Está en la página 1de 20

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/225760509

Determination of Total Free Sulphides in


Sediment Porewater and Artefacts Related to
the Mobility of Mineral Sulphides

Article in Aquatic Geochemistry · November 2011


DOI: 10.1007/s10498-011-9137-0

CITATIONS READS

5 70

4 authors, including:

Kristina A. Brown Jay Thomas Cullen


Woods Hole Oceanographic Institution University of Victoria
16 PUBLICATIONS 264 CITATIONS 57 PUBLICATIONS 1,757 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Kristina A. Brown on 19 March 2015.

The user has requested enhancement of the downloaded file.


Aquat Geochem (2011) 17:821–839
DOI 10.1007/s10498-011-9137-0

ORIGINAL PAPER

Determination of Total Free Sulphides in Sediment


Porewater and Artefacts Related to the Mobility
of Mineral Sulphides

Kristina A. Brown • Eric R. McGreer • Bernie Taekema •

Jay T. Cullen

Received: 4 May 2011 / Accepted: 12 May 2011 / Published online: 3 June 2011
Ó Springer Science+Business Media B.V. 2011

Abstract A field and laboratory study of the accuracy of a method commonly used to
determine free sulphide concentrations in the porewater of marine sediments is presented.
The method uses an ion-selective electrode (ISE), sensitive to the sulphide ion (S2-), in
sediments buffered to high pH ([12) and is commonly used in regulatory monitoring
programs to assess the impacts of open net-pen finfish aquaculture on local marine habitats.
Here we report that on the timescale of field measurements, the accepted protocol can lead
to significant bias of free sulphide measurements, with orders of magnitude higher con-
centration detected in the buffered sediment–porewater slurry than in porewater samples
isolated and analysed separately. Laboratory experiments with model marine sediments
and analysis of sediment composition indicate that this bias is likely introduced by the
dissolution of particulate sulphides and/or sulphur present in the sediments under the
intense alkaline conditions of the protocol. Recommendations for the modification and
continued use of this commonly applied field methodology are discussed.

Keywords Free sulphide  Aquaculture  Environmental monitoring  Method  Toxicity

1 Introduction

Finfish aquaculture in British Columbia (BC) is responsible for over 50% of Canada’s total
production of farmed fish, a country that is the fourth leading producer of farmed salmon
globally (Auditor General 2004; Statistics Canada 2007). Aquaculture facilities in BC’s
coastal waters operate predominantly using an open net-pen system, consisting of

K. A. Brown  J. T. Cullen (&)


University of Victoria, School of Earth and Ocean Sciences, Bob Wright Centre, Victoria, BC, Canada
e-mail: jcullen@uvic.ca
K. A. Brown
e-mail: kbrown@eos.ubc.ca

E. R. McGreer  B. Taekema
Environmental Protection Division, British Columbia Ministry of Environment,
Nanaimo, BC V9T 6J9, Canada

123
822 Aquat Geochem (2011) 17:821–839

submerged pen arrays of approximately 30 9 30 m in size and 15 m depth and produce in


the order of 25,000 metric ton of fish during each 18–24-month production cycle (B.
Taekema, unpublished data). The dense population of fish in net enclosures results in an
increased flux of organic material, in the form of waste feed and semisolid faeces, to the
underlying sediments (Brooks 2001a, b). Some estimates suggest that 0.094 kg of solid
organic material is introduced to the water column per kg of fish produced (Nash 2001;
Brooks 2001a, b). Scaled to British Columbia’s 2006 salmon production numbers, the
industry is responsible for an allochthonous flux of over 6,500 metric tonnes of organic
material to the coastal environment per annum (Statistics Canada 2007).
As organic matter settles to the ocean floor, heterotrophs consume a portion of the
material aerobically, resulting in a depletion of dissolved oxygen within the sediment
porewater and at the sediment–water interface. As oxygen becomes scarce, further
remineralization continues anaerobically through the use of alternative electron acceptors
including nitrate and sulphate (Brooks and Mahnken 2003). Under these conditions,
bacterial sulphate reduction becomes quantitatively important in the sediments and results
in the formation of hydrogen sulphide (H2S) according to Eq. (1).
2CH2 O þ H2 SO4 ! 2 CO2 þ 2H2 O þ H2 S ðFonselius 1983Þ ð1Þ
At micromolar concentrations, the presence of sulphide in sediment porewater can be toxic
to many macroscopic marine organisms (Wang and Chapman 1999; Bagarinao 1992),
ultimately interfering with the function of the electron transport chain and other enzyme
functions (Nicholls 1975; Bagarinao 1992 and references therein). As H2S concentrations
increase in sediment porewater, its toxic properties elicit a change in the benthic biological
community, with many organisms experiencing acute toxicity. The resulting changes in
macrobenthic community structure have been widely used as an indicator of the physi-
cochemical changes occurring in the sediment and have thus been incorporated into the
aquaculture monitoring programs of many private and government organizations (Brooks
2001b; Hargrave et al. 2008 and references therein). Brooks (2001a, b) demonstrated a
correlation between total volatile solids (TVS), redox potential, sediment free sulphides
and changes in macrobenthic community structure. Specifically, statistical measures of
taxa diversity within the net cage array footprint and at reference stations were observed to
correlate well with both free sulphides and redox potential of sediments (Brooks 2001a, b).
Subsequent work illustrated that the accumulation of sediment-free sulphides and negative
shifts in redox potential correlated significantly with salmon farm biomass, feeding rates
and the deposition rates of TVS beneath net pens (Brooks and Mahnken 2003), indicating
that relatively straightforward, direct measurements of the sediment physical–chemical
conditions could be used as cost-effective tools to monitor the impacts on the benthic
biological community.
British Columbia’s Ministry of Environment (MOE) currently utilizes the measurement
of free sulphide or ‘sulphide ions not chemically bound to any other chemical constituent’
(BCMOE 2002a), within sediment porewater, as part of a larger monitoring framework to
determine aquaculture facility compliance with the Finfish Aquaculture Waste Control
Regulation (FAWCR BCMOE 2002a) portion of the provincial Environmental Manage-
ment Act (BCMOE 2003). The existing protocol for measuring free sulphide was devel-
oped to maximize accuracy and precision while using commercially available, portable
instruments whose ease of calibration and operation would allow non-professionals to
determine free sulphide efficiently in the field. The measurement of porewater free sul-
phide is carried out by mixing a sample of surface sediments collected from the sea floor in

