You are on page 1of 15

Applied Geochemistry 23 (2008) 3597–3611

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: www.elsevier.com/locate/apgeochem

Environmental isotopes (N, S, C, O, D) to determine natural


attenuation processes in nitrate contaminated waters:
Example of Osona (NE Spain)
Laura Vitòria a,*, Albert Soler a, Àngels Canals a, Neus Otero a,b
a
Grup de Mineralogia Aplicada i Medi Ambient, Departament de Cristal.lografia, Mineralogia i Dipòsits Minerals, Facultat de Geologia, Universitat de
Barcelona, Martí i Franquès, s/n, 08028, Barcelona, Spain
b
Departament de Geologia Ambiental, Institut de Ciències de la Terra ‘‘Jaume Almera”, CSIC C/Lluís Solé i Sabarís s/n, Barcelona 08028, Spain

a r t i c l e i n f o a b s t r a c t

Article history: Nitrate-contaminated groundwater from an aquifer in the Osona region (NE Spain) was
Received 25 May 2006 chemically and isotopically (d15 NNO3 ; d18 ONO3 ; d34 SSO4 ; d18 OSO4 , dD, d18 OH2 O and d13 CDIC ) char-
Accepted 4 July 2008 acterized. Diffuse- NO3 contamination reached values of 366 mg/L. Nearly 75% of the 37
Available online 30 August 2008
sampled sites had higher concentrations than the 50 mg/L in NO 3 limit for drinking water.

Editorial handling by R.S. Harmon


To identify the source of pollution d15 NNO3 and d18 ONO3 were used with results, for most
samples, in the range of pig manure NO 3 . Nitrification processes were evaluated by means
of the d18O of NO3 and water. Isotopic data suggested that natural attenuation of NO3 was


taking place. This process was confirmed using the d18 ONO3 coupled with the NO 3 =Cl ratio,
avoiding the influence of continuous NO 3 inputs. A further insight on denitrification pro-
cesses was obtained by analyzing the ions involved in denitrification reactions. Although
the role of organic matter oxidation could neither be confirmed nor discarded, this
approach showed a link between denitrification and pyrite oxidation. Therefore, in areas
with no adequate infrastructure (e.g. multipiezometers), such as the one studied, this
approach could be useful for implementing better water management.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction several aquifers have NO 3 concentrations up to 10 times


higher than the 50 mg/L of NO3 threshold permitted for
Groundwaters are one of the best water resources for human consumption (EC, 1998). The Catalan Water
human consumption due to their normally high chemical Agency wants to characterise NO 3 contamination in the
quality. However, there is increasing groundwater NO 3 region, and discriminate between the different sources of
pollution worldwide due to the intensive use of synthetic pollution, in order to carry out the basic principle of the
and organic fertilizers, livestock piles, leaks from septic European Union Treaty regarding communitarian policies
tanks, and sewer systems in poor conditions. As for concerning environment – namely, that the ‘‘polluter
Europe, in 1991, an EU Council Directive (91/676/EEC Offi- should pay” (EC, 1997).
cial Journal, 1991), known as the Nitrates Directive, was Classical chemical water analyses only show the magni-
formulated to protect waters against NO 3 contamination tude of NO3 contamination not the source of the pollution.
in agricultural areas. Each State member must identify Therefore, in order to identify and/or quantify the origin of
its vulnerable zones – zones draining into water affected this type of contamination new tools, such as stable iso-
by or vulnerable to N pollution – and establish a Codes topes, are needed. Over the last 30 a, different studies have
of Good Agricultural Practice. In Catalonia (NE Spain), used the d15N and/or d18O of dissolved NO 3 in waters as
tracers of organic (e.g. animal and human wastes) and
* Corresponding author. Fax: +34 93 402 13 40. inorganic (e.g. synthetic fertilizers) contamination with
E-mail address: lvitoria@ub.edu (L. Vitòria). relative success (Wassenaar, 1995; Kendall and McDonnell,

0883-2927/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apgeochem.2008.07.018
3598 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

1998; Widory et al., 2004, among others). One of the con- with its 1263.8 km2, has more than 60,000 sheep, 100,000
clusions that emerge from these studies is the need for a cows and around 1,000,000 pigs in more than 1000 pig
clear chemical and isotopic characterization of the contam- farms (IDESCAT, 1999), most of them concentrated in
ination sources, as well as the identification of the fraction- small areas. This intensive activity produces great amounts
ation processes in order to interpret the isotopic data. For of organic animal wastes (more than 10,000 tones N/a),
example, denitrification takes place under specific phys- mainly as pig manure. A small part (around 3% in 2001)
ico-chemical conditions, where bacteria (Pseudomonas, is processed in treatment plants and the rest is spread
Bacillus, Thiobacillus, Propionbacterium and others) reduce out on the fields. As the application of pig manure is
NO 15
3 to N2, increasing the isotopic composition (d N and banned near urban areas, due to the bad smells, synthetic
d18O) of the remaining NO 3 in waters (Böttcher et al., fertilizers are also used. The municipalities of the area
1990). This is considered a drawback in the application of (Manlleu, Torelló and St. Hipòlit de Voltregà, with over
NO 3 isotopes as tracers of sources, nevertheless, it can be 30,000 people) are connected to two sewage stations with
applied to trace processes. Indeed, the isotopic composi- N elimination treatment designed to treat two and three
tion of dissolved NO 3 is used to distinguish between times more wastewater than they actually do. As a result,
dilution and denitrification in areas where a decrease in the contribution of other N sources, such as domestic sew-
NO 3 concentration in groundwater samples is observed age, is negligible compared to agricultural sources. Resid-
(Grischek et al., 1998; Mengis et al., 1999; Cey et al, ual waters are discharged into the Ter River that crosses
1999). A further step in the investigation of denitrification the area (Fig. 1).
processes is to determine the factors controlling the reac- From a geological perspective, this hydrogeological sys-
tions. This has been done coupling chemical data with tem lies on Paleogene sedimentary materials overlaying
the d15N and/or d18O of dissolved NO 3 and the isotopic Hercynian crystalline (igneous and metamorphic) rocks.
composition of the ions involved in denitrification reac- The stratigraphic sequence primarily consists of carbonate
tions such as d34S and d18O of dissolved SO2 4 , and/or formations, with alternations of calcareous, marl and car-
d13C, of dissolved inorganic C (Aravena and Robertson, bonate sandstone layers (Reguant, 1967). These up to
1998; Pauwels et al., 2000). 300 m-thick formations, show a quite uniform dip of about
The aim of this study was to integrate chemical and sta- 7–10° W, and they only have an antiform structure at their
ble isotope data of groundwater in the Osona region (NE northern limit. It is also worth mentioning the presence of
Spain) – one of the vulnerable zones – in order to evaluate framboidal pyrite in these formations (Urquiola, 1994).
the state of the water resources and predict their evolution. Hydrogeologically, the system consists of a series of con-
In the area, pig manure had been intensively used as or- fined aquifers located in the carbonate and carbonate-
ganic fertilizer producing diffuse NO 3 contamination, sandstone layers, whose porosity is mainly related to the
although synthetic fertilizers have also been occasionally fracture network. Marl strata act as confining layers, as ob-
used. Therefore, the multi-isotopic approach (15N, 34S, served in boreholes and hydrogeological prospecting. Main
13
C, 18O, D) coupled with the chemical data, was applied production wells for agriculture and farm demand usually
first to validate their use as tracers of NO 3 sources, and, reach depths of more than 100 m, searching for the most
secondly as markers of N processes. In particular, the nitri- productive confined aquifers. Alluvial aquifers are scarce
fication process was evaluated using the d18 ONO3 value; the and generally non-productive in the area, except those lo-
amount of denitrification was estimated taking into con- cated at the Ter river terraces.
sideration both the chemical and the d15 NNO3 and d18 ONO3
data. Furthermore, the denitrification reactions were 3. Sampling and analysis
determined to some extent applying a multi-isotopic ap-
proach (d34 SSO4 ; d18 OSO4 and d13 CDIC ). This knowledge is Thirty-seven groundwater samples (31 from private
important when designing water and agricultural control parapet wells and 6 from natural springs) were collected
and management plans. between 9 and 26 October 2001 (Fig. 1). Sampled ground-
waters were shallow, with piezometric depths from 0 to
2. Study area 45 m, and were collected with the pumps already installed
in wells or with portable pumps. In both cases, the sam-
The Osona region (NE Spain) is one of the areas classi- pling was carried out after extracting the routine 3 well
fied as vulnerable to nitrate contamination from agricultural volumes of water. The piezometric level, pH, conductivity
sources by the European Union (EEC, 1991). The region is and water temperature were measured in situ. Samples
situated to the North of Barcelona, at the East of the Ebro were kept at 4 °C for their subsequent chemical and isoto-
basin, between the Montseny Mountain and the Pyrenees pic analyses. In the laboratory, pH and conductivity at
range (Fig. 1). The Ter river is the major stream in the area, 25 °C were measured again; NHþ 4 concentration was
with a mean discharge of 509 hm3/a at the gauging station analyzed by colorimetry (continuous flow analyzer, Tech-
of Roda de Ter. The area has a sub-Mediterranean climate nicon) and total organic C (TOC) by organic matter com-
with an annual mean temperature of 13 °C, precipitation bustion (TOC 500 SHIMADZU). Afterwards a split of the
of around 530 mm/a, and potential evapotranspiration sample was filtered through a 0.45 lm MilliporeÒ filter;
3 2
greater than rainfall (740 mm/a). Therefore, Osona is a the anion (NO 
3 , NO2 , PO4 , SO4 and Cl) content was
dry area that mainly relies on ground water resources to measured by High Performance Liquid Chromatography
fulfil its agricultural and farm demand. Surface water from (WATERS 515 HPLC, with WESCAN and UV/VIS KONTROL
the Ter River covers most of the urban demand. The region, detectors); concentrations of Na, Ca, K, Mg, Sr, Fe, B, Zn
L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611 3599

