Sunteți pe pagina 1din 23

Available online at www.sciencedirect.

com

Marine Chemistry 107 (2007) 388 – 410


www.elsevier.com/locate/marchem

Analysis of the water column oxic/anoxic interface in the Black and


Baltic seas with a numerical model
E.V. Yakushev a,b,⁎, F. Pollehne b , G. Jost b , I. Kuznetsov a,b , B. Schneider b , L. Umlauf b
a
Shirshov Institute of Oceanology RAS, Southern Branch, Gelendzhik-7, Krasnodarskii Kray, 353467, Russia
b
Baltic Sea Research Institute Warnemuende, Seestrasse 15, 18119 Rostock, Germany
Received 29 August 2006; received in revised form 31 May 2007; accepted 4 June 2007
Available online 16 June 2007

Abstract

A 1D hydrophysical-biogeochemical model was developed to study the cycling of the main elements in the pelagic redox layer
in seas with anoxic conditions. The formation and decay of organic matter; the reduction and oxidation of nitrogen, sulfur,
manganese, and iron species; and the transformation of phosphorus species were parameterized. Temporal and spatial
developments of the model's variables were described by a system of horizontally integrated vertical diffusion equations for non-
conservative substances. The calculated spatial and temporal distributions of the above-mentioned parameters were in good
agreement with observed vertical distribution patterns of these processes.
To study the influence of seasonal variability on the chemical structure of the pelagic redox layer, specific and distinct
hydrophysical scenarios for the Black Sea and for the Baltic Sea were used. The results clearly showed that organic matter, formed
during the bloom periods by phytoplankton, exerts a major and direct influence on the structure of the remote redox interface in
both seas as well as on the processes occurring in them. This is due to competition for dissolved oxygen between its consumption
for oxidation of organic matter originating in the euphotic layer and the consumption for oxidation of reductants supplied from the
anoxic deep water. As a result, ammonification, nitrification, denitrification and sulfate reduction dominate in the spring and
summer, while the oxidation of reduced forms of metals and of hydrogen sulfide dominates in the winter.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Biogeochemical modelling; Oxic/anoxic interface; Redox processes; Anoxic conditions; Baltic Sea; Black Sea

1. Introduction continue due to bacterial activities employing electron


acceptors other than oxygen. In some cases, this results
Anoxic conditions are a natural feature of water in the reduction of sulfate, a major constituent of
columns in many of the world's seas. They arise from an seawater, and in the production of hydrogen sulfide,
imbalance in the transport rates of organic matter (OM) which is toxic to most higher life forms.
and oxygen into deeper layers, such that oxygen is The energy derived from the oxidation of reduced
depleted and an excess of organic material is left to be inorganic compounds in the anoxic zone fuels microbial
decomposed. Nonetheless, decomposition processes communities that produce OM via chemosynthesis
(Nealson and Stahl, 1997; Fenchel et al., 1998; Sorokin,
⁎ Corresponding author. 2002; Canfield et al., 2005). This process together with
E-mail address: e_yakushev@yahoo.com (E.V. Yakushev). the oxic, anoxic, and suboxic mineralization of OM as
0304-4203/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.marchem.2007.06.003
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 389

well as chemical reactions between reduced and the oxic/anoxic interfaces found in the Black and Baltic
oxidized compounds accounts for the complexity of Seas are shown in Fig. 1. In both of these interfaces, the
the redox layer. nitrate maximum is located at a depth where the vertical
The above-mentioned imbalance between OM trans- gradient of oxygen decreases (lower part of the
port and oxygen occurs when a hydrophysical structure oxycline). The depths of the onset of ammonia and
with a pronounced pycnocline is created. The existence dissolved manganese correspond to the position of the
of such structures can be temporary or permanent, and most pronounced phosphate minimum in the Black Sea.
they create corresponding zones of temporary or This depth is identical to the one at which oxygen
permanent anoxia. Permanent anoxic conditions are depletion occurs, whereas the hydrogen sulfide onset is
observed in numerous lakes, fjords, and some regions of 5–10 m deeper. The vertical distribution of transmission
the World Ocean (Black Sea, Baltic Sea Deeps, Cariaco (Xmiss, Fig. 1) is characterized by the presence of a
Basin). The processes that affect the formation of each turbidity layer in the vicinity of sulfide onset.
system's hydrophysical structure vary in scale, ranging In the present study, a model designed to analyze the
from molecular diffusion to climatic variability. various systems that are active in the water column is
The redox interfaces of marine basins have many described. This 1D, hydrophysical-biogeochemical, O–
features in common. The hydrochemical structures of N–S–P–Mn–Fe model simulates the main distribution

Fig. 1. Vertical distributions of hydrochemical parameters in the Black (upper panel) and Baltic (lower panel) Seas. “Xmiss” corresponds to Dr
Haardt, turbidity (A) and to Chelsea transmissometer (B) data. The arrows shows the depths of: (1) NO3 maximum and the lower portion of the
oxycline; (2) O2 depletion, the onsets of Mn(II) and NH4, and the PO4 minimum; and (3) the onset of H2S and the maximum PO4 gradient.
390 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

patterns of the biogeochemical variables and processes (S2O3), sulfate (SO4), ammonia (NH4), nitrite (NO2),
at the redox interface. Accordingly, the water column nitrate (NO3), particulate organic nitrogen (PON),
between the surface and a depth of about 100 m below dissolved organic nitrogen (DON), phosphate (PO4),
the oxic/anoxic interface was considered. In contrast to particulate organic phosphorus (POP), dissolved organic
previous versions of this model (Yakushev, 1992, 1999; phosphorus (DOP), bivalent manganese (Mn(II)), triva-
Yakushev and Neretin, 1997) and to other models lent manganese (Mn(III)), quadrivalent manganese (Mn
devoted to analyzing oxic/anoxic interface processes in (IV)), bivalent iron (Fe(II)), trivalent iron (Fe(III)),
the water column (Oguz et al., 1998; Konovalov et al., phytoplankton (Phy), zooplankton (Zoo), aerobic het-
2006), we parameterized OM formation during photo- erotrophic bacteria (Bhe), aerobic autotrophic bacteria
synthesis and chemosynthesis. Hence, our current (Bae), anaerobic heterotrophic bacteria (Bha), and
model includes feedbacks between the upward fluxes anaerobic autotrophic bacteria (Baa).
of nutrients and OM production. The time-space evolution of the model's variables is
This model was used to analyze key processes driving described by a system of horizontally integrated vertical-
the formation of the redox-layer structure, i.e., sulfide diffusion equations for non-conservative substances:
oxidation, oxygen consumption, and phosphate distri-
bution. The structure and seasonality of a redox interface ACi A ACi AððWC þ WMe ÞCi Þ
¼ Kz  þ RCi ð1Þ
were simulated based on two hydrophysical scenarios— At Az Az Az
a simplified one for the Black Sea and a more
complicated one for the Baltic Sea that was calculated where Ci is the concentration of a model variable, Kz the
using the General Ocean Turbulent Model (GOTM; vertical turbulent diffusion coefficient, WC the sinking
Burchard et al., 1999). rate of particulate matter; WMe the acceleration of the
sinkingPrate due to the settling of Mn hydroxides, and
2. The model RCi ¼ j RateRjCi the rate of biogeochemical produc-
tion/consumption of Ci, which is an algebraic sum of
The following biogeochemical parameters (Ci) were local fluxes (sources and sinks) caused by biogeochem-
considered (Fig. 2): dissolved oxygen, (O2), hydrogen ical interaction (RateBjCi). A table with RCi is presented
sulfide (H2S), total elemental sulfur (S0), thiosulfate in the Appendix (Table A1).