123
Aquat Geochem (2011) 17:821–839 823

the vicinity of aquaculture enclosures with a sulphide antioxidant buffer (SAOB) (a


solution composed of 20.0 g sodium hydroxide, 17.9 g EDTA and 8.75 g L-ascorbic acid
per litre) in equal parts, homogenizing the slurry with a spatula and measuring the S2- ions
present with a standard silver/sulphide ion-selective electrode (ISE), according to the
Protocols for Marine Environmental Monitoring (PMEM, BCMOE 2002b). This protocol,
or slightly modified versions, for determining free sulphide is widely used by government
and the aquaculture industry as a component of aquaculture monitoring programs (e.g.
Wildish et al. 1999; Brooks 2001a, b; Sutherland et al. 2007; Hargrave et al. 2008). The
ISE used (typically a Thermo Orion 96-16 ionplus) operates with a silver/sulphide sensing
element; when the element is in contact with a solution containing sulphide ions, an
electrode potential develops across the sensing element and is measured against a constant
reference potential. This measured potential (corresponding to the amount of the measured
ion in solution) is described by the Nernst equation:
E ¼ Eo þ S log A ð2Þ
where E = measured electrode potential
Eo = reference potential.
A = ion activity (silver or sulphide) in solution.
S = electrode slope determined by standard calibration:
RT
S ¼ 2:3 ð3Þ
nF
where: R and F are constants
T = temperature in degrees K.
n = ionic charge.
The ion activity (A) is related to free-ion concentration through multiplying the free ion
concentration (Cf) by its activity coefficient (c), where A = cCf. The ionic activity
coefficient depends largely on the total ionic strength of the solution and can be highly
variable between solutions (Morel and Hering 1993). In order to keep the activity
coefficient constant, the background ionic strength must be kept high and constant relative
to the sulphide ion concentration; this ensures the activity of the ion is directly proportional
to its concentration. The addition of SAOB to the sample regulates the ionic strength of the
solution, slows the oxidation of free sulphide and maintains alkaline solution conditions for
optimal ISE function (Hseu et al. 1968; Glaister et al. 1985). The electrode slope used in
the Nernst equation is determined by calibrating with a set of external sulphide standards
(BCMOE 2002b).
Since the ISE detects sulphide (S2-), the SAOB addition to the sediment slurry is
required not only to maintain the ionic strength of the solution but also to ensure that a
larger proportion of hydrogen sulphide species present in the solution are in a form of S2-.
Experimental determinations of the second dissociation constant of hydrogen sulphide (K2)
yield a value of pK2 % 17±1 (pK2 = -log K2) (Meyer et al. 1983; Migdisov et al. 2002).
Adjusting the pH to [13 in a precise manner can allow the use of a single instrument to
determine the approximate total free sulphide present in the sediment porewater.
In practice, commercial and government laboratories report that the reproducibility of
total free sulphide measurements in aqueous standards (standard error ± 12%) tends to be
markedly better than measurements in natural marine sediments containing high sulphide
levels (standard error ± 55%, data not shown). The reasons for increased error in

123
824 Aquat Geochem (2011) 17:821–839

measurements in the presence of marine sediments may reflect natural spatial variability in
sediment total free sulphide, variability in subsampling of surface sediments or analytical
artefacts related to the presence of sediments during processing and measurement.
Here, we report the results of a study combining field and laboratory experiments
designed to investigate whether mineral sulphides and/or elemental sulphur in marine
sediments are mobilized and contribute to total free sulphide concentrations when mea-
sured using the existing MOE protocol.

2 Materials and Procedures

2.1 Sample Collection and Processing

2.1.1 Aquaculture Site Samples

Sites were accessed using the MOE R/V Grizzly Coast, and samples were collected on day
trips at Cyrus Rocks (28–29 June 2005: Lat 50°15.374, Lon 125°12.686) and Thorpe Point
(12–14 July 2006: Lat 50°34.792, Lon 127°36.504). Both sites were in fallow after a
rotation of production and were sampled according to the PMEM protocols for soft bottom
sediments (BCMOE 2002b). As outlined in the PMEM, sediment sampling within the
footprint of the salmon farm was carried out by taking samples of soft bottom sediment
using a Van Veen type grab. Samples were collected at the edge of the net pen (0 m), 30 m
from the net pen, and at the perimeter of the farm tenure which is typically on the order of
100 m from the net pen.

2.1.2 Model Anoxic Marine Sediment Samples

Marine sediment samples were collected for use in laboratory experiments on 9 November
2005, in Saanich Inlet on the University of Victoria R/V John Strickland. Saanich Inlet is a
235-m-deep glacial fjord with restricted circulation and high primary productivity,
resulting in intermittently anoxic bottom waters at depths below 70–150 m (Mosher and
Moran 2001). A Van Veen grab of bottom sediments (220 m depth) was obtained (Lat
48°37.149, Lon 123°29.858). The sediments in the grab sample were black in colour, and
organoleptic observation indicates they were rich in H2S. Samples of the top 2 cm of the
grab were isolated immediately after the grab returned to the deck and stored in 50-ml
capped vials (polypropylene screw cap, conical bottom tubes, SARSTEDT) on dry ice.
Once returned to the laboratory on shore, samples were transferred to a -20°C freezer until
used in laboratory tests.

2.2 Instrumentation and Analysis

2.2.1 Total Free Sulphide Determination

Sample total free sulphide concentrations reported here were determined using two inde-
pendent methods. Field measurements were carried out according to the detailed protocol
of the PMEM. Briefly, a 50-ml subsample of the upper 2 cm of the sediment grab is
homogenized in a low-density polyethylene (LDPE) beaker, and an equivalent volume of
SAOB is introduced and further homogenized with a stainless steel spatula. A Thermo
Orion 96-16 ion plus Silver/Sulphide ISE is then inserted and swirled until a stable reading

123
Aquat Geochem (2011) 17:821–839 825

is achieved, the unit flashes ‘Ready’ and the total free sulphide value is recorded. Standard
calibration was carried out according to the PMEM using previously made sulphide
standards provided by the MOE. Accuracy was generally ±9.7% for a 10 lM standard, and
precision of the ISE was found to be ±4.7% for replicate analyses of a natural sediment
sample with [H2S]T = 430 lM.
Laboratory measurements of total free sulphide were made using a flow injection
analysis (FIA) adaptation (Sarradin et al. 1999), of the methlylene blue method first
described by Cline (1969). A Rainin peristaltic pump (Rainin Dynamax RP-1, 8 channels)
was used to pump reagents through tygon tubing (Ismatech, ID) into various flow streams.
Flourinated ethylene-propylene (FEP, Cole-Palmer, ID 0.8 mm, OD 1.6 mm, wall
0.4 mm) tubing was used to make the mixing coil, plumb the injection valve (VICI, 10-port
injection valve) and to make tubing used to connect flow streams. Polyetheretherketone
(PEEK) fittings (Upchurch Flangeless fittings, SPE Ltd.) were used to connect FEP and
tygon tubing. The injection valve was electronically controlled through the WA Spec.3
Software (Waterville Analytical, ME USA). The reagent and carrier streams meet and mix
in a 2-m knotted reaction coil (Global FIA SSR-2, Fox Island WA, USA) and are carried to
the flow cell (Hellma Precision Cell, Quartz Suprasil window 176.700-QS, 10-mm path
length) to be measured by the detector. The light source used is a combination of an RF-
excited deuterium UV light source and a tungsten halogen VIS–NIR light source (DT-
mini-2-GS, Mikropack). The combination of two light sources into a single optical path
gives a spectral range of 200–1,100 nm. An Ocean Optics Spectrophotometer (USB2000),
with a wavelength range of 250-800 nm, detects the coloured product methylene blue as it
passes through the flow cell. The WA Spec.3 Software (Waterville Analytical, ME USA)
also controls the Ocean Optics Spec and logs data output.
Reagents were prepared according to the optimized methylene blue method developed
by Sarradin et al. (1999). All reagents were prepared in deionized water (18 MX cm) from
a Milli-Q analytical reagent-grade water purification system (Milli-Q Element; Millipore),
hereafter referred to as MQ water. The sulphide stock solution was standardized by id-
iometric titration.
A sulphide stock solution was prepared by dissolving sodium sulphide crystals in
deoxygenated (purged for [1 h per litre with ultra-high-purity (UHP) N2 gas) MQ water.
Since sulphide crystals were not stored under nitrogen, the stock solution concentration
was determined through slight modifications to the iodometric method described by
Clesceri et al. (1998) and Parsons et al. (1984). The stock solution was preserved with 2M
zinc acetate and refrigerated (Cassella et al. 1999). Standards used for the FIA system were
made via serial dilution of the stock solution. Standards were made at 10,000, 5,000, 2,500,
1,000, 500, 100, 50 and 25 lM. Precision of the method was found to be ±13% at 25 lM
which improved in standards [50 lM (typically \ ± 5%).