Fig. 1. Geological map of the study area, between the municipalities of Manlleu and Torelló. The October 2001 groundwater chemistry is represented with
Stiff diagrams.
3600 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

were measured by Inductively Coupled Plasma-Optical were analyzed in a Carlo Erba Elemental Analyzer (EA)
Emission Spectrometry (ICP-OES, Thermo Jarell ASH 61- coupled in continuous flow to a Finnigan Delta C IRMS.
E); Ni, Co, Mn, Ba, Zn, Cu, Cr, As, Se, Cd, V, Hg, Pb by Induc- The d18 ONO3 and d18 OSO4 were analyzed in duplicate with a
tively Coupled Plasma-Mass Spectrometry (ICP-MS, Perkin ThermoQuest TC/EA (high temperature conversion-ele-
Elmer Elan 6000); and alkalinity ðHCO 3 Þ by titration mental analyzer) unit coupled with a Finnigan Matt Delta
(METROHM 702 SM Titrino). The quality of the chemical C IRMS using the method of Kornexl et al. (1999). Notation
analysis was checked by doing an ionic mass balance, is expressed in terms of d ‰ relative to the international
accepting an error lower than 5%. standards V-SMOW (Vienna Standard Mean Oceanic Water)
For the d15 NNO3 and d18 ONO3 analysis, dissolved nitrates for dD and d18O; AIR (atmospheric N2) for d15N, V-CDT
were concentrated using anion-exchange columns Bio (Vienna Canyon Diablo Troilite) for d34S; and V-PDB (Vienna
RadÒ AG 1-X8(Cl) 100–200 mesh resin, after extracting Peedee Belemnite) for d13C. Precision (1r) of the samples
the sulphates and phosphates by precipitation with BaCl2 calculated from international and internal standards sys-
and filtration (Mayer et al., 2001). Afterwards dissolved ni- tematically interspersed in the analytical batches was
trates were eluted with HCl and converted to AgNO3 by ±1.5‰, ±0.2‰, ±0.3‰, ±0.4‰, ±0.4‰, ±0.5‰ and ±0.2‰ for
adding silver oxide. The AgNO3 solution was then freeze- dD, d18 OH2 O ; d15 NNO3 ; d18 ONO3 ; d34 SSO4 ; d18 OSO4 and d13 CDIC ,
dried to purify the AgNO3 for analysis (modified from Silva respectively. A summary of international and internal
et al., 2000). For the SO2 4 isotopic analysis (d34 SSO4 and standards used for analysis and the assumed values is de-
d18 OSO4 ), dissolved SO2
4 was precipitated as BaSO4 by add- tailed in Table 1. All chemical and isotopic analyses were
ing BaCl2  2H2O after acidifying the sample with HCl and prepared in the Mineralogia Aplicada i Medi Ambient
boiling it to prevent BaCO3 precipitation following stan- Research Group laboratory and analyzed at the Scientific
dard methods (e.g. Dogramaci et al., 2001). Unfiltered Services of the University of Barcelona.
splits of samples were treated with NaOH–BaCl2 solution
to precipitate carbonates together with sulphates and 4. Results and discussion
phosphates. Later, they were filtered at 3 lm and the
d13C of the total dissolved inorganic C, mainly HCO 3 , was 4.1. Groundwater quality
analyzed (Bishop, 1990).
Deuterium and O isotopes of water were analyzed in a Chemical results are presented in Table 2 (water
Finnigan Matt Delta S Isotope Ratio Mass Spectrometer temperature, pH, conductivity and major elements) and
(IRMS) coupled to an automated line based on the equilibra- Table 3 (minor and trace elements). Water pH values were
tion between H-water and H2 gas with a Pt catalyst, and all above 7.3, temperatures ranged from 14.3 to 18.8 °C and
between O-water and CO2 gas following standard methods conductivities varied from 801 for the most pristine to
(Epstein and Mayeda, 1953). The d15 NNO3 ; d34 SSO4 and d13 CDIC 2474 lS/cm for the most polluted in terms of NO 3.

Table 1
Standards used in the isotopic determinations

Ratio Scale Reference materials Matrix d (‰) Used for the determination of
2
H/1H VSMOW (Vienna Standard Mean Ocean Water) NEUa Water 71.3 dD–H2O
LAB 6a Water 48.0 dD–H2O
MARa Water 2.2 dD–H2O
13
C/12C VPDB (Vienna Peedee Belemnite) IAEA-CH6b Sucrose 10.43 ± 0.13 d13C–DIC
IAEA-CH7b Polyethylene 31.83 ± 0.11 d13C–DIC
USGS 24b C (Graphite) 15.99 ± 0.11 d13C–DIC
15
N/14N AIR-N2 (Atmospheric N2) IAEA-N1b (NH4)2SO4 +0.43 ± 0.07 d15N–NO3
IAEA-N2 (NH4)2SO4 +20.41 ± 0.12 d15N–NO3
IAEA-NO3b KNO3 +4.72 ± 0.13 d15N–NO3
USGS-25b (NH4)2SO4 30.25 ± 0.38 d15N–NO3
18
O/16O VSMOW (Vienna Standard Mean Ocean Water) NEUa Water 10.63 d18O–H2O
LAB 6ª Water 7.29 d18O–H2O
MARa Water +0.48 d18O–H2O
NBS127b BaSO4 +9.3 ± 0.4 d18O–NO3, d18O–SO4
YCEMa BaSO4 +17.6 ± 0.5 d18O–NO3, d18O–SO4
H2SO4a BaSO4 +12.9 ± 0.6 d18O–NO3, d18O–SO4
Camb 35-4c AgNO3 +0.5 d18O–NO3
Chilean KNO3c KNO3 +45.2 d18O–NO3
EGC-1c KNO3 +28.0 d18O–NO3
EIL61c KNO3 +11.0 d18O–NO3
EIL62c AgNO3 +20.5 d18O–NO
34
S/32S VCDT (Vienna Canyon Diablo Troilite) IAEA-S1b Ag2S 0.30 (Exact.) d34S–SO4
IAEA-S2b Ag2S +22.67 ± 0.15 d34S–SO4
IAEA-S3b Ag2S 32.55 ± 0.12 d34S–SO4
NBS127b BaSO4 +21.1 d34S–SO4
YCEMb BaSO4 +12.5 d34S–SO4
a
Internal standard (Serveis Cientificotecnics de la Universitat de Barcelona).
b
Internal Standard.
c
Environmental Isotopes Laboratory, Universitty of Waterloo, Canada.
Table 2
Water temperature, pH, conductivity, and major elements of sampled groundwater in Osona, October 2001

Sample Water Temp (°C) pH at 25 (°C) Cond. at 25 °C (lS/cm) NO


3 (mg/L) SO2
4 (mg/L) HCO3 (mg/L) Cl (mg/L) Na+ (mg/L) K+ (mg/L) Mg2+ (mg/L) Ca2+ (mg/L)
PO-001 16.3 7.9 1623 173.8 252.1 431.1 76.8 37.1 13.5 48.7 216.8
PO-004 17.0 7.8 1144 64.1 170.3 509.3 38.0 21.1 19.8 43.3 172.8
PO-012 17.3 7.7 1585 140.8 223.2 400.4 101.9 42.5 8.4 44.2 211.7
PO-020 15.1 7.5 1364 80.8 204.9 458.5 67.8 29.1 3.5 43.9 187.5
PO-036 16.1 7.5 1249 109.5 135.2 425.7 45.3 22.1 7.1 36.3 171.3
PO-037 18.5 7.5 1112 83.1 105.4 378.0 51.0 23.2 15.1 24.6 148.5
PO-043 16.7 7.6 675 15.6 101.6 265.9 3.0 4.3 2.7 18.8 97.4
PO-051 15.7 1.1 1098 58.7 162.5 387.1 24.7 16.8 3.9 37.7 138.9