Fig. 2. Flow-chart of biogeochemical processes represented in the model.


E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 391

A detailed description of this model is given in the effect of increased sinking rates due to the formation of
Yakushev et al. (2006a). Mn(IV) and Fe(III) oxides and their association with
particulate organic matter (POM) was parameterized.
2.1. Hydrophysical scenarios Yakushev and Debolskaya (2000) found that the
precipitation of particulate Mn oxide significantly
For the Black Sea Kz in the redox layer was increases the flux of sinking particles, which, in turn,
determined based on the vertical density structure and affects the overall distribution of particles. This effect was
calculated according to Gargett (1984): parameterized in Eq. (1) as an additional term reflecting
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi this effect:
g Aq
Kz ¼ a0 N q ; where N ¼  : ð2Þ
q Az WMe ¼ WMe
max MnðIVÞ
ð4Þ
MnðIVÞ þ KMe
where N is the buoyancy frequency, g is the acceleration of
gravity, ρ is the mean density, and a0 and q are empirical max
Coefficients W Me and KMe are given in Table A2.
coefficients. Estimates for the Black 0:5
Sea redox layer
obtained with Kz ¼ 1:62d103 d qg dq yielded values
dz 2.3. Boundary conditions
of about 1 · 10− 5 m2 s− 1 (Samodurov and Ivanov, 1998).
The estimates of Kz obtained by Stokozov (2004),
A 1D water column that ranged between the sea
who analyzed the spreading of 137Cs after the Chernobyl
surface (upper boundary) and a water depth of 200 m
accident, were 1–3 · 10− 5 in the Black Sea Central Basin
(lower boundary) was considered.
and 3–11 · 10− 5 m2/s in the region of the rim current.
At the upper boundary except for O2, PO4, and inorganic
These larger estimates are more realistic because they
nitrogen compounds, the surface fluxes of the chemical
were calibrated using the flux of a passive tracer and
constituents considered in the model were assumed to be
may therefore reflect processes not considered in the
zero. O2 exchange is given by the flux equation:
Gargett (1984) approach. In order to match this higher
Kz, the coefficients were revised accordingly and used to QO2 ¼ k660 ðSc=660Þ0:5 ðOxsat  O2 Þ ð5Þ
re-calculate Kz:
  where Oxsat is the oxygen saturation concentration as
2 g dq 0:5
Kz ¼ 1:94d10 d ð3Þ a function of temperature and salinity, according to
q dz UNESCO (1986); Sc is the Schmidt number; and k660 is
the reference (Sc= 660, CO2 at 20 °C) gas-exchange
Seasonal variabilities in light and hydrophysical
transfer velocity. To describe k660 as a function of wind
structure were considered as external parameters. To
speed, the following equation (Schneider et al., 2002) was
describe the variability between the surface and the cold
used:
intermediate layer, the changes between two typical
distributions observed in the central Black Sea in winter k660 ¼ 0:365 u2 þ 0:46 u ð6Þ
and summer were considered (Fig. 3). It was assumed
that the changes between typical summer and typical Simulations were carried out based on a mean wind
winter structures occur according to a sinusoidal speed of 5 m s− 1 for both the Black Sea and the Baltic Sea.
function. Daily calculated density values were used for The input of phosphorus (QP) and total inorganic
estimations of Kz according to Eq. (3). nitrogen (QN) from rivers and atmospheric precipitates was
For the Baltic Sea, temperature, Kz, and irradiance taken into account. QP and QN for the Black Sea were
were determined based on the results of the GOTM calculated with data from Fonselius (1974) and amounted
(Burchard et al., 1999) for 1992, when the vertical to QP =0.13 mmol m− 2 d− 1 and QN =1.5 mmol m− 2 d− 1.
distribution in the Gotland Basin was stable and anoxia For the Baltic Sea, QP = 0.0085 mmol m− 2 d− 1 and Q-
had developed. The calculated arrays reflecting the daily −2 −1
NO3 = 0.46 mmol m d were determined on the basis
changes in these variables were directly applied in the of estimates for total-N (990000 t/year) and total-P
model's equations. (40000 t/year) input (HELCOM, 2002) and an area of
415,266 km2.
2.2. Particle sinking At the lower boundary since this model does not
consider the nutrient dynamics below 200 m, the
Constant WCi values were assumed for phytoplankton, concentrations of the main reductants at the lower
zooplankton, bacteria, and detritus (Table A1). However, boundary were assumed to be constant. According to
392 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

Fig. 3. Typical temperature (dotted) and salinity (solid) distributions for the central Black Sea in winter (A) and summer (B). Data from RV“Knorr”
and RV “Akvanavt” cruises.

observations, the values for the Black Sea are: these decompositions are needed for modeling long-term
NH4 = 20 μM, H2S = 60 μM, Mn(II) = 8 μM, Fe(II) = processes in sediments, and they can be described with
0.4 μM, PO4 = 4.5 μM. Those for the Baltic Sea were: ‘multi-G’ models, in which OM is divided into several
NH4 = 10 μM, H2S = 40 μM, Mn(II) = 10 μM, Fe(II) = classes according to its degradability (Westrich and
0.4 μM, PO4 = 4.5 μM. For the other parameters, an Berner, 1984; Boudreau, 1996).
absence of flux was assumed. In this model, a simplified approach was used in that
OM was divided into dissolved organic matter (DOM)
2.4. Parameterization of biogeochemical processes and POM, with different rates of mineralization. POM
was considered as a detrital labile OM that can be
The chemical and biological pathways (shown in mineralized directly. This approach was widely used in
Fig. 2) were formally described by using our previously previous models (e.g., Fasham et al., 1990; Gregoire et al.,
reported parameterizations (Yakushev, 1992, 1999; 1997; Fennel and Neumann, 2004). Here, the stoichiom-
Yakushev and Mikhailovsky, 1995; Yakushev and etry of mineralization reactions proposed by Richards
Neretin, 1997) as well as those of others (Fasham et al., (1965) was used in modeling oxic and anoxic conditions.
1990; Fennel and Neumann, 2004; Ayzatulllin and The main goal of this model was to explain component
Leonov, 1975; Savchuk and Wulff, 1996; Boudreau, processes in the redox layer. This was achieved with a
1996; Oguz et al., 1998; Gregoire et al., 1997; Konovalov model of living organisms that is much simpler than other
et al., 2006). The values for the coefficients in the rate models for the Black (Gregoire et al., 1997; Oguz et al.,
equations were obtained either from the literature or from 1998) and Baltic (Savchuk and Wulff, 1996; Fennel and
fitting the model to measured concentration profiles. A Neumann, 2004) Seas. Since the main role of the Phy and
detailed description of the model is given in (Yakushev Zoo parameters was to describe the seasonality of OM
et al., 2006b). Here, the biogeochemical processes production, they were not further subdivided. It was also
considered, their chemical equations, and formulas and assumed that the uptake rate of inorganic nutrients by
coefficients for the parameterizations are given in the phytoplankton equals the growth rate of the phytoplank-
Appendix list (Tables A1, A2). ton. The role of bacteria at the redox interface is very
The mineralization of OM is a key process in the important; hence, based on the results of a series of
modeling of oxygen-deficient and anoxic conditions, experiments, bacteria were divided into four groups
because the electron acceptors for this reaction change according to their relation to a particular energy source
from oxygen to nitrate, metal oxides, and finally sulfate. and to OM transformation (Tables A1 and A2).
In addition, the rates of mineralization vary between the The formulation for the biogeochemical production/
different electron acceptors. Microbial degradation of the consumption rates, RCi, as an algebraic sum of reactions
various types of OM, with different stabilities, covers time affecting the concentrations of certain compartments, is
scales from hours to millions of years. Detailed kinetics of presented in Table A3.
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 393