2.2.2 Reduction–Oxidation Potential (Eh)

Redox potential measurements were obtained with an Orion combination platinum redox
electrode (Orion Model 290A metre and 9678BN combination redox electrode), according
to the methodology outlined in the PMEM and by Wildish et al. (1999). The redox
electrode was filled with a 4M Ag/AgCl filling solution (Orion 900011) 24 h prior to use
and calibrated against the Orion ORP standard 967961. Potentials were measured in
sediment subsamples prior to the addition of SAOB, corrected for the potential of the
reference electrode and reported relative to the normal hydrogen electrode (NHE).

123
826 Aquat Geochem (2011) 17:821–839

2.2.3 pH Measurements

The pH of experimental treatments (below pH 12) was measured with an IQ120 pH meter
with silicon chip sensor (pH range 2.0–12.0, IQ Scientific Instruments). Experimental
treatments exceeding a pH of 12 were assessed using ColorpHast strips (pH range 0–14,
EMD Chemicals).

2.2.4 Scanning Electron Microscope-Energy Dispersive X-ray Analysis (SEM–EDS)

Marine sediments from a subset of laboratory experiments were analysed for qualitative
relative elemental composition using SEM–EDS (Hitachi S-3500N Scanning Electron
Microscope, Tungsten filament with an attached Oxford Instruments Link ISIS EDS X-Ray
Microanalysis System). Sediments were prepared for SEM–EDS observation as follows.
Sediments were first washed with distilled water and then dehydrated with varying
strengths of acetone (25, 50, 75, 100, 100 and 100%) at 25-min intervals and left overnight
in 100% acetone solution. Samples were then dried in porcelainized crucibles (cleaned
with laboratory-grade detergent, sonicated in distilled water for 10 min and rinsed with
isopropynol) at 100°C for 45 min. Dried samples were transferred to a greased capsule
(Dow Corning High Vacuum grease, silicon based) and Buehler Epoxide Resin, and Epoxy
Hardener was added at 5:1 w/w ratio (i.e. 15 g resin/ 3 g hardener). This mixture was
allowed to set for 4–6 h. Sample pucks were then extracted and sanded down to 5-mm
height with aluminium–diamond paste. The final polishing was completed by sonicating
pucks with distilled water for 3-5 min and washing them twice with ethanol.
The EDS can measure elements from B (5) to U (92), with a detection limit for all
elements of 1,500–2,000 ppm. Samples were placed in the system individually and viewed
in variable pressure mode (Backscatter) 20 kV, under 20-Pa vacuum, from a working
distance of *15 mm, and at 35 or 45 times magnification. Element spectra from several
areas of the sample were taken to get a good sense of the composition of the sample, and
then a representative a 3.6 9 2.7 mm area of the puck surface was chosen and mapped
with the EDS overnight (up to 17 h). The major elements silica (Si), aluminium (Al), iron
(Fe) and sulphur (S) were selected for the EDS mapping.

3 Results

3.1 Experimental Analysis

3.1.1 Field Experiment

Once Van Veen grab samples were deemed acceptable under the PMEM guidelines,
multiple subsamples of the upper 2 cm of undisturbed sediments were removed with a
stainless steel spatula for total free sulphide and Eh determinations. One set of duplicates
for free sulphide was processed by the MOE according to the PMEM, where samples were
homogenized and combined with equal parts of SAOB, mixed with a spatula, and then S2-
measured with a Thermo Orion 96-16 ion plus Silver/Sulphide ISE. A second set of
quadruplet samples was processed by centrifuging homogenized subsamples at
*4,000 rpm for 10 min. For one set of centrifuged duplicates, the supernatant/porewater
was immediately filtered (\0.45 lm). To control for potential loss of total free sulphide
through oxidation during centrifugation, the other duplicates were shaken to resuspend

123
Aquat Geochem (2011) 17:821–839 827

sediments. The filtered porewater samples and resuspended porewater/sediment mixtures


were then combined with equal parts SAOB, mixed, and S2- measured with a Thermo
Orion 96-16 Ion Plus Silver/Sulphide ISE, in accordance with the PMEM.

3.1.2 Laboratory Experiment

To test whether adjusting the pH of anoxic marine sediments to [13 mobilized metal
sulphide, several tubes of Saanich Inlet sediment samples were thawed overnight in a
generic refrigerator. Once thawed, porewater was separated from sediments by centri-
fuging at 4,000 rpm for 15 min (SORVALL Legend RT, MANDEL Scientific Co.) at 20°C
(room temperature); porewater was decanted, discarded, and 3.16 ± 0.02 g subsamples of
the sediment were measured into separate 50-ml conical bottom tubes. All subsamples for
each trial were taken from one sediment vial to minimize variability related to sediment
heterogeneity in the original Van Veen grab sample. To each tube containing a sediment
subsample, 50 ml of de-oxygenated salt water (30 ppt NaCl, purged with ultra high purity
N2 gas for 1 h) was added and the sediment was resuspended in solution by vigorous
mixing. Hereafter, this mixture will be called ‘model marine sediments’. A flow diagram
describing the experimental procedure is illustrated in Fig. 1.
The PMEM calls for an addition of SAOB to sediment in a 1:1 ratio (BCMOE 2002b),
testing this with the sediment collected from Saanich Inlet indicated a solution pH of *14
(ColorpHast 0-14, EMD Chemicals). In order to duplicate this shift in pH, a solution of 4N
NaOH (EMD Chemicals) was added to the model marine sediments until the pH was
brought up to an * pH = 14 (ColorpHast strips indication). This experiment was initially
attempted by combining model marine sediments with SAOB, following the procedure for
the use of the ISE, instead of the 4N NaOH used in this trial. This proved to be problematic
as the use of SAOB caused complications with the chemistry of the methylene blue
method. When tested, the porewater isolated from sulphide antioxidant-buffered model
marine sediment appeared to inhibit the reaction between DPD and the H2S present,
resulting in little colour change of the solution. Organoleptic observation concluded that
H2S was indeed being released by the sulphide antioxidant-buffered model marine sedi-
ments (and not the control sediments); however, our methods could not accurately quantify
this change, and further experimentation required that NaOH alone be used to adjust
sample pH.
Since the 1:1 volume ratio of SAOB to sediment is not always easily achieved, given
uncertainty in sediment porosity and the challenge of working in the field, an intermediate
of pH 10 was also tested in the experimental treatments (IQ120 pH meter with silicon chip
sensor, pH 2.0–12.0, IQ Scientific Instruments). In order to maintain the sediment-to-liquid
volume ratio between treatments, an appropriate volume of deoxygenated MQ water was
added to the control treatment and pH 10 treatments.
Once MQ water and/or NaOH were added to the model marine sediments (5-min
handling time), they were recapped, mixed again and let to sit for 10 min; this 15-min total
handling time was used to mimic the handling time of sediments in the field by MOE
which is typically not more than 10 min). At the end of the ‘handling time’ period, samples
were centrifuged for 10 min at 4,000 rpm (SORVALL Legend RT, MANDEL Scientific
Co.) to isolate the model marine sediment porewater. This porewater was then carefully
decanted into vials, so as not to resuspend sediment. Porewater for each model marine
sediment treatment was analysed for total free sulphides by the flow injection analysis
adaptation of Cline’s (1969) methylene blue method after Sarradin PM et al. (1999).