L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611


PO-052 17.7 1.1 1011 30.5 90.6 465.1 21.9 13.1 1.8 24.8 148.5
PO-059 16.0 8.2 1514 77.3 239.3 471.3 74.9 53.0 7.0 50.3 163.8
PO-062 15.6 7.6 1265 96.8 140.0 445.2 52.7 18.4 4.1 38.0 169.1
PO-063 18.5 7.8 1486 107.3 214.0 491.7 57.9 26.6 10.3 52.7 185.0
PO-065 14.3 7.6 917 91.5 75.0 328.9 28.6 11.9 4.9 30.7 113.8
PO-070 16.9 7.3 1261 83.6 200.9 429.6 51.3 22.6 5.6 42.0 163.6
PO-084 16.8 7.6 1054 38.9 137.4 384.8 35.2 41.6 4.2 38.6 105.5
PO-086 17.9 7.5 1043 78.2 125.2 367.2 25.6 18.6 3.3 35.8 125.5
PO-100 16.7 7.5 2474 365.7 244.1 480.4 210.5 78.0 77.8 55.8 279.1
PO-101 18.8 7.8 989 68.7 156.0 253.0 47.3 27.4 3.6 17.3 136.5
PO-102 n.d. 7.8 791 18.8 78.7 281.0 38.9 23.5 6.0 20.4 90.5
PO-103 18.7 8.1 801 <0.2 152.5 373.6 17.2 44.4 2.8 35.5 92.6
PO-104 16.1 1.1 1155 55.5 194.0 418.1 25.7 23.9 5.8 39.5 144.4
PO-105 15.7 7.6 1405 95.2 234.0 435.0 49.5 66.2 5.3 50.0 141.8
PO-106 17.2 8.0 1097 12.5 140.0 409.7 40.9 174.9 4.8 16.3 31.0
PO-107 18.7 7.5 1055 52.2 149.3 388.3 27.9 23.1 6.0 35.3 140.0
PO-108 16.1 7.4 1365 176.9 146.0 372.1 70.1 29.0 10.4 34.1 184.6
PO-109 16.2 7.5 1225 75.4 197.6 391.6 36.0 31.2 3.4 42.7 144.9
PO-110 16.5 7.5 1463 127.8 188.2 457.1 68.1 37.3 6.6 43.5 178.0
PO-111 15.6 7.4 1157 105.1 95.0 439.4 37.6 20.5 3.5 24.0 174.2
PO-112 16.8 8.0 835 17.0 91.6 387.9 11.7 12.3 1.9 24.8 119.2
PO-113 16.9 1.1 879 46.5 103.7 343.9 19.2 12.6 1.8 24.6 123.8
PO-114 17.5 1A 1285 149.2 118.1 422.4 44.0 28.6 5.7 31.5 170.8
PO-115 15.6 7.3 1752 325.3 151.2 408.0 83.4 35.4 20.9 39.0 233.0
PO-116 16.3 7.4 976 28.9 154.6 392.1 11.3 8.8 4.1 31.5 140.2
PO-117 15.3 7.5 1400 72.8 252.7 396.3 53.1 46.9 7.3 41.1 171.0
PO-118 15.3 7.9 882 6.8 148.8 377.1 15.3 8.8 6.7 23.5 137.1
PO-119 17.8 1.1 1091 62.7 98.6 405.6 44.4 56.8 3.7 26.4 117.6
PO-120 17.6 7.9 1092 66.5 170.3 373.8 38.4 16.2 2.1 37.2 150.1
PO-121 18.6 7.8 1101 78.5 110.6 431.2 40.2 13.2 3.2 38.2 145.3
98/83/CEE treshold P6.5 and 69.5 <2500 at 20 °C 50 250 – 250 200 – – –

3601
3602 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

Table 3
Minor and trace elements of sampled groundwater in Osona (October 2001)

Site NHþ
4 (mg/L) NO
2 (mg/L) TOC (mg/L) Sr (mg/L) Ba (mg/L) Mn (mg/L) Cu (mg/L) Zn (mg/L) B (mg/L) Pb (mg/L)
PO-001 <0.1 <0.2 1.7 1.701 0.095 0.0004 0.006 0.031 <0.025 0.0002
PO-004 1.08 0.7 3.2 1.717 0.081 0.1812 0.004 0.008 <0.025 <0.0001
PO-012 <0.1 <0.2 1.3 1.652 0.094 0.0005 0.005 0.015 <0.025 <0.0001
PO-020 <0.1 <0.2 1.2 1.734 0.098 0.0289 0.004 0.007 <0.025 0.0002
PO-036 <0.1 <0.2 1.5 1.454 0.083 0.0003 0.002 0.002 <0.025 0.0005
PO-037 <0.1 <0.2 1.2 0.900 0.094 0.0002 0.005 0.008 <0.025 <0.0001
PO-043 <0.1 <0.1 5.3 0.839 0.026 0.0009 0.008 0.015 <0.025 0.0002
PO-051 <0.1 0.6 <1 2.007 0.102 0.0006 0.002 0.002 <0.025 0.0002
PO-052 <0.1 <0.1 1.2 0.751 0.053 0.0003 0.002 0.006 <0.025 0.0004
PO-059 <0.1 <0.2 1.2 3.060 0.080 0.0017 0.005 0.956 <0.025 <0.0001
PO-062 <0.1 <0.2 1.6 1.626 0.083 0.0007 0.002 0.009 <0.025 <0.0001
PO-063 <0.1 <0.2 1.0 2.192 0.071 0.0003 0.012 0.048 <0.025 0.0003
PO-065 <0.1 <0.1 <1 0.536 0.091 0.0002 <0.001 0.002 <0.025 <0.0001
PO-070 <0.1 <0.2 1.2 1.617 0.069 0.0003 0.004 0.016 <0.025 <0.0001
PO-084 <0.1 <0.2 2.4 4.770 0.034 0.0047 0.004 0.011 <0.025 0.0006
PO-086 <0.1 <0.1 3.2 3.728 0.226 0.0118 0.007 1.461 <0.025 0.0004
PO-100 <0.1 <0.3 4.1 1.480 0.127 0.0015 0.007 0.025 <0.025 <0.0002
PO-101 <0.1 0.4 1.6 0.826 0.072 0.0069 0.006 0.035 <0.025 0.0011
PO-102 <0.1 0.2 <1 0.817 0.041 0.0002 0.004 0.010 <0.025 0.0002
PO-103 <0.1 <0.2 <1 3.680 0.046 0.0087 0.002 0.005 <0.025 <0.0001
PO-104 0.14 0.3 1.6 5.037 0.042 0.0022 0.002 0.183 <0.025 <0.0001
PO-105 <0.1 <0.2 <1 12.065a 0.050 0.0067 0.002 0.018 <0.025 <0.0001
PO-106 <0.1 <0.2 <1 3.041 0.028 0.0014 0.002 0.041 0.061 <0.0001
PO-107 <0.1 <0.2 1.1 1.557 0.046 0.0019 0.009 0.030 <0.025 0.0005
PO-108 <0.1 <0.2 2.7 1.026 0.084 0.0013 0.003 0.008 <0.025 <0.0001
PO-109 <0.1 0.3 1.1 1.700 0.040 0.0006 0.002 0.005 0.062 <0.0001
PO-110 <0.1 <0.2 1.7 2.219 0.079 0.0025 0.004 0.009 0.052 <0.0001
PO-111 <0.1 <0.2 <1 0.654 0.078 0.0004 0.002 0.002 0.085 <0.0001
PO-112 <0.1 <0.1 <1 0.972 0.047 0.0007 0.002 0.447 <0.025 0.0004
PO-113 0.11 <0.1 1.5 0.822 0.051 0.0004 0.003 0.011 <0.025 0.0002
PO-114 0.21 0.9 4.3 1.419 0.068 0.0133 0.033 0.051 0.026 0.1525
PO-115 <0.1 0.2 12.6 1.174 0.142 0.0218 0.027 0.117 0.069 0.0017
PO-116 <0.1 <0.1 2.2 1.407 0.057 0.0011 0.005 0.013 <0.025 0.0001
PO-117 0.13 2.3 2.0 1.940 0.105 0.0014 0.004 0.007 0.079 0.0002
PO-118 <0.1 <0.1 5.5 0.654 0.038 0.0003 0.006 0.024 <0.025 0.0001
PO-119 <0.1 <0.2 1.5 1.387 0.056 0.0060 0.008 1.752 0.119 0.0023
PO-120 <0.1 <0.2 1.5 1.151 0.059 0.0001 0.003 0.008 <0.025 0.0001
PO-121 <0.1 <0.2 <1 0.875 0.053 0.0002 0.002 0.002 <0.025 0.0001
Analytical precision ±0.005 ±0.0001 ±0.0001 ±0.001 ±0.001 ±0.025 ±0.0001
98/83/CEE treshold 0.5 0.5 – – – 0.05 – – 1 0.01

PO4, V, Cr, Fe, Co, Ni, Cd, Hg, As and Se concentrations were under their detection limits (10, 0.004, 0.002, 0.005, 0.001, 0.01, 0.0002, 0.0004, 0.002 and 0.01,
respectively) except for Fe in sample PO-115 with 0.007 mg/L.
a
Analytical precision of ±0.025 mg/L.