2.5. Computational aspects described hydrophysical scenario was simulated by the


model (Fig. 4). The similarities in the basic chemical
The initial calculations employed a uniform distribu- structures of the redox interfaces of the Black Sea and
tion of the considered variables. Numerical integration was the Baltic Sea allowed a comparison of the results.
conducted with the Eurelian scheme and by process
splitting. The time steps were 0.00125 d− 1 for diffusion 3.1. Distributions of biogeochemical parameters
and 0.0025 d− 1 for biogeochemical processes and
sedimentation. The vertical resolution was 2 m. A quasi- In the model, the dissolved oxygen structure was
stationary solution with seasonal forced oscillations was characterized by a uniform distribution in the upper
reached. There were no changes in the year-averaged mixed layer (270 μM), the presence of a pronounced
concentrations of the variables for at least 100 model years. oxycline with a vertical gradient of 7 μM m− 1, and a
decreased vertical gradient in the deeper layers.
3. Results and discussion The vertical profiles of nitrogen compounds (NH4,
NO2, NO3), as calculated by the model, reflected the main
Redox-layer development in response to seasonal features of the distributions of these compounds as ob-
changes and, for the Black Sea, according to the above- served in nature. Nitrate concentrations in the maximum

Fig. 4. The vertical distribution of the parameters in summer (Black Sea) as calculated by the model.
394 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

layer reached 4.5 μM in the model, as is the case in nature. concentrations of Fe(III) (0.01 μM) corresponded to
The NO2 maximum peak was situated at approximately previous observations (Pakhomova, 2005) and were
the same depth, where the concentrations of NH4 and NO3 located deeper than the Mn(IV) maximum. Therefore,
were equal. the modeled “iron interface” was slightly deeper than the
According to the model, H2S was present about 5– “manganese interface”, as is observed in nature (Lewis
10 m below the onset of the increase in NH4, while and Landing, 1991).
maximum absolute values (1.5 μM) for S0 occurred at The biological characteristics, defined in terms of the
the depth at which H2S appeared. The S2O3 concentra- biomasses of phytoplankton and zooplankton, ranged
tions were uniform, with a small maximum of 0.2 μM from 50 to 250 mg m− 3 depending on the season, in
at a depth slightly above the H2S onset. The Mn(II) agreement with the observations for the Black Sea
onset occurred at the same depth as the NH4 onset. (Sorokin, 2002). The calculated bacterial biomass in the
The maxima of Mn(III) and Mn(IV) also formed redox zone (5–10 mg m− 3) as a whole also corre-
there. The vertical distribution of Fe was similar to sponded to the observed values (Pimenov and Neretin,
that of Mn. 2006). The model parameters PON and DON (0.3–
The P distribution was characterized by a slight 2.0 μM and 2–5 μM, respectively) were lower than the
minimum at the depth of the Mn(II) onset and an reported observed values from the Black Sea (Konova-
intensive concentration increase that extended down to lov et al., 2006; Yakushev et al., 2006b) and the Baltic
the depth of the H2S appearance. Sea (2.9 μM and 14.6 μM, Nagel, personal communi-
In general, the vertical structure calculated by the cation, 2006). The difference can be explained by the
model reflected the basic structural features that fact that our model computes only autochthonous OM
characterize the redox interfaces of the natural water and does not consider allochtonous material derived
column, in particular: (i) the correspondence between from rivers (and during North Sea water inflow events in
the depth of the nitrate maximum and the changing O2 the Baltic Sea).
concentrations, (ii) the similarity of the depth of onsets
of Mn(II) and NH4, and (iii) the positions of both the 3.2. Processes
H2S onset and a layer with maximum gradients of
phosphate, located several meters below. Therefore, the The similarities of the basic structural features of the
model confirmed that the sequences of the disappear- redox interface in the Black and Baltic Seas were related
ance of electron acceptors from the vertical profile to the comparable rates of the respective biogeochemical
(O2 → NO3 → Mn(IV) → Fe(III)) and of the appearance processes (Table 1).
of electron donors (Mn(II) → NH4 → Fe(II) → H2S) The model results for primary production and dark
correspond to the sequences of the theoretical “electron CO2 fixation were close to the observed values. The
tower” (Nealson and Stahl, 1997; Fenchel et al., 1998; model calculated a maximum rate of chemosynthesis of
Canfield et al., 2005). 0.3–0.7 μM C l− 1, which corresponded well to the
Estimations of the maxima of the various parameters results of actual measurements. As determined from the
and of the vertical gradients were comparable with observed values, the rate of chemosynthesis is 0.8–
observed values. The vertical gradient of H2S in the 1.0 μM C l− 1 in the Baltic Sea and 0.2–2 μM C l− 1 in
model ranged from 0.66 to 0.5 μM m− 1, with a the Black Sea (Table 1).
maximum several meters below the onset (compared to The model's ratio of the annually integrated rate of
0.63 μM m− 1 in the Black Sea, according to Volkov photosynthesis and the rate of chemosynthesis was about
et al., 1992). The modeled vertical gradient, in which 4–5. This corresponded to estimates that the rate of
dissolved Mn(II) was located more shallowly than the chemoautotrophic production in the central Black Sea is
H2S onset, was about 0.6 μM m− 1, whereas the values 10–32% of that of surface photoautotrophic production
obtained from the field measurements (0.3–0.5 μM m− 1) (Yilmaz et al., 2006). A similar ratio was reported for the
were smaller (Pakhomova, 2005; Yakushev et al., Baltic Sea (Detmer et al., 1993).
2006b). In the model, Mn(IV) concentrations reached The processes of sulfide oxidation and sulfate
0.1–0.2 μM in the zone between the onset of oxygen and reduction described by the model corresponded well to
the onset of hydrogen sulfide. These values are in observations from the Black Sea (Table 1).
agreement with data for the Black (Erdogan et al., 2003) The modeled rate of ammonification was similar to
and the Baltic (Neretin et al., 2003) Seas. The model- the value obtained for the Black Sea by Sorokin (2002),
calculated Mn(III) concentrations correspond to those but much higher than the estimates by Kuypers et al.
measured by Trouwborst et al. (2006). The maximum (2003). The model's nitrification rate was close to the
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 395