123
828 Aquat Geochem (2011) 17:821–839

Fig. 1 Flow diagram of the experimental procedure to determine the effect of pH on the measured
concentration of free sulphide in marine sediments

Standards and samples were analysed in triplicate (from the same container), and results
are reported as average absorbance units (integrated peak area). Standard deviations are
reported for three replicate injections from each container. H2S concentrations were
determined based on an average of three replicate samples from the same experimental
treatment vial and then calibrated by comparison with a standard curve run inline with
treatment samples ([H2S]T = 0.9005(abs) - 18.923, r2 = 0.9969).

3.2 Field Experiments

A comparison of total free sulphide determinations by the PMEM protocol, centrifuged and
resuspended sediments by PMEM protocol, and centrifuged and filtered porewater samples
from Cyrus Rocks are given in Table 1. Concentrations of total free sulphide detected

123
Aquat Geochem (2011) 17:821–839 829

Table 1 Comparison of porewater and sediment slurry sulphide measurements made with an ISE at Cyrus
Rocks BC
Stations Treatment [S2-] lM SD

Reference SE #1 PMEM 36.5 18


Reference SE #1 Centrifuged: resuspended 67.1 –
Reference SE #1 Centrifuged: porewater ND 0
Reference SE #2 PMEM 45.7 –
Reference SE #2 Centrifuged: resuspended 51.5 0.6
Reference SE #2 Centrifuged: porewater ND 0
Om SE PMEM 1,415 –
Om SE Centrifuged: resuspended 954 42
Om SE Centrifuged: porewater 1,039 439

using the PMEM for samples collected 0 m from the southeast (SE) edge of the net pen
were 1,415 lM and decreased to *40 lM at the reference station where the impact of
salmon waste on sediments was likely minimal (Table 1). Similar values were obtained for
samples that were first centrifuged to separate sediment and porewater, homogenized and
analysed following the PMEM. Samples treated in this manner had determined total free
sulphide concentrations of 954 ± 42 lM at 0 m SE from the net pen and approximately
60 lM at the reference station. When porewater was isolated by centrifugation and sub-
jected to the PMEM (with no sediment present), total free sulphide concentrations were
1,039 ± 439 lM at 0 m SE from the net pen and total free sulphide was not detected at the
reference station.
Observations from field experiments at the Thorpe Pt. aquaculture facility in July 2006
are summarized in Table 2. Grab samples were taken from stations along a transect from

Table 2 Comparison of total free sulphide measurements made by ISE in surface sediments and pore water
collected at Thorpe Pt, BC
Stations (Grab#) Total free sulphide (lM) Eh(NHE) (mV)

Sediment initial Sediment stabilized Porewater average SD

105 mE 18.3 18.3 ND – 107


90 mE 52.3 70.4 ND – 147.4
75 mE 278 459 ND – 28.3
60 mE 204 380 ND – 70.2
30 mE 538 1,130 237 75.7 -59.3
OmW 863 1,080 305 3.54 -181.3
15 mW 1,100 1,240 42 4.3 -162.2
30 mW (l) 719 889 ND – 91.2
30 mW (2) 592 851 ND – 63.2
60 mW (l) 158 223 ND – 373.2
60 mW (2) 258 270 ND – –
75 mW (l) 2.08 2.11 ND – 164.2
75 mW (2) 2.11 2.18 ND – –
ND not detected

123
830 Aquat Geochem (2011) 17:821–839

the net pen (0 mW) to the west (Station 0–75 mW) and then from the net pen to the east
(Station 0–105 mE). Stations are named for their distance and compass direction from the
aquaculture net pen. When homogenized sediment subsamples were measured in this trial, it
was observed that the ISE measurements increased with time in porewater when it was in
contact with the sediment slurry. The Sediment Initial Measurement is the S2- reading given
by the ISE when it first equilibrates with the slurry; however, it was often found that after
this reading, the measurement continued to climb until it reached a value where it became
stable, recorded as the Sediment Stabilized Measurement (Table 2). The current practice
with the existing method is to record the free sulphide reading after the initial stabilization.
Total free sulphide concentrations were highest in sediments collected nearest to the edge of
the net pen with stabilized measurements of 1,240 lM at station 15 m west of the edge and
1,080 lM at the western perimeter of the pen. Concentrations generally decreased with
distance from the net pen to minimum values of 2.11 lM 75 m to the west and 18.3 lM in
sediments collected 105 m to the east (Table 2). The second set of duplicate sub samples of
the sediment was taken and centrifuged down, separating porewater from the sediment.
Porewater was then combined with equal parts SAOB and measured with the same ISE,
giving the Porewater Measurement. In all cases, the porewater measurement provided a
stable measurement. Porewaters were found to have total free sulphide concentrations that
were lower than the sediment slurry measurements at all of the stations. Concentrations
were highest at the net pen edge (305 ± 3.54 lM) and dropped to 237 ± 75.7 lM 30 m to
the east and 42 ± 4.53 lM 15 m to the west of the pen edge. Total free sulphide was not
detected in porewater collected from stations greater than 15 m from the net pen edge and
was only detectable in samples with a corresponding negative Eh(NHE) measurement
(Table 2). Sediment Eh values (Table 2) were found to be lowest at stations nearest to the
net pen (0, 15 mW, 30 mE), indicating reducing conditions relative to NHE, and were
observed to increase with station distance from the net pen in both west and east directions.

3.3 Lab Experiments

Figure 2 reports the results of the sediment buffering experiment carried out on samples of
Saanich Inlet sediment. In all cases, natural, anoxic sediments where pH was increased
relative to the control were observed to have significantly higher H2S concentrations as
measured with the flow injection adapted methylene blue method. Control treatments
showed the lowest H2S concentration (24.0 ± 2.03 lM, n = 3), whereas treatments that
were adjusted to pH 10 had concentrations almost fourfold higher (81.4 ± 20.2 lM,
n = 4) than the control. Samples that were exposed to the typical PMEM pH of *14 had
total H2S concentrations of 282 ± 50.7 lM (n = 5), more than an order of magnitude
greater than the controls.