Groundwaters in the area are mainly HCO3–Ca type, in thetic fertilizers (Vitòria et al., 2004a; Otero et al., 2005).
accordance with the lithology of the aquifer. Only 3 sam- Sulphate content of waters ranged between 75 and
ples are HCO3–SO4–Ca, one HCO3–Cl–Ca and one HCO3– 253 mg/L with a mean value of 157 mg/L, however, no
Na water type (Fig. 1). Fig. 2a shows the piezometric clear correlation was observed between SO2 4 concentra-
groundwater surface and Fig. 2b the NO 3 concentration tions and other major or trace elements. No evaporite out-
which varies from undetected up to 366 mg/L, with mean crops exist in the studied area, and groundwater SO2 4 is
values of 90 mg/L of NO3 . Around 75% of the studied sam- assumed to be related to oxidation of disseminated pyrite.
ples had NO 3 concentration above the drinking water Potentially harmful elements (As, Cd, Co, Cr, Hg, Ni, Se and
threshold. Maximum NO3 contamination was concentrated V) were below the detection limit; only Cu, Zn, Mn and Pb
in small areas that could coincide with higher pig manure were detected. In a few samples Pb and Mn were found to
application or, locally, with wells supplying shallower be slightly above the threshold given by Council Directive
groundwater, though no relationship was observed be- 98/83/EC (EC, 1998) (Table 3). Pig manure and synthetic
tween well depth and NO 3 concentration. Ammonium con- fertilizers had different trace element contents: higher
tent, detected in only 5 samples, was below the drinking contents of Fe, Zn, Cu and Mn in manure, and higher con-
water threshold, except for one sample; NO 2 was unde- tents of As, Cd, Cr or U among others, in chemical fertilizers
tected in most samples, but exceeded the 0.5 mg/L thresh- (Vitòria et al., 2004a; Otero et al., 2005, and references
old in 4 samples. TOC ranged from undetected to 12.6 mg/ therein). Although the measured potentially harmful ele-
L, with a mean value of 1.9 mg/L. Nitrate showed a positive ment concentrations could suggest a link with pig manure,
tendency with Cl (Fig. 3a), Ca (Fig. 3b) and Ba – elements trace element contents are not a conclusive tracer, as soil
found in high concentrations in both pig manure and syn- processes and plant uptake were not evaluated. Conse-
L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611 3603

Fig. 2. (a) Groundwater piezometric surface in meters above the sea level (masl) corresponding to October 2001. (b) Isoconcentrations of NO
3 in mg/L of the
same date.

a 225
b 300

200
250
175

150 200
r 2 = 0.70
Ca 2+ (mg/l)

r 2 = 0.71
Cl- (mg/l)

125
150
100

75 100

50
50
25

0 0
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
NO3- (mg/l) NO3- (mg/l)

Fig. 3. (a) NO  


3 versus Cl , (b) NO3 versus Ca.

quently, the water chemistry alone did not allow identifi- and dissolved inorganic C (d13 CDIC ) are shown in Table 4.
cation of the source of N in these groundwaters. Compositions of 2H and 18O fall on the local meteoric water
Chemical data have also been used to outline processes, line LMWL (Fig. 4) calculated with data from the Global

e.g. NO
3 =Cl ratio is used to distinguish between dilution Network of Isotopes in Precipitation (GNIP) from stations
and denitrification in groundwater samples (Altman and 0818001 and 0818002 located in Barcelona (IAEA/WMO,
Parizek, 1995). This approach could not be applied in the 2004). Groundwater dD and d18 OH2 O values in the studied
studied area, where sampling relied on production wells, area corresponded to the actual meteoric waters, from lo-
and piezometers were not available. The lack of well cas- cal recharge, and were not affected by appreciable evapo-
ing, unnecessary from a construction point of view but ration. Because the weighted local rain was sampled in
which brought about the simultaneous exploitation of Barcelona, near the coast, it was shifted towards higher
the different aquifer layers, made the interpretation of isotopic compositions.
chemical data difficult. Consequently, in this context stable The d13 CDIC values ranged from 16.4‰ to 12.5‰ with
isotopes become an essential tool to determine the origin a mean value of 14.1‰ – values within the known range
and fate of N compounds. of d13 CDIC for groundwater (16‰ to 11‰, Vogel and
Ehhalt, 1963). d34 SSO4 showed a wide range of values (from
4.2. Stable isotopes 20.2‰ to +5.8‰), with a mean value of 9.4‰, and
d18 OSO4 varied from 1.3‰ to +8.7 ‰, with a mean of
The isotopic compositions of water (dD, d18 OH2 O ), dissol- +3.3‰. The d34 SSO4 and d18 OSO4 confirmed an origin of
2
ved NO 15 18 34 18
3 (d NNO3 ; d ONO3 ), dissolved SO4 (d SSO4 ; d OSO4 ) SO2
4 related to oxidation of disseminated pyrite. Most
3604 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

Table 4
Isotopic data of sampled groundwater in Osona, October 2001

Sample d2H % d18 OH2 O % d15 NNO3 % d18 ONO3 % d34 OSO4 % d18 SSO4 % d13 CHCO3 %
PO-001 41.2 6.7 12.8 7.0 0.4 4.0 14.7
PO-004 39.8 6.6 14.1 7.6 12.2 3.5 14.8
PO-012 40.2 6.7 16.5 7.1 3.2 3.3 13.2
PO-020 42.5 6.6 16.2 9.6 10.3 4.0 15.3
PO-036 41.1 6.6 13.0 6.9 9.5 1.5 15.1
PO-037 48.0 7.4 14.1 7.4 0.1 4.6 13.8
PO-043 37.5 6.0 2.2 7.3 13.8 1.1 13.7
PO-051 43.0 6.7 12.9 7.3 14.5 1.1 15.4
PO-052 42.6 6.7 12.0 7.7 14.4 2.0 15.4
PO-059 43.3 6.9 20.9 9.6 10.8 1.9 13.5
PO-062 42.0 6.5 12.8 7.7 9.4 1.9 14.4
PO-063 41.6 6.6 14.1 6.9 13.2 2.2 12.5
PO-065 41.6 6.6 7.2 4.7 0.8 3.0 12.5
PO-070 41.9 6.7 11.2 6.9 10.3 4.8 14.0
PO-084 42.5 6.7 16.6 7.8 16.2 4.2 13.1
PO-086 44.1 6.8 15.8 8.9 19.1 3.1 14.7
PO-100 43.3 6.7 13.8 4.7 5.8 6.2 13.1
PO-101 43.7 6.7 18.4 8.9 5.0 8.7 15.5
PO-102 47.4 7.7 10.5 5.6 2.7 7.5 13.3
PO-103 44.5 7.0 – – 17.9 3.9 11.6
PO-104 40.7 6.6 14.2 7.4 2.0 6.0 14.1
PO-105 40.4 6.6 16.3 8.4 16.8 3.1 13.1
PO-106 46.2 7.2 – – 12.1 5.0 11.0
PO-107 38.7 6.5 12.1 8.3 14.5 3.0 14.2
PO-108 39.0 6.5 12.8 6.5 5.4 -3.2 15.1
PO-109 41.4 6.8 13.8 7.2 11.3 2.9 14.5
PO-110 40.0 6.6 14.4 7.1 13.5 2.2 14.8
PO-111 41.8 6.8 10.4 4.6 4.1 4.2 13.4
PO-112 41.2 6.6 9.5 8.0 18.6 8.7 13.8
PO-113 44.9 7.0 7.6 5.4 13.0 2.4 14.0
PO-114 40.4 6.5 14.1 8.6 10.5 2.9 13.9
PO-115 36.6 5.9 15.0 8.0 3.3 3.6 16.4
PO-116 37.1 6.4 9.2 6.1 20.2 1.0 16.2
PO-117 41.6 6.7 10.3 9.7 14.9 0.7 14.4
PO-118 37.5 5.7 – – 14.8 3.7 13.7
PO-119 41.5 6.5 14.7 8.3 9.4 4.9 12.7
PO-120 43.7 6.8 12.1 7.9 10.3 3.8 13.4
PO-121 43.4 6.8 13.6 8.2 12.0 2.7 13.6
Analytical precision ±1.5 ±0.6 ±0.3 ±0.4 ±0.4 ±0.5 ±0.2

samples showed negative d34 SSO4 values, in accordance caused by NH3 volatilisation result in d15 NNH4 from +8‰ to
with disseminated pyrite in the aquifer (d34 SFeS2 between +15‰ (Vitòria et al., 2004a), giving a narrower range of val-
5‰ and 28‰; Viñals et al., 2002; Pierre et al. 1994), ues than those found in the literature.
coupled with low d18 OSO4 values. The positive d34 SSO4 sam- As for the NO 15
3 sources, most of the samples had d NNO3
ples could be related to anthropogenic SO2 4 sources, most

values higher than +8‰, suggesting a NO3 origin linked to
likely linked to agriculture: fertilisers and pig manure. Re- pig manure (only 3samples had lower values, which could
sults are further discussed in Section 4.4. be influenced by synthetic fertilizers, organic soil NO
3 , or a
Whereas the d15N of groundwater dissolved NO 3 ranged mixture of the two sources). The d18 ONO3 values were high-
between +2.2‰ and +20.9‰, with a mean value of +13‰, er than expected, and the samples showed a coupled in-
d18 ONO3 showed less variability, ranging from +4.6‰ to crease in d15 NNO3 and d18 ONO3 indicating the occurrence of
+9.7‰, with a mean value of +7.4‰. Results are plotted denitrification processes. However, to validate these
in a d15N vs d18 ONO3 diagram (Fig. 5a), together with the assumptions regarding NO 3 sources, and estimate the
isotopic composition of NO 3 from different sources occurrence of natural attenuation, the isotopic fraction-
(Mengis et al., 2001, and references therein). The main ation affecting N compounds must be evaluated.
sources of N in the area are linked to agriculture, and the
isotopic composition of locally used fertilisers and manure 4.3. Evaluation of fractionation processes affecting N
were in agreement with literature values. Chemical fertilis- compounds
ers commonly used in the area have d15 NNO3 between 8‰
and +7‰, with a median value of 0.0‰, and d18 ONO3 values When pig manure or ammoniacal synthetic fertilizers
around +22‰ (Vitòria et al., 2004b). By law, in the studied were applied on the fields, hydrolysis of urea and NH3 vol-
area pig manure can be stored in pits for no more than 4 atilization took place (Reaction (1)) increasing the isotopic
months, after which it is used as an organic fertiliser. Dur- composition of the residual NHþ 4 (Letolle, 1980). Although
ing the storage period, changes in both d15 NNH4 and ½NHþ 4 , it is a known fact that the fractionation factor between
L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611 3605