Table 1
Comparison of the rates of biogeochemical processes in model, Black Sea and Baltic Sea
Process (units) Model Black Sea Baltic Sea
Primary production 90 40–240 (Sorokin, 2002; Yilmaz et al., 150 (Wasmund et al., 2005)
(g C m− 2 yr− 1) 2006)
Dark CO2 fixation (μM C d− 1) 0.3–0.7 0.2 (Pimenov and Neretin, 2006), 0.8–1.0 (Jost, p.c. 2006)
0.4–2 (Yilmaz et al., 2006),
0.2–0.6 (Morgan et al., 2006)
Sulfide oxidation (μM S d− 1) 1.9 0.3–4.5 (Jorgensen et al., 1991; Sorokin, 2002)
S0 oxidation (μM S d− 1) 0.3 0.6–0.9 (Sorokin, 2002),
S2O3 oxidation (μM S d− 1) 0.2 0.2–0.6 (Sorokin, 2002)
Sulfate reduction (μM S d− 1) 0.02 0.003–0.036 (Jorgensen et al., 1991);
0.2–0.6 (Pimenov and Neretin, 2006)
Ammonification in oxic zone 0.1–0.5 0.1–0.5 (calculated on the data of
(μM N d− 1) Sorokin et al., 1991), 0.005–0.05
(Kuypers et al., 2003)
Nitrification, (μM N d− 1) 0.2 0.75 NH4 oxidation — 0.005–0.05 (Ward and 0.001–0.28 (Enoksson, 1986)
(deep NO2 max) Kilpatrick, 1991), 0.02–0.05 (Sorokin, 2002) 0.017–0.48 (Bauer, 2003)
NO2 oxidation — 0.05–0.24
(Ward and Kilpatrick, 1991)
Denitrification (μM N d− 1) 0.2 0.002 (Ward and Kilpatrick, 1991). 0 μM N2 d− 1 (Hannig et al., 2006)
0.044–0.11
(Brettar and Rheinheimer, 1991)
Nitrogen fixation (μM N d− 1) 0.1–0.5 0.02–0.04 (Sorokin, 2002) 0.08–2.3 mmol N m− 2 d− 1
(Wasmund et al., 2005)
0.68–0.74 mmol N m− 2 d− 1
(Stal and Walsby, 2000)
Thiodenitrification (μM N d− 1) 0.2 Is possible (Sorokin, 2002) 0–2.7 (Hannig et al., 2006)
Anammox (μM N d− 1) 0–0.03 0.007 (Kuypers et al., 2003) 0–0.05 (Hannig et al., 2006)
Mn oxidation μM Mn d− 1 1.0 0.18–1.9 (Tebo, 1991)
Mn reduction 0.9 0.96–3.6 (Nealson et al., 1991)

rates of the Baltic and Black Seas. Denitrification was analysis of results from other studies. In this section,
larger than observed, while thiodenitrification was at the examples of applications to specific questions con-
limit of possible values for the Baltic Sea. Modeled cerning the function of the redox layer are provided.
values of anammox corresponded well to the observed
results, whereas the rate of nitrogen fixation was twice 3.3.1. Oxidation of H2S
as large as that measured in the Black Sea. The rates of In the 1990s, it became evident that the potential sink
manganese oxidation (1.0 μM Mn d− 1) and manganese of H2S is not balanced solely by the O2 supply (Murray
reduction (0.9 μM Mn d−1) were in accordance with et al., 1995). This discrepancy continues to challenge
observations (Table 1). current thinking regarding the biogeochemistry of
Generally, the model's estimates agreed well with the marine redox interfaces. It has been generally assumed
observed values for the same processes, but there were that the oxidation of H2S is connected with the activity
also a few differences. These may have been due to the fact of chemolithotrophic bacteria (Zopfi et al., 2001).
that the model represents an averaged balanced picture, While experiments with labeled S have demonstrated
while observational data frequently reflect a transient that the maximum rate of H2S oxidation is 0–20 m
situation, due to intrusions or temporal variability. below the sulfide onset, the electron acceptor of this
reaction below the sulfide interface has not been
3.3. Model experiments identified experimentally. Maximum dark CO2 fixation
is also observed in the 10 to 20-m layer below the
Models are developed in order to analyze data sulfide interface (Pimenov and Neretin, 2006).
obtained through observations. It was shown in the The model was therefore used to analyze the roles of
previous section that the calculated spatial and temporal different electron acceptors. H2S oxidation was repre-
distributions of the parameters of interest agreed sented by the following reactions:
reasonably well with those data. The extent of this
agreement supports application of the model to the 2H2 S þ O2 →2S0 þ 2H2 O ð7Þ
396 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

3H2 S þ 4NO−3 þ 6OH− →3SO2−


4 þ 2N2 þ 6H2 O ð8Þ conditions or following intrusions. The role of anoxygenic
photosynthesis should be subject to change depending on
H2 S þ MnO2 →So þ Mn2þ þ 2OH− ð9Þ the depth of the redox interface. In any case, a dominant
role in the oxidation of sulfide seems to be played by
H2 S þ 2FeOOH→2Fe2þ þ S0 þ 2OH− ð10Þ oxidized Mn species, which are formed several meters
higher than the sulfide onset via the reaction of dissolved
The modeled vertical profiles of these processes rates Mn(II) with O2. The precipitation of Mn(IV) leads to an
are presented in Fig. 5A. increase in the particles density and thus an accelerated
Estimates obtained with the model showed that H2S rate of sinking (Yakushev and Debolskaya, 2000).
oxidation with O2 is very slow (less than 0.01 μM d− 1, or Consequently, the depths of sulfide and Mn(II) onset
0.02 μM S d− 1). The maximum rates of Mn reduction differ. Our model demonstrated that an acceleration of the
were 0.8 μM d− 1, Fe reduction 0.10 μM d− 1, and NO3 sinking rate significantly affects one of the key features of
reduction about 0.30 μM d− 1. Therefore, H2S oxidation the oxic/anoxic interface, i.e., the 5 to 10-m difference
was due primarily to the reduction of Mn(IV) — 69%, between the depths of the Mn(II) and sulfide onsets
NO3 — 24%, Fe(III) — 6%, and O2 — 1% (Fig. 5A). The (Fig. 6A). In the absence of an accelerated sinking of Mn
oxidation of elemental sulfur and thiosulfate to sulfate hydroxide (WMn = 0), the oxic/anoxic interface became
with dissolved oxygen occurred at depths that were shallower, such that the H2S onset was located at the same
several meters shallower (Fig. 5A, central column). depth as the Mn(II) onset, and a layer appeared in which
In addition to the above-mentioned reactions, H2S both O2 and H2S were present (Fig 6B).
can be oxidized through anoxygenic photosynthesis The intense vertical transport of detrital particles and
(Canfield et al., 2005; Overmann and Manske, 2005): heavy Mn components explains why there is a zone
where oxygen and sulfide are absent. This, in turn,
2H2 S þ CO2 →CH2 O þ 2S0 þ H2 O ð11Þ facilitates processes such as anammox (Dalsgaard et al.,
2003; Kuypers et al., 2003) and the oxidation of Fe with
Simulations that parameterized all dark CO2 fixation Mn, and allows the presence of Mn in the form of Mn
occurring in connection with the above-mentioned (III), which otherwise is quickly oxidized by O2 or
processes were carried out using the model. The results reduced by H2S (Webb et al., 2005).
are presented in Fig. 5B. In this case, CO2 consumption
would have resulted in an oxidation rate of 0.1–0.5 μM 3.3.2. Consumption of O2 in the suboxic layer
S d− 1. This was a significant share of the total sulfide We used our model to analyze the roles of different
oxidation (32%) but, the highest share of sulfide was processes in O2 consumption. The vertical distributions of
still oxidized with Mn(IV). these processes from the surface to 200 m are presented in
Recently, the production of dissolved oxidized Mn, Fig. 7A. Both OM mineralization (49%) and the
in the form of Mn(III), by Mn(II)-oxidizing bacteria and respiration of living organisms (29%) were found to
by incubations with Black Sea suboxic zone water was play dominant roles in O2 consumption. These were the
described (Webb et al., 2005). Dissolved Mn(III) has main processes that compensated for the production of O2
also been directly detected in the suboxic zone by photosynthesis. Nitrification consumed 11% of the O2
(Trouwborst et al., 2006). Mn(III), in dissolved or available in the lower part of the oxic zone, while the
solid form, is an important intermediate product of the oxidation of reduced compounds in the water column of a
Mn cycle (Kostka et al., 1995). Therefore, the model sea with anoxic conditions was estimated at less than 10%.
was employed to take into account Mn reduction Under oxygen-deficient conditions, i.e., from O2
through the intermediate Mn(III): concentrations b 40 μM to the sulfidic boundary
(Fig. 7B), less O2 was consumed in connection with the
2Mn4þ þ HS− þ OH− →2Mn3þ þ S0 þ H2 O ð12Þ mineralization of OM and respiration (about 30%), because
in this layer OM decomposition was mainly associated with
2Mn3þ þ HS− →2Mn2þ þ S0 þ Hþ ð13Þ denitrification. The largest amount of O2 was consumed by
nitrification (mainly ammonia oxidation to nitrate −13%,
In this case, sulfide oxidation was due to a decrease in and nitrite oxidation to nitrate −8%). Consumption by the
Mn(IV) — 50%, Mn(III) — 43%, NO3 — 5%, Fe(III) — oxidation of reductants was about 50%, with a dominant
1%, and O2 — 1% (Fig. 5C). share for Mn(III) (20.1%) and Mn(II) (17.9%) and less for
These numerical experiments likely reflected specific reduced sulfur species. The amount of O2 consumed by the
situations that occur in natural water, under stable oxidation of Fe(II) was negligible.
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 397