3.4 SEM–EDS

Results of the SEM–EDS elemental mapping (iron (Fe) and sulphur (S)) carried out on
control, and pH *14-treated sediments are reported in Fig. 3, and atomic percentages for
all major and minor elements are summarized in Table 3. Differences between atomic
percentages between the control and pH 14 treatment were less than 7.6% for all major
elements (C, O, Al, Si and Fe) with the exception of S which was 31% lower in the pH
14-treated sediments than the control (Table 3). The SEM–EDS images indicate that the
spatial distribution of S on individual sediment grains (Fig. 3b, d) was distinct from that of
the other major elements represented here by Fe (Fig. 3a, c). Spatial maps of Al, Si and Fe

123
Aquat Geochem (2011) 17:821–839 831

Fig. 2 Results of the sediment buffering experiment depicted in Fig. 1, testing the effect of pH on the
measured concentration of free sulphide in marine sediments. Error bars represent the standard deviation of
the replicates

for both control and pH 14 treatments indicated that these elements tend to be concentrated
in the centre of sediment grains and decrease in concentration towards the grain periphery
(Fig. 3a, c). However, the distribution of S is markedly different with highest concentra-
tions of the element in a distinct ring on the grain periphery (Fig. 3b, d). The atomic
percentage is noticeably lower for S in the pH 14 treatment compared with control, and the
visible rings on the SEM–EDS images are less distinct than in control treatments.

4 Discussion

BC MOE, along with other government agencies, aquaculture companies and its consul-
tants, utilizes the measurement of porewater free sulphide associated with marine sedi-
ments as a benchmark or environmental standard to evaluate the impact of finfish
aquaculture waste on local benthic habitat (e.g. Wildish et al. 1999; Brooks 2001a, b;
Sutherland et al. 2007; Hargrave et al. 2008). The rationale for this methodology is that at
sufficient porewater concentration, sulphides that are unbound and free to react at in situ
pH can be toxic to marine benthic organisms and thereby reduce species diversity at the
community level. The combination of laboratory and field investigations reported here
suggests that the current protocol used for handling and analysis of marine sediments can
lead to an artificially high measurement of porewater free sulphide at some sites depending
upon the geochemical characteristics of the natural sediments. We hypothesize that the
measurement bias is likely introduced through the mobilization of normally biologically
inert solid phase amorphous metal sulphides and native sulphur, which accumulate in
response to redox gradients in marine sediments receiving large fluxes of organic matter.

4.1 Mobility of Mineral Sulphides Under High pH

Acid volatile sulphides (AVS) are a well-known category of sedimentary sulphides derived
from the addition of a strong acid (1 N HCl) to sediments to yield hydrogen sulphide (H2S)

123
832 Aquat Geochem (2011) 17:821–839

Fig. 3 Sediment microstructure characterization using SEM–EDS X-ray mapping of Saanich Inlet
sediment samples from the pH experiment outlined in Fig. 1: a, b control treatment (359 magnification),
and c, d pH 14 treatment (459 magnification). The distribution and relative elemental concentrations of Fe
(panels a-control treatment and c-pH 14 treatment) and S (panels b-control treatment and d-pH 14 treatment)
can be visually distinguished as numerically equivalent grey levels in the image, with brighter grey levels
indicating higher relative concentration of the element

Table 3 Elemental composition


Elements Control atomic%a pH 14 atomic%a
of control and pH 14 samples as
determined by EDS–SEM
C 52.27 53.24
O 38.20 37.38
Si 5.85 5.41
Al 1.32 1.42
Fe 0.64 0.60
S 0.42 0.29
Cl 0.37 0.29
a
Atomic % = weight percent/ Na 0.34 0.59
atomic weight (the sum of all K 0.18 0.20
atomic weights in a sample is
normalized to 100%) calculated Mg 0.27 0.38
from 3 analyses of separate Ca 0.13 0.18
2.8 9 2.0 mm area from control Total 100 100
and pH 14 sample pucks

gas (Rickard and Morse 2005). The measurement of AVS in sediments is widely used to
infer the concentration of the pyrite precursor iron (II) monosulphide (a major component
of solid sulphides), as FeS is solubilized when combined with strong acid (Rickard and

123
Aquat Geochem (2011) 17:821–839 833

Morse 2005). Similarly, combining sediments with a strong base should result in the
competition of hydroxides (OH-) with the sulphide (S2-) complexed as authigenic metal
sulphides in the sediments, releasing S2- to solution and spuriously increasing the free
sulphide component of sediment porewaters, as illustrated in the FeS stability diagram
depicted in Fig. 4.
Experiments carried out in this investigation demonstrate that the practice of buffering
sediment–porewater slurries to pH to values C10 yields higher free sulphide concentrations
than found in controls or in samples where the porewater was isolated before the addition
of SAOB (Figs. 2, 3; Tables 2, 3). A reasonable explanation for this bias is that by
buffering sediments to a high pH using the SAOB mixture, particulate sulphide compo-
nents of the sediment likely become unstable, dissolve and contribute to the concentration
of total free sulphide measured by the ISE. To further assess how imposed alkaline con-
ditions affect sediment microstructure and elemental composition, samples of the model
sediments used in the buffering experiment (Fig. 1) were observed under a scanning
electron microscope with an attached energy-dispersive X-ray spectrophotometer (SEM–
EDS) to gain qualitative insight into the stoichiometry of the model sediments and to
determine whether significant mobility of solid phase sulphides and/or sulphur was
evident.

Fig. 4 Eh(NHE) versus pH diagram for the Fe–S system in seawater. Major ion activity calculated after
Whitfield (1974), temperature (10°C), pressure (1.013 bar), with [Fe]T = 10-6 M, for seawater at pH = 8.1
according to the total concentrations reported in Holland (1978). The red arrow indicates the trajectory of
the sediment slurry pH as modified by the addition of SAOB according to the PMEM (observed in this
study), and the red-highlighted area describes an area allowing for partial oxidation of the sediments during
handling and analysis. Following the trajectory of the red arrow, pyrite (FeS2) will undergo dissolution,
releasing S2- into solution as pH increases