18
O (o/oo)
0
-14 -12 -10 -8 -6 -4 -2 0

Rains >20 mm
Rains <20 mm -20
Local Meteoric Water Line
Groundwaters Oct 01
Weighted Local Rain
-40

D ( o/oo)
46 -60
.32
+7
18 O -8 -7,5 -7 -6,5 -6 -5,5
·
8 97 -30
7.2 -80
D=
-35

-40 -100

-45

-120
-50

Fig. 4. Isotopic compositions (dD, d18O) of groundwater together with the Local Meteoric Water Line and the Weighted Local Rain, calculated from
precipitation values since 1985 at meteorological stations 0818001 and 0818002 of Barcelona, of the Global Network of Isotopes in Precipitation – GNIP –
Data Base (IAEA/WMO, 2004).

NH3 gas and NHþ 4 is 1.034 at 25 °C (Mariotti et al., 1981), to (1984), among others, the d18 ONO3 followed Reaction (3),
try to give a fractionation rate for a study area of only where it is assumed that no fractionation occurred during
38 km2 is senseless. Volatilization fractionation will de- the incorporation of O into the new molecules of NO 3
pend on the d15 NNH4 applied, the air temperature, humid- (Mayer et al., 2001).
ity, wind speed, etc.
d18 OðNO3Þ ¼ 2=3d18 OðH2OÞ þ 1=3d18 OðO2Þ ð3Þ
ð1Þ Mayer et al. (2001) suggest that Reaction (3) takes place
during chimiolithoauthotrophic nitrification, not during
heterotrophic nitrification, when less than ? of the O came
As explained in the previous section, the expected d15 NNH4 from water. This can be observed in systems where the
values for pig manure, taking into account volatilization NHþ 4 content is limited and when a low rate of nitrification
processes during the storage period, ranged from +8‰ to exists, but this was not the case in the study area, as pig
+15‰ (Vitòria et al., 2004a). Further volatilization could manure application on the fields provided an unlimited
occur once pig manure or fertilisers were applied on the amount of NHþ 4 . Therefore, Eq. (3) can be applied, and
fields, resulting in higher d15N values. Except for 3 samples NO 18
3 formed should have d ONO3 values between +2.7‰
(with d15N < 8‰), all the Osona samples showed d15N in and +4.0‰. Instead, all of the samples had higher isotopic
agreement with pig manure values, and a group of samples compositions (>+4.6‰) (Fig. 5a). This can only be explained
showed even higher values (d15N > +15%). In order to by: (a) a higher d18 ONO3 resulting from the nitrification
determine if the d15N increase is due to volatilization or reaction, (b) a mixture with another source of NO 3 (e.g.
to denitrification processes, the behaviour of the d18 ONO3 synthetic NO3-fertilisers, or atmospheric NO3), or (c) the
values must be also verified. occurrence of denitrification processes.
In the unsaturated zone, under aerobic conditions, the Slight enrichments in d18 ONO3 could be explained taking
residual NH4+ was completely nitrified (Reaction (2)). For into account the nitrification reaction, that is, differences in
this reason it was accepted that there was no difference be- the d18O of the soil O2 and/or the d18O of soil water, or be-
tween NH4–d15N and the newly formed NO 3 composition
cause during the nitrifying reaction water of a different ori-
(Heaton, 1986), which eventually becomes mixed with gin and composition was involved. Bacterial respiration
the initial pool of NO 
3 , from other N inputs, such as NO3 -
could also enrich the soil O2 in 18O, with a fractionation
fertilizers. factor of 1.015 (Guy et al., 1993). This higher soil O2 com-
position could explain the higher d18 ONO3 in groundwater.
2NHþ4 þ 3O2 ! 2NO2 þ 2H2 O þ 4Hþ
ð2Þ According to Gat (1996), Hsieh et al. (1998), among others,
2NO2 þ O2 ! 2NO3 water retained in soil as humidity is usually affected by
evaporation. Desert soils can show an increase in d18O in
During Reaction (2), according to laboratory experiments
soil water of up to 10‰ compared to the adjacent ground-
carried out by Anderson and Hooper (1983), Hollocher
water; this difference can reach 8‰ in low rain areas and
3606 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

a 24
NO3- from
Springs
NO3- in Springs Low d18O
20 fertilisers
Wells
Wells Low d18O
16
δ O-NO3 ( ‰ )
Aluvial
N/ O= 2.1 (1)
12 N/ O= 1.3 (2)
18

4 NO3- from NO3- from


NH4+ in soil organic NO3- from manure or
fertilisers N sew age

0
-3 0 3 6 9 12 15 18 21 24
15
N-NO3 ( ‰ )

b 12
Springs
Springs Low d18O
10
Wells
Springs Low d18O
δ N-NO3 ( ‰ )

8
Aluvial

6
15

0
0 50 100 150 200 250 300 350 400
-
NO3 (mg/l)

Fig. 5. (a) d15N vs. d18 ONO3 diagram. Boxes represent the isotopic composition of NO3- from different sources (Mengis et al. 2001, and references therein),
lines represent the changes in the isotopic composition due to denitrification processes. The different slopes represent the extreme values from the
literature for eN/eO (1) Böttcher et al., 1995, (2) Fukada et al. (2003) (see explanation in the text). (b) Nitrate concentration vs d15N.

2–3‰ in rainy areas. Effectively, if for the study area an an- d18 ONO3 (Kendall and McDonnell, 1998). Studied samples
nual enrichment of 2–3‰ in d18O-soil water is assumed, show an opposite trend with a coupled increase in
the theoretical composition of NO 3 formed by nitrification d18 ONO3 and d15 NNO3 suggesting that denitrification reac-
would range from +4‰ to +6‰ – closer to the lower tions could have occurred. Therefore, coupling d15N with
d18 ONO3 values found in the studied samples. Finally, the d18 ONO3 values, it became apparent that part of the variabil-
presence of water of a different origin could come from ity in the N isotopic values was due to volatilization, but
the pig manure itself, which is 70–90% water in volume. changes could also be caused by denitrification.
Although the measured d18 OH2 O from an open-air manure The coupled use of d15N and d18 ONO3 confirms a NO 3 ori-
pit of the area was in the range of groundwater (5.0‰, gin linked to organic manure. Although most samples had
Vitòria et al., 2004a), the evaporation of pig manure stored isotopic values in the range of pig manure NO 3 , one sample
in pits under atmospheric conditions could increase its iso- had a clear signature of synthetic fertilisers and 4 showed
topic composition. A combination of these 3 causes could intermediate values, suggesting a mix between these two
explain why even the lowest d18 ONO3 values detected are sources. Furthermore, a significant number of samples
slightly above the calculated values using Eq. (3). Regard- had high d18 ONO3 values coupled with high d15 NNO3 suggest-
ing mixing or denitrification processes to explain the ob- ing that denitrification is taking place.
served d18O values (up to +9.7‰), a mixture between
manure and fertilisers or atmospheric N2 would produce 4.4. Evaluation of the natural attenuation: the denitrification
higher d18 ONO3 and lower d15 NNO3 . Available mean values
for these sources are close to 0‰ and 3‰, respectively, Several processes, such as sorption, dilution and/or
for d15N, and around +20‰ and +45‰, respectively, for denitrification, can produce a natural attenuation of NO
3.
L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611 3607