Fig. 5. Left column: rates of oxidation of hydrogen sulfide with O2, NO3, Mn(IV), Fe(III) (A), O2, NO3, Mn(IV), Fe(III), CO2 (B) and O2, NO3, Mn
(IV), Mn(III), Fe(III) (C). Middle column: corresponding rates of oxidation of S0 and S2O3 with O2. Right column: potential electron acceptors and
their extent of involvement in sulfide oxidation.

These estimates are important for the modification of well-defined shallow minimum about 5–10 m shallower
ecological models (Fennel and Neumann, 2004; Sav- than the sulfide onset, a maximum below the sulfide onset,
chuk and Wulff, 1996) that describe oxygen processes and a second deep minimum about 30–50 m deeper
but do not consider S, Mn, and Fe cycles. (Fig. 1A). This structure is called the “phosphate dipole”
(Shaffer, 1986). A similar phosphate distribution is
3.3.3. “Phosphate dipole” observed in the Baltic Sea (Fig. 1B). The reason for the
The vertical distribution of phosphate in the Black Sea is formation of this dipolar structure is unclear. Previously, it
characterized by increased concentrations in the oxycline, a was thought to be connected with chemosynthesis (Sorokin,
398 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

Fig. 6. Vertical distribution of O2, H2S, Mn(II), Mn(III), Mn(IV), and PO4 due to the accelerated sinking of particles with Mn hydroxides; WMn = 18 m
d− 1 (A) and WMn = 0 m d− 1 (B).

2002) and/or the co-precipitation of phosphate with metal et al., 2005). Mn(III) present in the Black Sea suboxic zone
hydroxides (Shaffer, 1986). However, both theories are (Trouwborst et al., 2006) might easily form complexes with
probably incorrect because the maximum values of pyrophosphate. The phosphorus minimum is located at the
chemosynthesis occur below the sulfide onset, where the same depth where Mn(II) is depleted due to possible
phosphate content is also maximal, and the co-precipitation oxidation with oxygen, and its maximum is located about
of phosphate with iron hydroxides is not large enough to 5 m below the sulfide interface. Pyrophosphate particles
account for its consumption. The Fe:P ratio during co- were observed (T. Leipe, personal communication, 2006) at
precipitation with iron hydroxides has been reported to be 4 the redox interface of the Baltic Sea, and we observed a
(Savenko, 1995) or 2.7 (T. Leipe, personal communication, maximum of polyphosphate in the same layer in the
2006). Laboratory experiments showed very high ratios of Northeastern Black Sea in 2006. These depths coincide
Mn:P=1000 in the co-precipitation of Mn hydroxides with the likely limits of the Mn(III) maximum.
(Savenko and Baturin, 1996). Therefore, phosphorous The influence of PO4 distribution on its consumption
removal by precipitation of Mn hydroxides can be ignored. by (1) chemosynthesis, (2) co-precipitation with Fe(III),
It is possible, however, that Mn(III), an intermediate and (3) the formation of complexes with Mn(III) was
product between Mn(IV) and Mn(II), plays a key role in therefore studied in a series of numerical experiments
precipitation of phosphate. Known Mn(III) ligands that employing our model. The results are presented in Fig. 8.
bind with enough strength to stabilize Mn(III) in solution In the absence of these three factors, there were no
include inorganic ligands, such as pyrophosphate (Webb anomalies in phosphate distribution in the vicinity of the
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 399

Fig. 7. Dissolved oxygen consumption in a water column of 0–200 m (A) and in the layer below 70 m (O2 b 40 μM) (B).

oxic/anoxic interface (Fig. 8A). Chemosynthesis resulted 3.3.4. Seasonal changes in the redox-layer
in negligible changes (Fig. 8B), and co-precipitation The model was also employed to study the effects of
with Fe(III) at Fe:P= 2.7 led to a minor decrease in the seasonal variability on structural differences in the redox
phosphate concentration (Fig. 8C). Inclusion of the layer in the Black Sea and the Baltic Sea with respect to
formation of complexes with Mn(III) yielded a vertical two different scenarios.
distribution very similar to the observed one (Fig. 8D). In the simplified scenario applied to the study of the
According to Davies (1969), the ratios of Mn(III) Black Sea, the seasonality of phytoplankton develop-
pyrophosphate complexes may be Mn:P = 0.25 for Mn ment led to a summer-time increase in the flux of POM
(HP2O7)23− or Mn:P = 0.17 for Mn(H2P2O7)33−. In our into the anoxic zone. The oxygen content in the suboxic
model, Mn:P was set at 0.66, about 2.5–4 times higher zone changed seasonally depending on the amount of
than those values. At this ratio, only about 25% of the Mn OM degradation (Fig. 9).
(III) was predicted to form complexes with polypho- Calculations for the Baltic Sea were based on the
sphate, while the rest would complex with other ligands. results of GOTM model (Burchard et al., 1999). In this
The Mn(III) concentration calculated by the model was more realistic scenario, the details regarding seasonal
0.2 μM (Fig. 4), i.e., smaller than the observed variability differed from those of the Black Sea: the spring
concentration range (0.5–1.5 μM, Trouwborst et al., bloom was more intense and shorter, and in autumn a
2006). This concentration could explain the phosphate second phytoplankton bloom occurred (Fig. 10). As was
dipole, even if only a fraction of the Mn(III) formed the case for numerical experiments involving the Black
complexes with P. Further study of the relationships Sea, the model reproduced the seasonal changes in the
between Mn(III), pyrophosphate, and polyphosphate are depths of the redox layer.
important for a better understanding of the ecologies of the For both scenarios, the calculations clearly showed that
Black Sea and the Baltic Sea, since the upward flux of OM formed in the euphotic layer significantly influences
phosphate limits the amount of photosynthesis during the structure of the redox interface and the processes that
certain periods. take place in it. The redox zone is most likely marked by a
400 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