123
834 Aquat Geochem (2011) 17:821–839

4.2 SEM–EDS Mapping

SEM–EDS mapping is a qualitative tool that can be used to characterize sediment


microstructure and determine elemental distribution of a sample. The elemental stoichi-
ometry of the model marine sediments allows a qualitative determination of mineral
composition and the identity of minerals likely to dissolve at high pH, resulting in the
release of sulphide ions to sediment porewaters.
The images for both the control and treated (pH *14) samples (Fig. 3) indicate that
while Al, Si, and Fe content vary, they are concentrated in the centre of sediment grains (Al
and Si data not shown). The clustering of Al, Si and Fe in the same area of the sediments
indicates that these sediments are predominantly composed of silicates and alumino-sili-
cates. This composition is consistent with previous studies that report Saanich Inlet sed-
iments have on average 69 wt% detrital material, 29 weight % opal and negligible
carbonate (Johnson and Grimm 2001).
Reduced S compounds, such as H2S, are produced through SO42- reduction within the
sediments and will accumulate until porewaters become supersaturated with respect to
metal sulphide precipitates (see Billon et al. 2001; Rickard and Luther 2007). Iron is
released from reducible oxides/hydroxides and silicate minerals in the layers of sediment
where SO42- reduction is occurring to yield ferrous iron (Fe2?), which then reacts with
dissolved H2S to produce amorphous FeS and/or crystallized FeS (see Billon et al. 2001;
Rickard and Luther 2007). These precipitates would be found on the periphery of grains,
adsorbed to the surfaces available and could account for the higher concentrations of S
seen at the grain margins in Fig. 3b, d. Figure 3b illustrates the precipitation of Fe ? S
minerals at the surface of sediment grains, and the lower concentrations of S in Fig. 3d
imply that these Fe ? S mineral precipitates have undergone an amount of dissolution at
high pH, releasing reduced S and precipitating Fe oxyhydroxides (OH-) and oxides
(Fig. 4).
Iron and sulphur associations are a common observation in anoxic sediments, where
sulphide-rich zones at depth have been seen to include a variety of precipitated and
dissolved Fe and S species simultaneously (Rickard and Morse 2005). Work investigating
the AVS component of marine sediments has determined Fe ? S associations that con-
tribute to this measurement include dissolved substances (dissolved Fe, S species (HS-,
H2Saq, FeHS?) and their complexes (FeS clusters and FeS nanoparticles)); aqueous iron
sulphide clusters (FeSaq); metastable iron sulphides (Mackinawite, FeS and Greigite,
Fe3S4); and a fraction of the solid phase pyrite, FeS2, can also contribute depending on the
extraction methods used (Rickard and Morse 2005; Rickard and Luther 2007). These
compounds would all be susceptible to dissolution under strong alkaline conditions (Fig. 4)
and bias measurements of free sulphide to higher values in studies applying methodology
similar to the MOE protocol (e.g. Wildish et al. 1999; Brooks 2001a, b; Brooks and
Mahnken 2003; Brooks et al. 2004; Sutherland et al. 2007; Hargrave et al. 2008).
To illustrate this point further, we compare the relationship between a subset of marine
sediment redox potentials Eh(NHE) and free sulphide concentrations reported in the liter-
ature (Fig. 5). In sediments containing H2S, the Eh(NHE) is predicted to be controlled by the
overall reversible redox reaction relating the two electron reduction in rhombic sulphur
(S0) to sulphide (S2-) (Berner 1963; Whitfield 1969). The thermodynamic relationship
between Eh(NHE) and free sulphide according to Berner (1963) is plotted in Fig. 5a, b
assuming porewater pHs of 8 (lower bold dashed line) and 6 (upper dashed line). It is
important to point out that at highly negative or positive Eh(NHE) values, other redox
couples will control sediment redox potential (e.g. O2/H2O at positive Eh(NHE)) causing

123
Aquat Geochem (2011) 17:821–839 835

Fig. 5 a Eh versus total free sulphide concentration [S2-] as measured in field studies of organically loaded
sediments. Filled symbols illustrate measurement of sediment-free sulphide carried out with a 1:1 addition
of SAOB to subsampled sediments (Sutherland et al., 2007 (circles); this study (squares)). Open symbols
refer to samples where sediment-free sulphide was measured in porewaters at natural pH (Thamdrup et al.
1994 (triangles); Whitfield 1969 (circles), or where porewaters were combined with SAOB after being
removed from the sediment matrix (this study). Reference lines are derived from the Eh versus free sulphide
relationship defined by Hargrave et al. 2008 (solid line) and thermodynamic equilibria described by Berner
(1963) for pH 6 and pH 8 conditions (upper and lower dashed lines). b As in a except some results from
Whitfield’s (1969), field studies at natural sediment pH are excluded (open circles) to focus on a smaller
range of free sulphide concentrations. Data from the Thorpe Point aquaculture facility analysed in this study
are circled in red. Arrows indicate the difference between the measurements made on sediments following
the MOE protocol and on isolated porewaters in the presence of SAOB. Error bars represent the standard
error of replicate measurements

deviations from the S0/S2- thermodynamic relationship (Berner 1963). Data are plotted for
free sulphide measured by ISE in SAOB sediment–porewater slurries, according to the
PMEM (closed symbols), or at native pH in the sediments and porewaters mixed with
SAOB in the absence of sediments (open symbols). In this study, when free sulphide values
were compared with Eh(NHE) measurements from the same grab sample (Fig. 5b, solid and
open squares), a clear difference was observed between porewater ? SAOB and sedi-
ment ? SAOB mixtures. Free sulphide measured in the SAOB ? sediments samples
according to the PMEM was on average roughly sixfold higher than free sulphide deter-
mined in porewater ? SAOB mixtures at a given Eh(NHE). The relationship between
Eh(NHE) and free sulphide for studies measuring sediment-free sulphide at in situ pH
(Whitfield 1969 (open circles); Thamdrup et al. 1994 (open triangles)) tends to follow the
thermodynamic relationship of Berner (1963). Indeed, at positive values of Eh(NHE), free
sulphide was usually not detected and the maximum free sulphide measured at Eh(NHE) [ 0
(21 ± 19 mV) was only 1.4 lM (Thamdrup et al. 1994). In contrast, measurements made
with the MOE protocol where sediments are exposed to high pH SAOB, and free sulphide
is measured with the ISE (e.g. Sutherland et al. 2007 (solid circles); Hargrave et al. 2008
(solid regression line)), tend to have higher free sulphide concentrations for a given
Eh(NHE) and lie significantly above the predicted thermodynamic relationship of Berner
(1963) as sediment Eh(NHE) increases. Free sulphide concentrations as high as 98 lM were
reported in sediments with highly oxidizing Eh(NHE) of [ ?300 mV (Sutherland et al.
2007; Fig. 5b). The slope of the regression line describing the Eh(NHE) versus free sulphide
relationship for sediments in the Bay of Fundy and coastal Maine (Fig. 5 solid line in a and
b; Hargrave et al. 1997; Wildish et al. 1999; Hargrave et al. 2008) has a slope that is
significantly more negative than the Berner (1963) relationship predicting free sulphide

123
836 Aquat Geochem (2011) 17:821–839

concentrations [1,600 lM at Eh(NHE) = 0 and [50 lM at oxidizing Eh(NHE) = ?200.


Taken together, this comparison suggests that on the timescale of measurements made with
MOE protocol (e.g. Wildish et al. 1999; Brooks 2001a, b; Sutherland et al. 2007; Hargrave
et al. 2008), there is a significant overestimate of free sulphide present in marine sediment
porewater, despite precautions taken by the investigators, likely from the dissolution of
solid phase metal sulphides and potentially elemental sulphur in the presence of the high-
pH SAOB.