Whereas only denitrification causes a real attenuation, pared to Cl. Focusing on the NO 3 =Cl ratio instead of the
sorption and dilution produce a decrease of the concentra- NO 3 concentration dispenses with the effect of new NO3


tion but the total amount of contaminant remains in the input, hence, this correlation indicates that a denitrifica-
system. Denitrification requires anoxic conditions (Clark tion reaction is taking place, producing an increase in the
and Fritz, 1997), where NO 3 is reduced to N2 or N2O by or- isotopic composition and a decrease in the concentration.
ganic matter (Reaction (10)) or sulphide (Reaction (11)) The reactions controlling the denitrification process Eq.
oxidation. In both cases, the decrease in NO 3 concentration (4) and (5) can be revealed, to some extent, by studying the
is coupled with an increase in the d15N and d18O of residual isotopic composition of the solutes involved in the reac-
NO 3 with a eN/eO ratio that ranges from 1.3 (Fukada et al., tions: (d15 NNO3 ; d18 ONO3 ; d13 CDIC ; d18 OH2 O ; d34 SSO4 and
2003) to 2.1 (Böttcher et al., 1990), (Fig. 5a). Intermediate d18 OSO4 ). Denitrification by oxidation of organic matter
enrichment ratios were reported by Mengis et al. (1999), (the d13C of one pig manure sample analyzed was
Cey et al. (1999), DeVito et al. (2000). Sorption and dilution 23.8‰, Vitòria, 2004, in agreement with values for d13C
processes produce no change in the d15N and d18O of NO 3, organic matter – between 20‰ and 30‰ – reported
or, if so, it is negligible (Clark and Fritz, 1997). by Trumbore and Druffel, 1995) should result in a decrease
of NO 13
3 and d CDIC together with a DIC increase. However,
4NO3 þ 5C þ 2H2 O ! 2N2 þ 4HCO3 þ CO2 ð4Þ
in the present case, these relationships were not observed
14NO3 þ 5FeS2 þ 4Hþ ! 7N2 þ 10SO2
4 þ 5Fe

þ 2H2 O (Fig. 7). Instead, the measured values were in the range of
ð5Þ d13 CDIC in groundwater – within Vogel’s range of 16‰ to
11‰ (Vogel and Ehhalt, 1963)—, possibly resulting from
The Osona samples show a positive trend in a d15 NNO3 vs equilibrium between soil CO2 and carbonate dissolution
d18 ONO3 diagram (Fig. 5a), in accordance with denitrifica- (0‰; Urquiola, 1994). Because the d13C of the DIC in a
tion, thus ruling out that 15N variations are only due to carbonated aquifer is buffered, no relationship was ob-
NHþ 15
4 volatilization. A linear relationship between d NNO3 served. Although the organic matter could be playing a role
and d18 ONO3 as reported in previous studies was not ob- in the denitrification process, this can neither be confirmed
served, neither was there a negative correlation between nor discounted with the available information. Detailed
d15 NNO3 and NO 3 concentration (Fig. 5b), because samples sampling using multipiezometers, or temporal sampling,
were a mix from the distinct aquifer levels, and the initial should be applied to give further insight.
concentration and the d15N in the recharge area could dif- Although thermodynamically the reaction of denitrifi-
fer depending on volatilization processes and rates of cation by organic matter oxidation is more favourable than
application. In order to confirm the existence of this natu- by pyrite oxidation, some authors (Pauwels et al., 2000;
ral attenuation process, d18 ONO3 is used as a tracer of deni- Postma et al., 1991) have found that the kinetics of these
trification. Although a comparison between d18 ONO3 and reactions facilitate the latter. In an aerobic media, S oxida-
NO 3 concentrations could not be carried out due to the ini- tion can take place by the decrease of the dissolved O2 in
tial variability of the NO3 concentration and the different water
application rates. A comparison between d18 ONO3 and the
NO

3 =Cl ratio shows a negative correlation for most sam-
4FeS2 þ 14O2 þ 4H2 O ! 4Fe2þ þ 8SO2
4 þ 8H
þ
ð6Þ
ples (Fig. 6). Those with higher d18 ONO3 show a lower
 and, in an anaerobic media, by the nitrate reduction (Reac-
NO 
3 =Cl ratio pointing to a relative decrease in NO3 com- tion (11). Because this last reaction consumes protons, it is
believed that it can not occur in carbonated aquifers,
Springs Springs Low d18O where pH is higher than 7 (Moncaster et al., 2000). How-
12 Wells Wells Low d18O ever, if the Fe2+ produced is oxidized, following Reaction
Aluvial

10 Springs Springs Low d18O


Wells Wells Low d18O
-10
Aluvial Low NO3
δ O-NO 3 ( ‰ )

8 -11

-12
δ C-DIC ( ‰ )

6 -13
18

-14
r 2 = 0.67
4 -15
13

-16
2
-17
0 1 2 3 4 5 6
- -
NO3 /Cl -18
200 250 300 35 400 450 500 550
 -
Fig. 6. NO
ratio vs. d18 ONO3 . A negative linear correlation is observed
3 =Cl
HCO3 (mg/l)
(outliers were not considered for the regression calculation), indicating
that the increase in d18 ONO3 is linked to denitrification (see explanation in Fig. 7. HCO 13 13
3 vs d CDIC . Samples are in the range of d CDIC in ground-
the text). water (16‰ to 11‰, Vogel and Ehhalt, 1963).
3608 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

(13), a global Reaction (14) can be written where protons d18 OH2 O diagram, where SO24 derived from sulphide oxida-
could locally overcome the ones consumed by carbonate tion must fall. The lower limit of this area was d18 OSO4 ¼
dissolution. If that is the case, denitrification by pyrite oxi- d18 OH2 O , assuming no enrichment during the incorporation
dation could occur in a carbonate aquifer. Based on the of O from H2O (eH2O = 0); the upper limit, using experimen-
studies of Bottrell et al. (2000) regarding the reduction of tal data of abiotic and biotic oxidation of sulphides, was de-
SO2
4 , and (Lawrence and Foster 1986) concerning the fined as d18 OSO4 ¼ 0:62  d18 OH2 O + 9. Fig. 8b shows the area
reduction of NO 3 , the catalyst bacterial activity would take defined by Van Stempvoort and Krouse (1994) together
place on the surface of carbonate fissures, as the pore with the analyzed samples, showing that SO2 4 is derived
diameters of the carbonates are too small to allow the col- from sulphide oxidation in most samples.
onization of the matrix aquifer Due to the absence of evaporite outcrops in the area,
positive S values could be explained by an anthropogenic
10Fe2þ þ NO3 þ 14H2 O ! 10FeOOH þ N2 þ 18H2 O ð7Þ
input. In this case, the most likely source is related to
6NO3 þ 2FeS2 þ 2H2 O ! 3N2 þ 2FeOOH þ 4SO2
4 þ 2H þ
agricultural activities: fertilisers or pig manure. In the
ð8Þ d34 SSO4 versus d18 OSO4 diagram (Fig. 8a), samples plot
over a mixing line between SO2 4 derived from pyrite oxi-
In the studied area, most samples had negative values of dation and pig manure SO2 4 . Local pig manure samples
d34 SSO4 (Table 3), matching the composition of pyrite dis- had d34 SSO4 between 0% and +5% (Otero et al., 2007), in
seminated in the aquifer materials (d34 SFeS2 between 5‰ the range of literature values (0.9‰ to 5.8‰, Cravotta,
and 28‰; Viñals et al., 2002; Pierre et al. 1994). Further- 1997). To the authors’ knowledge, no data reports in
more, SO2 4 formed via sulphide oxidation must show a the literature are available for d18O of pig manure dis-
relationship between d18 OSO4 and d18 OH2 O , as water is the solved SO2 4 , except for one, +5.6‰, offered by Otero
source of 50–100% of the SO2 4 oxygen. Van Stempvoort et al. (2007). The relationship between sulphide oxida-
and Krouse (1994) defined the area, in a d18 OSO4 vs. tion and denitrification is evidenced in a d34 SSO4 vs.

a 10
Springs
Springs Low d18O
8 Wells
Wells Low d18O
Aluvial
6
δ O-SO4 ( ‰ )

Low NO3
Pig manure
Pig manure
4
18

0 Sulphates derived from


sulphide oxidation

-2
-22 -18 -14 -10 -6 -2 2 6 10
34
S - SO4(‰ )

16
b Springs
14 Springs Low d18O
Wells
12
Wells Low d18O
10 Aluvial
δ O-SO4 ( ‰ )

8 Low NO3

4
18

Experimental area of
2 sulphates derived from
sulphide oxidation
0

-2

-4
-12 -9 -6 -3 0
18
O - H2O (‰ )

Fig. 8. (a) d34 SSO4 vs. d18 OSO4 of dissolved SO4. The box represents the SO4 derived from sulphide oxidation, and the shadowed area the isotopic composition
of manure (Vitòria et al, 2004a; Cravotta, 1997, and references therein) (b) d18 OH2 O vs d18 OSO4 diagram. The experimental area of SO4 derived from sulphide
oxidation is defined by Van Stempvoort and Krouse (1994). See explanation in the text.
L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611 3609

Springs Springs Low d18O groundwater flow line, by means of a detailed temporal
Wells Wells Low d18O
study, or by using a comprehensive multipiezometer
8 Aluvial
network.
denitrification
4 not linked to To predict the future evolution of contamination, it is
pyrite oxidation important to determine the enrichment factors in the
0 study area, the reactions controlling the denitrification
process, and the kinetics of these reactions. Moreover,
δ S-SO4 ( ‰ )

-4
knowing the quantity of pyrite in the aquifer materials,
-8 its availability to react through porosity, and its oxidizing
kinetics, should allow considering the possibility of in-
34

-12
duced remediation techniques. Recent studies on chemi-
-16
pyrite oxidation cal mediums in permeable reactive barriers for
not linked to
detnirification groundwater nitrate remediation have shown that zero-
-20 valent Fe can produce NHþ 4 ions under laboratory condi-

-24
tions (Su and Puls, 2004). The present study showed the
2 4 6 8 10 12 potential use of pyrite for permeable reactive barriers, at
18
O - NO3( ‰) sites where the natural attenuation is limited. Future re-
search could focus on laboratory studies in order to
Fig. 9. d18 ONO3 vs. d34 SSO4 diagram. Samples with lower d18 ONO3 show determine the kinetics of pyrite dissolution in the deni-
positive d34 SSO4 values, and samples with higher d18 ONO3 show a trend to
trification process.
negative d34 SSO4 values, confirming the link between pyrite oxidation and
denitrification.