Fig. 8. Influence of chemosynthesis and co-precipitation with iron or manganese on the formation of the “phosphate dipole”. See text for a more
detailed explanation.

competition for consumption of dissolved oxygen order to maintain the system at steady state with regard
between OM supplied from the upper layers and reduced to dissolved inorganic nitrogen. According to the model,
species of S, Mn, Fe and N supplied from the anoxic zone. N2 fixation should occur during the summer in the low
The increase in OM flux in the summer causes more boundary of the photic layer. (Figs. 9, 10). The role of
intense activity of heterotrophic (Figs. 9, 10) (both in oxic N2 fixation in the nitrogen balance of the Baltic Sea is
and anoxic zones) and aerobic autotrophic (nitrifiers) very large, (Schneider et al., 2002; Wasmund et al.,
bacteria, whereas the activity of anaerobic chemolitho- 2005) testify that this process can contribute up to 60%
trophic organisms is reduced (Figs. 9, 10). In the winter of the annual nitrogen consumption in the central Baltic
months, the oxidation of reduced species of S, Mn, Fe Sea surface water. In the Black Sea there are no direct
and the upward flux of ammonia from the deep water observations on N2 fixation in the last decades, while in
intensifies. As a result, larger amounts of oxidized Mn the 1950s–1960s N2 fixation in the Black Sea was
species are formed, and sulfide oxidation reactions as well intensive (Pshenin, 1963). Its possible that the absence
as the growth of chemolithotrophic bacteria play more of N2 fixation in 1990s is connected with the intensive
important roles. Therefore, according to the model's eutrophication of the Black Sea in 1970s–1980s, and
estimates, processes such as ammonification, denitrifica- favorable conditions for the N2-fixation have now re-
tion, sulfate reduction, and nitrification predominate in the appeared. Yilmaz et al. (2006) reported that the
spring and summer, while the oxidation of reduced forms concentrations of nitrate decreased considerably at the
of metals and of hydrogen sulfide (with all possible beginning of 2000s and nitrogen became a limiting
electron acceptors) predominates in the winter. element. The d15N of suspended PON was around 0‰
The model demonstrated that in seas with anoxic in the surface waters of the NE Black Sea during the late
conditions N2 fixation should occur. The large NO3 June 2005 which could indicate the presence of N fixing
losses due to denitrification must be compensated for in organisms (C. Fuschman, persomal communication).
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 401

Fig. 9. Seasonal changes in vertical structure (Black Sea).


402
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410
Fig. 10. Seasonal changes in vertical structure (Baltic Sea).
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 403

4. Summary and conclusions observed structures are formed by biogeochemical


transformations as well as the processes included in the
On the basis of this model it was possible to dem- model, specifically, biogeochemical interactions, vertical
onstrate the relationships between biogeochemical turbulence, and sinking.
transformation processes and the distribution of bio- Our numerical experiments showed that the season-
geochemical parameters. In addition, three layers within ality of OM production results in: (1) competition for
the boundaries of the redox zone were distinguished: dissolved oxygen by consumption for OM mineraliza-
tion and for oxidation of reduced compounds (H2S, NH4,
4.1. Depth of the NO3 maximum Mn(II)), and (2) the seasonal activity of the functional
groups of bacteria involved in these processes. Hetero-
In the upper part of the redox zone, the concentra- trophic processes were shown to likely be more intense
tions of O2 decreased to 15–20 μM; the O2 vertical during the summer, and the development of heterotro-
gradient decreased and became equal to that of nitrate. phic aerobic and anaerobic bacteria as well as nitrifiers
At lower depths, nitrate, instead of oxygen, was the more pronounced. Chemolithotrophic bacteria, respon-
main electron acceptor in OM degradation. The reason sible for the intensity of redox processes, should be
for the decrease in the O2 gradient may have been the especially abundant in winter.
decrease in O2 consumption during OM mineralization The nitrogen balance of seas with anoxic conditions
in conditions of low O2 concentrations (Naqvi, 2006). includes compensation for the high level of denitrifica-
tion by N fixation, the intensity of which is, in turn,
4.2. Depth of O2 depletion influenced by the development of anoxic zones.
The scenarios included in the model offer an ex-
In the middle of the redox zone, the concentrations of planation for phosphate-dipole formation, as the results
oxidized chemical compounds diffusing downwards from indicated that Mn(III) accounts for a large fraction of the
the upper layer (O2 and NO3) decreased to zero. This dissolved Mn at that depth.
occurred simultaneously with the depletion of reductants The model also suggested that the redox zone
(“deep” NH4, Mn(II)) diffusing upwards from the anoxic contains a layer in which O2 and H2S are absent,
zone. A minimum of PO4 was also found in this middle which allows the presence of Mn, in the form of Mn(III)
region. According to the model, the redox reactions that and the anammox reaction. A dominant role in the
depleted the remaining O2 also resulted in the disappear- oxidation of H2S at its onset level was ascribed to
ance or a decrease of the above-mentioned reductants and oxidized Mn compounds. H2S oxidation occurs through
the formation of alternative electron acceptors, dissolved the reduction of several compounds in varying amounts:
Mn(III) and particulate Mn(IV) and Fe(III). In addition to Mn(IV) — 50%, Mn(III) — 43%, NO3 — 5%, Fe(III) —
their transport by diffusion, these latter compounds have a 1%, and O2 — 1%. By contrast, iron cycling in the
sinking rate that accelerates their downward transport. formation of the redox layer structure was determined to
Thus, a PO4 minimum could be established, depending on be insignificant.
the extent of its complexation with Mn(III). Our The results of this study allow us to conclude that
experiments showed that other factors, such as co-pre- mathematical modeling of the redox system is a useful
cipitation with Fe and Mn hydroxides and chemosynthe- tool for filling in the gaps in our knowledge and for
sis, are less important. determining the direction of further studies, including
analyses of redox-system reactions and the effects of
4.3. Depth of the H2S onset natural and anthropogenic forces upon them.