5 Conclusions and Recommendations

The practice of measuring free sulphide with an ISE in sulphide antioxidant-buffered


sediments (pH [10) is inconsistent with the accurate determination of the sediment
porewater-free sulphide defined as ‘sulphide ions not chemically bound to any other
chemical constituent’ (BCMOE 2002a). This bias does not necessarily invalidate the use of
empirical relationships between sulphide determined by the existing method and measures
of benthic invertebrate taxa richness and diversity that are applied as a proxy for alterations
in benthic habitat proximal to fish farms in response to organic matter loading (e.g. Brooks
2001a, b). However, modifications to the protocol to minimize or eliminate the observed
bias would likely improve the accuracy and precision of measurements thus strengthening
the predictive power of empirical relationships and allow for direct comparison of field
measurements of free sulphide with laboratory studies that report the toxicity of free
sulphide to marine organisms. Here, we recommend two courses of action to address the
issue of metal sulphide dissolution and eliminate or minimize the potential for such dis-
solution to bias measurements of free sulphide in marine sediments: (1) to obviate the use
of SAOB by conducting measurements of sediment-free sulphide at natural pH, using
paired measurement of S2- and pH; (2) isolation of porewater from the collected sedi-
ments, followed by immediate analysis with the current MOE protocol or preservation with
zinc acetate for later analysis in the laboratory. Although each of these options requires
some modification of the current measurement protocol, both allow for more accurate
measurements of sediment-free sulphide, while allowing the majority of the measurement
protocol to remain intact. The recommendations endeavour to strike a balance between
maximizing accuracy and precision of the measurements and minimizing instrumental and
protocol complexity given that non-professionals will be expected to collect samples and
make measurements in remote field locations from small platforms.
The experiments presented in this paper indicate that the addition of SAOB to sediments
results in the solubilization of S2- previously bound to metals in the sediment. This effect,
but not the likely cause, was noted in earlier investigations by Wildish et al. (1999), who
tested the effects of different treatments and storage times on porewater sulphide con-
centrations. Paired pH measurements would obviate the need for the addition of SAOB,
and indeed, early investigations into H2S chemistry in marine sediments prescribed the
determination of S2- with a silver/sulphide electrode and pH simultaneously (Berner 1963;
Whitfield 1969; Allam et al. 1972). The activity of the S2- ion in solution is then calculated
from derived equilibrium equations (Table 4), and the potential of the silver–silver sul-
phide electrode (ES2-) can be plotted against pS2- of standardized solutions to provide a
calibration curve (Berner 1963; Allam et al. 1972). The benefit of this method is that the
manipulation and measurement of sediments are carried out at natural sediment pH,
yielding free sulphide concentrations which are the representative of the biologically

123
Aquat Geochem (2011) 17:821–839 837

Table 4 Calculation of sediment free sulphide at in situ pH


Investigator Equilibrium equations Measured parameters

Berner (1963) pS2- = 21.9–2pH pH, ES2-


Allam et al. (1972) pH2S = 2pH?pS2- -20.9 pH, ES2-

available fraction independent of the potential bias of solid-phase metal sulphide disso-
lution at high pH.
In both Berner (1963) and Allam et al. (1972) methods, solution (and sediment) pH is
maintained at natural levels. While the total free sulphide present in a closed system is
essentially constant, the free sulphide ion activity is controlled by the pH-dependent
equilibrium of the H2S system (Latimer 1952). As the potential of the silver–sulphide
electrode is a function of the free sulphide ion activity, which is not a dominant ion below
pH 12, measuring the solution pH as well as the amount of S2- present allows the
determination of total [H2S] in sediment samples through the equilibrium equations
described by Berner (1963) and Allam et al. (1972) (Table 4). When compared with the
standard methylene blue method, the results of Allam et al. (1972) were highly correlated
(r2 = 0.9965), illustrating the accuracy of the approach. Advances in the construction of
microelectrodes that allow for the direct determination of free sulphide in marine sedi-
ments (Luther et al. 2001a) could prove quite useful for monitoring purposes in the
aquaculture industry. Indeed, microelectrodes have been applied with great success in
more extreme marine environments, including hydrothermal vents (Luther et al. 2001b),
and could find similar success in the marine sediments underlying aquaculture pens.
Another approach to minimize error associated with the free sulphide measurement is to
separate sediments from porewater before the introduction of SAOB. Centrifugation of the
collected sediment samples is likely the easiest and most field-adaptable method of sedi-
ment–porewater separation, but other methods have been widely applied to this problem
(for example Thamdrup et al. 1994 utilize a porewater squeezer technique within an anoxic
glove-bag). This measurement will yield a more accurate porewater total free sulphide
measurements by minimizing the potential contribution from the dissolution of mineral
sulphides in sediments. Note that filtration [0.45 lm will not exclude polysulfide nano-
particulates (Rickard and Luther 2007) that could still artificially raise free sulphide
determinations. Given the high pK2 of H2S (Meyer et al. 1983; Migdisov et al. 2002) and,
therefore, the pH dependence of S2- concentrations determined by ISE, the accuracy and
precision of total free sulphides determined by this method ultimately rely on a precisely
defined pH for SAOB–porewater mixtures. This method also provides the opportunity to
preserve centrifuged porewater samples for return to shore and measurement in the lab-
oratory (for example Cassella et al. 1999). Preserved samples could be analysed for
porewater-free sulphide with more accuracy and precision with a variety of methodologies
reducing some of the variability introduced through measurements with a field-adapted
ISE.

Acknowledgments The authors would like to thank the following people for their much-appreciated
assistance throughout the duration of this project: Kirsten White, the Ministry of Environment, Nanaimo
BC; field work assistance from Conservation Officers Dan Dwyer and Gord Gudbranson, Marine Technician
Conrad Cooper, Captain Ken Brown and the crew of the RV John Strickland; Sharon Blackmore (UVIC) for
preparation of SEM samples; and the SEM laboratory (Brent Gowen) at the University of Victoria. This
research was funded by a grant from the BC MOE and NSERC to JTC.