5. Conclusions
d18 ONO3 diagram (Fig. 9). Samples with higher d18 ONO3 ,
indicating a higher degree of denitrification, have lower In Osona, the use of pig manure as an organic fertiliser
d34 SSO4 . It is worth noting that two samples have has caused diffuse NO 3 pollution of groundwater with
d34 SSO4 below 12‰, and low d18 ONO3 indicating that, NO 3 concentration up to 366 mg/L. Nearly 75% of the
for these samples, sulphide oxidation is not linked to sampled groundwater had NO 3 contents that exceed the
denitrification. Finally there is one sample showing high 50 mg/L threshold for drinking water. The coupled use
d18 ONO3 , and positive d34S, this sample is from an alluvial of d15 NNO3 and d18 ONO3 confirmed the link between
aquifer. In this case, the denitrification reaction would be groundwater NO 3 pollution and pig manure. While most
related to organic matter oxidation, though using DIC samples had isotopic values in the range of pig manure
and d13C this could not be confirmed nor discounted. NO 3 one sample had a clear signature of synthetic fertilis-
A quantification of the attenuation could be done by ers and 4 showed intermediate values, suggesting a mix
taking into account that the denitrification reaction de- between these two sources. The study of d18O in nitrates,
scribes a Rayleigh distillation process (Reaction (9)) where groundwater and rain water has enabled evaluation of the
e is the isotopic enrichment factor that depends on the nitrification process and identification that there was a
aquifer materials and characteristics. pre-existing enrichment in 18O of the soil O2 and H2O in
the area, resulting from bacterial respiration and evapora-
d15 Nresidual ¼ d15 Ninitial þ e lnð½NO3 residual =½NO3 initial Þ ð9Þ
tion. Nitrate isotopic data suggested that denitrification
Lehmann et al. (2003) presented a summary of the pub- processes were taking place. Moreover, the occurrence
lished enrichment factors for both N and O in denitrifica- of this process was confirmed using d18 ONO3 coupled with

tion processes in experiments with pure culture, the NO 3 =Cl ratio, thus avoiding the influence of continu-

groundwater, and a marine environment. Values for the ous NO3 inputs. A further insight into denitrification pro-
eN ranged from 0‰ to 40‰, although in groundwater cesses was obtained using a multi-isotopic approach,
studies the range is narrower, between 4.7‰ and analyzing the ions involved in denitrification reactions
30‰. Only two values of eO are reported: 8‰ (Böttcher, (d15 NNO3 ; d18 ONO3 ; d34 SSO4 ; d18 OSO4 ; d18 OH2 O and d13 CDIC ). This
1990) and 18.3‰ (Mengis et al., 1999). approach showed a link between denitrification and pyr-
In the studied case, Eq. (9) cannot be applied, because ite oxidation; however, the role of organic matter oxida-
samples are a mix between different aquifer levels and a tion could not be confirmed nor discarded with the
flow path could not be determined. Therefore, only a rough available information. In order to quantify the natural
estimation of the degree of attenuation could be calcu- attenuation, to foresee the evolution of the contamina-
lated. For example, taking an initial d15N composition be- tion, and to implement induced remediation techniques,
tween +8‰ and +15‰ for groundwater contaminated by it is necessary to determine the enrichment factor for
pig manure (Vitòria et al., 2004a), and using the extreme the aquifer and perform detailed studies on pyrite avail-
enrichment factors from the literature, the natural attenu- ability and its oxidation kinetics. To conclude, it should
ation of contamination could be estimated varying be- be emphasized that in areas where aquifers are complex,
tween 20–80% for an e of 4.7‰ (Mariotti et al., 1988) or and with no good infrastructure (e.g. multi-piezometers)
between 5–25% for an e of 30‰ (Vogel et al., 1981). The to describe groundwater dynamics, isotopes provide
real enrichment factor for the Osona aquifer should be ob- effective tools to trace sources of contamination and
tained by studying the chemical and isotope data from a study processes affecting N.
3610 L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611