The peaks of oxidized Mn and Fe are located at this Acknowledgements


depth. There is an intense reduction of Mn(III) and Mn
(IV) by sulfide in reactions that are balanced by the H2S We appreciate the continuous support and critical and
flux from below. Either a decrease in the PO4 vertical useful discussions with our colleagues from the Baltic Sea
gradient or the formation of the deep phosphate maxi- Research Institute Warnemuende and the Shirshov Institute
mum occurs as a result of the degradation of phosphate of Oceanology, RAS. This research was supported by the
complexes with reducing Mn(III). Baltic Sea Research Institute Warnemuende, the Shirshov
The ability of this 1D model to reproduce the main Institute of Oceanology, Russian Foundation for Basic
features of the chemical structures of the central Black Sea Research grants 06-05-96676 yug, 07-05-01024 and
and the Gotland Deep of the Baltic Sea suggests that the CRDF grant RUG1-2828-KS06.
404 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

Appendix A
Table A1
Parameterization of the biogeochemical processes RatesBG
Name of process/reaction Parameterizations
Autolysis Autolis P = KPD POP
Autolis N = KPD ⁎ PON
Mineralization at oxic conditions DcDM_O2 = exp(Ktox ⁎ T ) ⁎ KND4 ⁎ DON ⁎ Fox
(CH2O)106(NH3)16H3PO4 + 06O2 = DcPM_O2 = exp(Ktox ⁎ T ) ⁎ KNP4 ⁎ PON ⁎ Fox
106CO2 + 16NH3 + H3PO4 + 106H2O 8
<¼ 0 for O2 VO2 ox
Fox ¼ ¼ ðO2  O2 oxÞ
: for O2 NO2 ox
ðO2  O2 ox þ KoxÞ
Denitrification Denitr1_PM = KN32 ⁎ Fdnox ⁎ FdnNO3 ⁎ PON
(CH2O)106(NH3)16H3PO4 + 84.8HNO3 = Denitr2_PM = KN24 ⁎ Fdnox ⁎ FdnNO2 ⁎ PON
106CO2 + 42.4N2 + 148.4H2O + 16NH3 + Denitr1_DM = KN32 ⁎ Fdnox ⁎ FdnNO3 ⁎ DON
H3PO4 Denitr2_DM = KN24 ⁎ Fdnox ⁎ FdnNO2 ⁎ DON
8
1/2CH2O + NO−3 → NO−2 + 1/2H2O + 1/2CO2 <¼ 0 for O2 NO2 dn
Fdnox ¼ ¼ 1  O2
O2 VO2 ox
: for
O2 dnðO2 dn þ 1  O2 Þ
3/4CH2O + H+ + NO−2 → 1/2N2 + 5/4H2O + 3/ (
¼0 for NO3 VNO3 mi
4CO2 FdnNO3 ¼ NO3  NO3 mi
¼1 for NO3 NNO3 mi
NO3  NO3 mi þ 1
(
¼0 for NO2 VNO2 mi
FdnNO2 ¼ NO2  NO2 mi
¼1 for NO2 NNO2 mi
NO2  NO2 mi þ 1
DcPM_NO3 = Denitr1_PM + Denitr2_PM
DcDM_NO3 = Denitr1_DM + Denitr2_DM
Sulfatereduction s4_rd_PM = K_s4_rd Fsox Fsnx SO4 PON
(CH2O)106(NH3)16H3PO4 + 53SO2−
4 = s4_rd_DM = K_s4_rd Fsox Fsnx SO4 DON
106CO2 + 106H2O + 16NH3 + H3PO4 + s23_rd_PM = K_s23_rd Fsox Fsnx PON S2O3
53S2− s23_rd_DM = K_s23_rd Fsox Fsnx DON S2O3

¼ 0 for O2 NO2 sr
Fsox ¼
¼ 1 for O2 VO2 sr

¼ 0 for ðNO3 þ NO2 ÞNNosr
Fsnx ¼
¼ 1 for ðNO3 þ NO2 ÞVNosr
DcPM_SO4 = s23_rd_PM + s4_rd_PM
DcDM_SO4 = s23_rd_DM + s4_rd_DM
Ammonification and phosphatification AmmonPON = DcPM_O2 + DcPM_NO3 + DcPM_SO4
AmmonDON = DcDM_O2 + DcDM_NO3 + DcDM_SO4
PhosPOP = AmmonPON / 16
PhosDOP = AmmonDON / 16
Nitrification
NH+4 + 1.5 O2 → NO−2 + 2H+ + H2O Nitrif1 = KN42 ⁎ NH4 ⁎ O2 / (O2 + O2nf)
NO−2 + 0.5 O2 → NO−3 Nitrif2 = KN23 ⁎ NO2 ⁎ O2 / (O2 + O2nf)
Nitrogen fixation 1 PO4
Nfixation ¼ Kmax  4 KNF dfi ðiÞdft ðT ÞdPhydSn
Nfix
NO3 þNO2 þNH4 PO4 þ 0:3
1þ 16PO4

Anammox NO2 + NH4 → N2 + 2H2O Anammox = NO2 ⁎ NH4 ⁎ Kannamox


Oxidation of reduced S forms with oxygen
2H2S + O2 → 2S0 + 2H2O hs_ox = K_hs_ox ⁎ H2S ⁎ O2
2S0 + O2 + H2O → S2O2− 3 + 2H
+
s0_ox = K_s0_ox ⁎ S0 ⁎ O2
S2O2−
3 + 2O 2 + 2OH −
→ 2SO 2−
4 + H2O s23_ox = K_s23_ox ⁎ S2O3 ⁎ O2
S0 disproportionation 4S0 + 3H2O → 2H2S + Disprop = Kdisp ⁎ S0
S2O2−3 + 2H
+

Thiodenitrification hs_NO3 = KT ⁎ H2S ⁎ NO3


3H2S + 4NO−3 + 6OH− → 3SO2− 4 + 2N2 + 6H2O hs_NO2 = KT ⁎ H2S ⁎ NO2
Mn oxidation and reduction
4Mn2+ + O2 + 4H+ → 4Mn3+ + 2H2O mn_ox = K_mn_ox ⁎ O2 ⁎ Mn(II)
4Mn3+ + O2 +6OH− → 4MnO2+6H2O mn_ox2 = K_mn_ox2 ⁎ O2 ⁎ Mn(III)
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 405

Table A1 (continued)
Name of process/reaction Parameterizations

2MnO2 + 7H + HS → 2Mn + 4H2O
+ 3+
mn_rd = K_mn_rd ⁎ Mn(IV) ⁎ H2S
2Mn3+ + HS− → 2Mn2+ + S0 + H+ mn_rd2 = K_mn_rd2 ⁎ Mn(III) ⁎ H2S
Fe oxidation and reduction
4Fe2+ + O2 + 2H2O → 4Fe3+ + 4OH− fe_ox = K_fe_ox ⁎ Fe(II) ⁎ O2
2Fe+2 + MnO2 + 2H2O → FeOOH + Mn2+ + 2H+ fe_mnox = K_fe_mnox ⁎ Fe(II) ⁎ Mn(IV)
10Fe2+ + 2NO−3 + 12H+ = 10Fe3+ + N2 + 6H2O fe_nox = K_fe_nox ⁎ Fe(II) ⁎ NO3
2FeOOH + H2S → 2Fe2+ + S0 + 4OH− fe_rd = K_fe_rd ⁎ Fe(III) ⁎ H2S
P sorption/desorption and complexation Coprecip = (fe_rd − fe_ox − fe_mnox) / 2.7 − (mn_ox − mn_ox2 + mn_rd − mn_rd2) / 0.66
Phy growth rate GrowthPhy = KNF ft(T)fi(i)min {fP(PO4), fN(NO3NO2, NH4)}
ft(T)=0.2 + 0.22(exp(0.21T) - 1) / (1 + 0.28exp(0.21T))
 
