2.


CE

S-12 0.04 3.1


12.0 44 8 25.5 04 8 5.1
61 1.4 23 - 0.17±0. 12±2.
S-11 0.04 3.0
AC

12.9 39 24.7 03 2 5.0


54 0.9 7.1 - 0.23±0. 24±7.
S-10 0.03 3.5
16.4 46 8 24.4 07 1 5.0
67 0.7 8.4 - 0.04±0. 4.9±1.
S-13 0.01 2.0
21.8 18 2 22.6 01 9 5.1
56 1.1 - - 0.04±0. 3.6±0.
S-14 0.04 3.5
24.7 44 2 23.1 01 7 4.7
55 0.6 6.2 - 0.04±0. 5.8±1.
S-15 0.03 3.8
27.3 45 9 22.8 01 9 4.8
58 0.8 8.7 - 0.01±0. 0.9±0.
S-8 0.03 3.7
35.5 42 9 22.8 00 6 3.9
64 0.6 6.8 - 0.04±0. 6.6±5.
S-7 0.02 3.0
38.1 36 4 23.0 04 9 4.3
S-9 48.5 57 43 0.6 6.8 - 0.05 0.08±0. 7.2 13±5. 4.5

42
ACCEPTED MANUSCRIPT

4 23.0 03 0
54 15 0.5 6.2 -
Min 0 25.9 0.01 0.01 2.0 0.9
3 3.3
83 46 1.4 23 -

PT
Max 0 22.5 0.11 0.23 15 34
7 5.1

RI
64 34 0.8 9.2 -
Mean 8 24.3 0.04 0.09 5.1 11

SC
1 4.6
9 10 0.2 4.6 1.28
SD 0.03 0.08 3.4 10
6 0.5

NU
a
Graphitic-Black Carbon.
b
Benzenepolicarboxylic Acids (replicates per sample n=3).
c
MA
Average number of acids (N) = Σ[BnCA]*n/ Σ[BnCA], where [BnCA] is the concentration of a
determined acid and n is the number of benzene rings forming the acid (Ziolkowski and Druffel,
2010).Table 2. Comparison of the BC burial fluxes (see Table 1 for acronyms) and estimate of
the annual burial sink in the GoC.
D
P TE
CE
AC

43
ACCEPTED MANUSCRIPT

Table 2

Sedimentation BC burial flux


Sample density
rate (μg/cm2∙yr)

T
(g/cm3) (cm/yr) GBC BPCAa

IP
S-1 2.76 0.43 262 31±13
S-3 2.62 0.44 227 108±101

R
S-2 2.35 0.43 285 376±273

SC
S-4 2.59 0.43 85 21±12
S-5 2.63 0.43 77 404±214

NU
S-6 2.70 0.43 53 491±367
S-12 2.50 0.43 115 466±139
S-11 2.50 0.42 111 462±123
MA
S-10 2.56 0.42 94 633±227
S-13 2.74 0.43 42 105±45
S-14 2.53 0.41 103 105±29
D

S-15 2.60 0.42 71 109±42


TE

S-8 2.63 0.42 90 23±15


S-7 2.72 0.43 56 124±113
P

S-9 2.53 0.42 122 213±94


CE

mean±SD 120±76 245±205


BC sink (103 t/yr)b 13±8.6 26±23
AC

a
The uncertainty of the 15 BPCA burial fluxes is estimated from propagating the analytical
uncertainty of the BPCA concentrations (n=3).
b
Accounting for an ARGIS-determined accumulation area of 10,788 km2, between the San
Vicente Cape and Gibraltar Rock (Fig. 1). The uncertainty for the BC sink is estimated from
the spatial variability of BC in the 15 marine sediments (i.e. propagated from the standard
deviation of the mean BC burial flux in the GoC).

44
ACCEPTED MANUSCRIPT

Different pools of black carbon in sediments from the Gulf of Cádiz (SW Spain): method
comparison and spatial distribution
Laura Sánchez-García, José R. de Andrés, Yves Gélinas, Michael W.I. Schmidt, Patrick
Louchouarn

T
IP
Research highlights:

R
SC
 We assess the terrestrial input of BC to the Gulf of Cadiz Continental Shelf.
 Bulk (thermal) and molecular (biomarkers) level techniques are combined.

NU
 The relative contribution of two end-members of the BC set is accounted.
 We derive regional burial and sink fluxes of the two measured pools of BC.
MA
 This provides new insights on the origin and nature of the sedimentary BC.
D
P TE
CE
AC

45

You might also like