Acknowledgements Hollocher, T.C., 1984. Source of the oxygen atoms of nitrate in the
oxidation of nitrite by Nitrobacter agilis and evidence against a P–O–N
anhydride mechanism in oxidative phosphorylation. Arch. Biochem.
This study was financed by the CICYT, project REN-2002- Biophys. 233, 721–727.
04288 and CGL2005-08019-CO4-CO1/HID, granted by the Hsieh, J.C.C., Chadwick, O.A., Kelly, E.F., Savin, S.M., 1998. Oxygen isotopic
composition of soil water: quantifying evaporation and transpiration.
Spanish Government, and partially by the -SGR2005-0933
Geoderma 82, 269–293.
from the Catalan Government. We would like to thank the IAEA/WMO, 2004. Global Network of Isotopes in Precipitation. The GNIP
Agencia Catalana de l’Aigua (Catalan Water Agency) for Database. [on-line]. Available at: <http://isohis.iaea.org>. (Accessed:
October 2004).
their support (especially A. Ginebreda) and the Serveis
IDESCAT, 1999. Cens agrari del Banc d’estadístiques de municipis i
Cientificotècnics of the Barcelona University (Spain) for comarques [on line]. Institut d’Estadística de Catalunya. Available
their analytical support. from: <http://www.idescat.es>.
Kendall, C., McDonnell, J.J., 1998. Tracing nitrogen sources and cycling in
catchments. In: Kendall, C., McDonnell, J.J. (Eds.), Isotope Tracers in
References Catchment Hydrology. Elsevier Science, BV, Amsterdam.
Kornexl, B., Gehre, M., Höfling, R., Werner, R., 1999. On-line d18O
Altman, S.J., Parizek, R.R., 1995. Dilution of nonpoint-source nitrate in measurement of organic and inorganic substances. Rapid Commun.
ground water. J. Environ. Qual. 24, 707–718. Mass Spectrom. 13, 1685–1693.
Anderson, K.K., Hooper, A.B., 1983. O2 and H2O are each the source of one Lawrence, A.R., Foster, S.S.D., 1986. Denitrification of a limestone aquifer
O in NO2 produced from NH3 by Nitrosomonas:15N evidence. FEBS in relation to the security of low-nitrate water suppliers. J. Inst. Water
Lett. 164, 236–240. Eng. Sci. 40, 159–172.
Aravena, R., Robertson, W.D., 1998. Use of multiple isotope tracers to Lehmann, M.F., Reichert, P., Bernasconi, S.M., Barbieri, A., McKenzie, J.A.,
evaluate denitrification in ground water: study of nitrate from a 2003. The stable isotope composition (d15N and d18O) of nitrate was
large-flux septic system plume. Ground Water 36, 975–981. measured during Summer 1999 in the anaerobic hypolimnion of
Bishop, P.K., 1990. Precipitation of dissolved carbonate species from eutrophic Lake Lugano (Switzerland). Geochim. Cosmochim. Acta 67,
natural waters for d13C analysis – Acritical appraisal. Chem. Geol. 2529–2542.
(Isotope Geosci. Section) 80, 251–259. Letolle, R., 1980. Nitrogen-15 in the natural environment. In: Fritz, P.,
Böttcher, J., Strebel, O., Voerkelius, S., Schmidt, H.-L., 1990. Using Fontes, J.C. (Eds.), Handbook of Environmental Isotope Geochemistry
isotope fractionation of nitrate-nitrogen and nitrate-oxygen for The Terrestrial Environment, vol. 1. Elsevier, Amsterdam, pp. 407–
evaluation of microbial denitrification in sandy aquifer. J. Hydrol. 433.
114, 413–424. Mariotti, A., Germon, J.C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, P.,
Bottrell, S.H., Moncaster, S.J., Tellam, J.H., Lloyd, J.W., Fisher, Q.J., Newton, 1981. Experimental determination of nitrogen kinetic isotope
R.J., 2000. Controls on bacterial sulfate reduction in dual porosity fractionation: some principles, illustration for the denitrification
aquifer systems: the Lincolnshire aquifer, England. Chem. Geol. 6, and nitrification processes. Plant Soil 62, 413–430.
319–333. Mariotti, A., Landreau, A., Simon, B., 1988. 15N isotope biogeochemistry
Cey, E., Rudolph, D., Aravena, R., Parkin, G., 1999. Role of the riparian zone and natural denitrification process in groundwater; application to the
in controlling the distribution and fate of agricultural nitrogen near a chalk aquifer of northern France. Geochim. Cosmochim. Acta 52,
small stream in southern Ontario. J. Contam. Hydrol. 37, 45–67. 1869–1878.
Clark, I.D., Fritz, P., 1997. Environmental Isotopes in Hydrogeology. Lewis Mayer, B., Bollwerk, S.M., Mansfeldt, T., Hütter, B., Veizer, J., 2001. The
Publishers, New York. oxygen isotope composition of nitrate generated by nitrification in
Cravotta, C.A., 1997. Use of Stable Isotopes of Carbon, Nitrogen, and Sulfur acid forest floors. Geochim. Cosmochim. Acta 65, 2743–2756.
to Identify Sources of Nitrogen in Surface Waters in the Lower Mengis, M., Schiff, S.L., Harris, M., English, M.C., Aravena, R., Elgood, R.J.,
Susquehanna River Basin, Pennsylvania. US Geol. Surv. Water-Supply MacLean, A., 1999. Multiple geochemical and isotopic approaches for
Paper 2497. assessing ground water NO 3 elimination in a riparian zone. Ground
DeVito, K.J., Fitzgerald, D., Hill, A.R., Aravena, R., 2000. Nitrate dynamics in Water 37, 448–457.
relation to lithology and hydrologic flow path in a river riparian zone. Mengis, M., Walther, U., Bernasconi, S., Wehrli, B., 2001. Limitations of
J. Environ. Qual. 29, 1075–1084. using d18O for the source identification of nitrate in agricultural soils.
Dogramaci, S.S., Herczeg, A.L., Schiff, S.L., Bone, Y., 2001. Controls on d34S Environ. Sci. Technol. 35, 1840–1844.
and d18O of dissolved SO4 in aquifers of the Murray Basin, Australia and Moncaster, S.J., Bottrell, S.H., Tellam, J.H., Lloyd, J.W., Honhauser, K.O.,
their use as indicators of flow processes. Appl. Geochem. 16, 475–488. 2000. Migration and attenuation of agrochemical pollutants: insights
EC, 1997. Treaty establishing the European Community (Amsterdam from isotopic analysis of groundwater sulphate. J. Contam. Hydrol. 43,
consolidated version) [on line]. Official Journal of the European 147–163.
Communities, C 340, of 10.11.1997, Brussels. Available from: <http:// Otero, N., Canals, A., Soler, A., 2007. Using dual-isotope data to trace
www.europa.eu.int/eur-lex>. the origin and processes of dissolved sulphate: a case study in
EC, 1998. Council Directive 98/83/EC, of 3 November 1998, on the quality Calders stream (Llobregat basin, Spain). Aquat. Geochem. 13, 109–
of water intended for human consumption [on line]. Official Journal of 126.
the European Communities, L 330, of 5.12.1998, Brussels. Available Otero, N., Vitòria, L., Soler, A., Canals, A., 2005. Fertilizer characterization:
from: <http://www.europa.eu.int/eur-lex>. major, trace and rare earth elements. Appl. Geochem. 20, 1473–
EEC, 1991. Council Directive 91/676/EEC, of 12 December 1991, 1488.
concerning the protection of waters against pollution caused by Pauwels, H., Foucher, J.-C., Kloppmann, W., 2000. Denitrification and
nitrates from agricultural sources [on line]. Official Journal of the mixing in a schist aquifer: Influence on water chemistry isotopes.
European Communities, L 75, of 31.12.1991, Brussels. Available from: Chem. Geol. 168, 307–324.
<http://www.europa.eu.int/eur-lex>. Pierre, C., Taberner, C., Urquiola, M.M., Pueyo, J.J., 1994. Sulphur and
Epstein, S., Mayeda, T.K., 1953. Variation of 18O content of waters from oxygen isotope composition of sulphates in hypersaline
natural sources. Geochim. Cosmochim. Acta 4, 213–224. environments, as markers of redox depositional versus diagenetic
Fukada, T., Hiscock, K., Dennis, P.F., Grischek, T., 2003. A dual isotope changes. Mineral. Mag. 58 (A), 724–725.
approach to identify denitrification in groundwater at a river-bank Postma, D., Boesen, C., Kristiansen, H., Larsen, F., 1991. Nitrate reduction
infiltration site. Water Res. 37, 3070–3078. in an unconfined sandy aquifer: water chemistry, reduction
Gat, J.R., 1996. Oxygen and hydrogen isotopes in the hydrologic cycle. processes, and geochemical modeling. Water Resour. Res. 27, 2027–
Ann. Rev. Earth Planet. Sci. 24, 225–262. 2045.
Grischek, T., Hiscock, K.M., Metschies, T., Dennis, P.F., Nestler, W., 1998. Reguant, S., 1967. El Eoceno marino de la plana de Vic (Barcelona) –
Factors affecting denitrification during infiltration of river water into a Investigaciones estratigráficas en el borde meridional de la depresión
sand and gravel aquifer in Saxony, Germany. Water Res. 32, 450–460. del Ebro. Memoria del Instituto Geológico y Minero de España. Tomo
Guy, R.D., Foge, M.L., Berry, J.A., 1993. Photosynthetic fractionation of LXVIII, Madrid.
stable isotopes of oxygen and carbon. Plant Physiol. 101, 37–47. Silva, S.R., Kendall, C., Wilkinson, D.H., Zieglerc, A.C., Chang, C.C.Y.,
Heaton, T.H.E., 1986. Isotopic studies of nitrogen pollution in the Avanzino, R.J., 2000. A new method for collection of nitrate from
hydrosphere and atmosphere: a review. Chem. Geol. 59, 87– fresh water and the analysis of nitrogen and oxygen isotope ratios. J.
102. Hydrol. 228, 22–36.
L. Vitòria et al. / Applied Geochemistry 23 (2008) 3597–3611 3611

Su, C., Puls, R.W., 2004. Nitrate reduction by zerovalent iron: effects of dD and 87Sr/86Sr) of nitrate polluted groundwater from agricultural
formate, oxalate, citrate, chloride, sulfate, borate, and phosphate. and livestock sources. PhD Thesis. Univ. Barcelona.
Environ. Sci. Technol. 38, 2715–2720. Vitòria, L., Grandia, F., Soler, A., 2004a. Evolution of the chemical (NH4)
Trumbore, S.E., Druffel, E.R.M., 1995. Carbon isotopes for characterizing and isotopic (d15N–NH4) composition of pig manure stored in an
sources and turnover of nonliving organic matter. In: Zepp, R.G., experimental pit. In: International Atomic Energy Agency (Ed.),
Sonntag, C. (Eds.), The Role of Nonliving Organic Matter in the Earth’s Isotope Hydrology and Integrated Water Resources Management.
Carbon Cycle. John Wiley & Sons, New York. Conf. Symp. Papers, Vienna, pp. 260–261.
Urquiola, M.M., 1994. Aplicación de indicadores geoquímicos al estudio Vitòria, L., Otero, N., Canals, A., Soler, A., tul=0?>Vitòria et al., 2004 b.
de las facies anóxicas del Eoceno Superior de la cuenca de antepaís Fertilizer characterization: isotopic data (N, S, O, C and Sr). Environ.
surpirenaica (sector oriental). Doctoral Thesis, Univ. Barcelona. Sci. Technol. 38, 3254–3262.
Van Stempvoort, D.R., Krouse, H.R., 1994. Controls of d18 O in sulphate. In: Vogel, J.C., Ehhalt, D.H., 1963. The use of carbon isotopes in groundwater
Alpers, C.N., Blowes, D.W. (Eds.), Environmental Geochemistry of studies. In: Proc. Conf. Isotopes in Hydrology, IAEA, Vienna, pp. 383–
Sulphide Oxidation. Am. Chem. Soc. Symp. Series, 550, pp. 446–480. 396.
Viñals, E., Canals, A., Soler, A., Teixidor, P., 2002. Influence of oil dump Vogel, J.C., Talma, A.S., Heaton, T.H.E., 1981. Gaseous nitrogen as evidence
remediation on d34S of dissolved sulphate in a Llobregat creek for denitrification in groundwater. J. Hydrol. 50, 191–200.
(Catalonia, NE Spain). In: European Society for Isotope Research Wassenaar, L.I., 1995. Evaluation of the origin and fate of nitrate in
(Ed.), VI. Isotope Workshop of the European Society of Isotope Abbotsford Aquifer using the isotopes of 15N and18O in NO 3 . Appl.
Research, Abstr. 122-123. Geochem. 10, 391–405.
Vitòria, L., 2004. Estudi multi-isotòpic (d15N, d34S, d13C, d18O, dD i87Sr/86Sr) Widory, D., Kloppmann, W., Chery, L., Bonnin, J., Rochdi, H., Guinamant, J-
de les aigües subterrànies contaminades per nitrats d’origen agrícola i L., 2004. Nitrate in groundwater: an isotopic multi-tracer approach. J.
ramader. Translated title: Multi-isotopic study (d15N, d34S, d13C, d18O, Contam. Hydrol. 72, 165–188.

You might also like