123
838 Aquat Geochem (2011) 17:821–839

References

Allam AI, Pitts G, Hollis JP (1972) Sulfide determination in submerged soils with an ion selective electrode.
Soil Sci 114(6):456–467
Auditor General (2004) Report 5, Salmon Forever: An assessment of the provincial role in sustaining wild
salmon, Report of the auditor general of British Columbia, October 2004. ISBN 0—7726—5216—0.
p 18
Bagarinao T (1992) Sulfide as an environmental factor and toxicant: tolerance and adaptations in aquatic
organisms. Aquatic Toxicology 24:21–62
BCMOE (2002a) Environmental management act: Finfish aquaculture waste control regulation B.C Reg.
256/2002 O.C.836/2002. Queen’s Printer, Victoria
BCMOE (2002b) Protocols for Marine Environmental Monitoring B.C Reg. 256/2002 O.C.836/2002
Schedule B. Queen’s Printer, Victoria
BCMOE (2003) Environmental management act, SBC 2003, Chapter 53. Queen’s Printer, Victoria, British
Columbia, Canada
Berner RA (1963) Electrode studies of hydrogen sulfide in marine sediments. Geochim Cosmo Act
27:563–575
Billon G, Ouddane B, Laureyns J, Boughriet A (2001) Chemistry of metal sulphide in anoxic sediments.
Phys Chem Chem Phys 3:3586–3592
Brooks KM (2001a) An evaluation of the relationship between salmon farm biomass, organic inputs to
sediments, physiochemical changes associated with those inputs and the infaunal response—with
emphasis on total sediment sulphides, total volatile solids, and oxidation-reduction potential as sur-
rogate endpoints for biological monitoring, Final Report. p 172. Aquatic Environmental Sciences, 644
Old Eaglemont Road, Port Townsend, Washington, USA
Brooks KM (2001b) Recommendations to the British Columbia farmed salmon waste management technical
advisory group for biological and physicochemical performance standards applicable to marine Net-
pens. For: the technical advisory group, BC MoE. pp 24
Brooks KM, Mahnken CVW (2003) Interactions of Atlantic salmon in the Pacific northwest environment II.
Organic wastes. Fish Res 62:255–293
Brooks KM, Stierns AR, Backman C (2004) Seven year remediation study at the Carrie Bay Atlantic salmon
(Salmo salar) farm in the Broughton Archipelago, British Columbia, Canada. Aquaculture 239:81–123
Cassella RJ, De Oliveira LG, Santelli RE (1999) On line dissolution of ZnS for sulfide determination in
stabilized water samples with zinc acetate, using spectrophotometry by methylene blue Spec. Letters
32(2):469–484
Clesceri LS, Greenberg AE, Eaton AD, Clesceri LS, Greenberg AE, Eaton AD (eds) (1998) Standard
methods for the examination of water and wastewater. United book press (Maryland) pp 4–167
Cline JD (1969) Spectrophotometric determination of Hydrogen sulphide in natural waters. Limnol Oceaogr
14(3):454–458
Fonselius SH (1983) Determination of hydrogen sulphide. In: Grasshoff K, Ehrhardt M, Kremling K (eds)
Methods of seawater analysis, 2nd edn. Verlag Chemie, Berlin, pp 73–80
Glaister MG, Moody GJ, Thomas JDR (1985) Studies on flow injection analysis with sulpihde ion-selective
electrodes. Analyst 110:113–119
Hargrave BT, Phillipps GA, Doucette LI, White MJ, Milligan TG, Wildish DJ, Cranston RE (1997)
Assessing benthic impacts of organic enrichment from marine aquaculture. Water Air and Soil Poll
99:641–650
Hargrave BT, Holmer M, Newcombe CP (2008) Towards a classification of organic enrichment in marine
sediments based on biogeochemical indicators. Mar Poll Bull 56:810–824
Holland HD (1978) The chemistry of the atmosphere and oceans. Wiley, New York
Hseu TM, Rechnitz GA (1968) Analytical study of a sulfide ion-selective membrane electrode in alkaline
solution. Anal Chem 40(7):1054–1060
Johnson KM, Grimm KA (2001) Opal and organic carbon in laminated diatomaceous sediments: Saanich
Inlet, santa barbara basin and the miocene monterey formation. Marine Geo 174:159–175
Latimer WM (1952) Oxidation potentials (2nd edn). Prentice-Hall, Englewood Cliffs, pp 1–392
Luther GW III, Glazer BT, Hohmann L, Popp JI, Taillefert M, Rozan TF, Brendel PJ, Theberge SM, Nuzzio
DB (2001a) Sulfur speciation monitored in situ with solid state gold amalgam voltammetric micro-
electrodes: polysulfides as a special case in sediments, microbial mats and hydrothermal vent waters.
J Environ Monitoring 3(1):61–66
Luther GW III, Rozan TF, Taillefert M, Nuzzio DB, Di Meo C, Shank TM, Lutz RA, Cary SC (2001b)
Chemical speciation drives hydrothermal vent ecology. Nature 410:813–816

123
Aquat Geochem (2011) 17:821–839 839

Meyer B, Ward K, Koshlap K, Peter L (1983) Second dissociation constant of hydrogen sulfide. Inorg Chem
22(16):2345–2346
Migdisov AA, Williams-Jones AE, Lakshtanov LZ, Alekhin YV (2002) Estimates of the second dissociation
constant of H2S from the surface sulfidation of crystalline sulfur. Geochem Cosmochim Acta
66(10):1713–1725
Morel FMM, Hering JG (1993) Principles and applications of aquatic chemistry. Wiley, New York.
pp 264–272, 398–404
Mosher DC, Moran K (2001) Post-glacial evolution of Saanich Inlet, British Columbia: results of physical
property and seismic reflection stratigraphic analysis. Marine Geo 174:59–77
Nash CE (2001) The net-pen salmon farming Industry in the Pacific Northwest. US Dept Commer, NOAA
Tech. Memo. NMFS-NWFSC-49 pp 125
Nicholls P (1975) The effect of sulfide on cytochrome aa3 isoteric and allosteric shifts of reduced alpha-
peak. Biochim Hiophyr Am 396:24–35
Parsons TR, Maita Y, Lalli CM (1984) A manual of biological and chemical methods for seawater analysis.
Publ Pergamon Press, Oxford, p 173
Rickard D, Luther GW (2007) Chemistry of iron sulfides. Chem Rev 107(2):514–562
Rickard D, Morse JW (2005) Acid volatile sulphide (AVS). Mar Chem 97(3–4):141–197
Sarradin PM, LeBris N, Birot D, Caprais JC (1999) Laboratory adaptation of the methylene blue method to
flow injection analysis: towards in situ sulfide analysis in hydrothermal seawater. Anal Commun
36:157–160
Statistics Canada (2007) Statistics Canada, Agriculture Division, Livestock Section: Aquaculture Statistics -
Catalogue no. 23–222-9. Minister of Industry, Ottawa
Sutherland TF, Petersen SA, Levings CD, Martin AJ (2007) Distinguishing between natural and aquacul-
ture-derived sediment concentrations of heavy metals in the Broughton Archipelago, British Columbia.
Mar Poll Bull 54:1451–1460
Thamdrup B, Finster K, Fossing H, Wurgler-Hansen J, Barker-Jorgensen B (1994) Thiosulfate and sulfite
distributions in porewater of marine sediments related to manganese, iron, and sulfur geochemistry.
Geochim Cosmo Acta 58:67–73
Wang F, Chapman PM (1999) Biological implications of sulphide in sediment—a review focusing on
sediment toxicity. Env Toxi Chem 18(11):2526–2532
Whitfield M (1969) Eh as an operational parameter in estuarine studies. Limnol & Oceanogr 14(4):547–558
Whitfield M (1974) The ion-association model and the buffer capacity of the carbon dioxide system in
seawater at 25 C and 1 atmosphere total pressure. Limnol & Oceanogr 19(2):235–248
Wildish DJ, Akagi HM, Hamilton N, Hargrave BT (1999) A recommended method for monitoring sedi-
ments to detect organic enrichment from mariculture in the Bay of Fundy. Canadian Technical Report
of Fisheries and Aquatic Sciences 2286, St. Andrews, NB

123
View publication stats

También podría gustarte