I0 I0
fi ðiÞ ¼ fu ðuÞ expðkhÞexp 1  expðkhÞ
Iopt Iopt
PO4
fp ðPO4 Þ ¼
KPO4 þ PO4
ðNO3 þ NO2 ÞexpðKpsi NH4 Þ NH4
fN ðNO3 ; NO2 ; NH4 Þ ¼ fNV ðNO3 ; NO2 Þ þ fNW ðNH4 Þ ¼ þ
KNO3 þ ðNO3 þ NO2 Þ KNH4 þ NH4
Phy excretion rate ExcrPhy = KFD ⁎ Phy
Phy mortality rate MortPhy = KFP ⁎ Phy
Grazing of Zoo Grazing = GrazPhy + GrazPOP + GrazBact
Grazing of Zoo on Phy GrazPhy = KFZ ⁎ Zoo ⁎ (Phy / Zoo) / (Phy / Zoo + KF)
Grazing of Zoo on detritus GrazPOP = KPZ ⁎ Zoo ⁎ (POP / Zoo) / (POP / Zoo + KPP / 0.001)
Grazing of Zoo on bacteria GrazBact = GrazBhe + GrazBae + GrazBhat + GrazBaa
GrazBhe = KBhaZ ⁎ Zoo ⁎ (Bhe / Zoo) / (Bhe / Zoo + KBhe)
GrazBae = KBaeZ ⁎ Zoo ⁎ (Bae / Zoo) / (Bae / Zoo + KBae)
GrazBha = KBhaZ ⁎ Zoo ⁎ (Bha / Zoo) / (Bha / Zoo + KBha)
GrazBaa = KBaaZ ⁎ Zoo ⁎ (Baa / Zoo) / (Baa / Zoo + KBaa)
Mortality of Zoo MortZoo = KZP ⁎ Zoo ⁎ Zoo
Growth rate of Bhe CBhe = Kmax
Bhe · (DcPM_O2 + DcDM_O2) · fBhe(DON + PON) · Bhe
PON þ DON
fBhe ðDON þ PONÞ ¼
PON þ DON þ KNBhe
Rate of mortality of Bhe MortBhe ¼ KMort 2
Bhe Bha
Growth rate of Bae CBae = Kmax 0 NP
Bae ·(Nitrif 1 + Nitrif 2 + S _ox + S2O3_ox + mn_ox ++ fe_ox)f Bae(NH4, PO4)·Bae
 
NH4 PO4
NP
fBae ð NH4 ; PO4 Þ ¼ min N ; PO þ K P
NH4 þ KBae 4 Bae
Rate of mortality of Bae MortBae = KMort
Bae Bae
2
max
Growth rate of Bha CBha = KBha (DcPM_NO3 + DcDM_NO3 + DcPM_SO4 ++ DcDM_SO4)fBha(DON + PON)Bha
PON þ DON
fBha ðDON þ PONÞ ¼
PON þ DON þ KBha
N

Rate of mortality of Bha MortBha = KMort


Bhat Bha
2

Growth rate of Baa CBaa = KBaa (mn_rd + fe_rd + hs_ox + hs_NO3 + hs_NO2 ++ anammox)f NP
max
Baa(NH4, PO4)Baa
 
NH4 PO4
fBaa ðNH4 ; PO4 Þ ¼ min
NP
N ;
NH4 þ KBaa PO4 þ KPBaa
Rate of mortality of Baa MortBaa = KMort
Baa Baa
2

Table A2
Parameters names, notations, values and units of the coefficients used in the model
Parameter Notation Units Value
−1
Specific rate of decomposition of POM to DOM KPD d 0.10
Mineralization in oxic conditions
Specific rate of decomposition of DON and DOP KND4 d− 1 0.1
Specific rate of decomposition of PON and POP KNP4 d− 1 0.04
(continued on next page)
406 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

Table A2 (continued)
Parameter Notation Units Value
Mineralization in oxic conditions
Temperature parameter for oxic mineralization Ktox °C− 1 0.15
Oxygen parameter for oxic mineralization O2ox μM 0
Half saturation constant for oxic mineralization Kox μM 15
Denitrification
Specific rate of 1st stage of denitrification KN32 d− 1 0.12
Specific rate of 2d stage of denitrification KN24 d− 1 0.20
Oxygen parameter for denitrification O2dn μM 25
NO3 parameter for denitrification NO3mi μM 0.001
NO2 parameter for denitrification NO2mi μM 0.0001
Sulfate reduction
Specific rate of sulfate reduction with sulfate K_s4_rd d− 1 2.5 * 10−7
Specific rate of sulfate reduction with thiosulfate K_s23_rd d− 1 1.2
Oxygen parameter for sulfate reduction O2sr μM 25
NO3 and NO2 parameter for sulfate reduction NOsr μM 0.5
Nitrification
Specific rate of the 1st stage of nitrification KN42 d− 1 0.9
Specific rate of the 2d stage of nitrification KN23 d− 1 2.5
Oxygen parameter for nitrification O2nf μM 1
Nitrogen fixation
Specific rate of nitrogen fixation Kmax
Nfix d− 1 20
Anammox
Anammox constant Kannamox d− 1 0.03
Oxidation of the hydrogen sulfide
Specific rate of oxidation of H2S with O2 K_hs_ox d− 1 0.2
Specific rate of oxidation of S0 with O2 K_s0_ox d− 1 4.0
Specific rate of oxidation of S2O3 with O2 K_s23_ox d− 1 1.5
S0 disproportionation
Specific rate of S0 disproportionation Kdisp d− 1 0.01
Thiodenitrification
Thiodenitrification constant KT μM− 1d− 1 0.8
Oxidation and reduction of Mn and Fe
Mn(II) oxidation with O2 constant K_mn_ox d− 1 2
Mn(IV) reduction with Sulfide constant K_mn_rd d− 1 22
Mn(III) oxidation with O2 constant K_mn_ox2 d− 1 18
Mn(IV) reduction with sulfide constant K_mn_rd2 d− 1 2
Fe oxidation with O2 constant K_fe_ox d− 1 4
Fe oxidation with Mn(IV) constant K_fe_mnox d− 1 1
Fe oxidation with NO3 constant K_fe_nox d− 1 5
Fe(III) reduction by sulfide K_fe_rd d− 1 0.05
Phytoplankton
Maximum specific growth rate KNF d− 1 1.86
Specific respiration rate KFN d− 1 0.05
Incident light I0 W m− 2 80
Optimal light Iopt W m− 2 25
Extinction coefficient K m− 1 0.07
Half-saturation constant for uptake of PO4 KPO4 μM 0.01
Strength of ammonium inhibition of nitrate uptake constant Kpsi 1.46
Half saturation constant for uptake of NH4 KNH4 μM 0.02
Half saturation constant for uptake of NO3 + NO2 KNO3 μM 0.03
Specific rate of mortality KFP d− 1 0.05
Specific rate of excretion KFD d− 1 0.05
Zooplankton
Specific respiration rate KZN d− 1 0.1
Maximum specific rate of grazing of Zoo on Phy KFZ d− 1 0.5
Half-saturation constant for the grazing of Zoo on Phy for KF 1
Phy/Zoo ratio
Maximum specific rate of grazing of Zoo on POP KPZ d− 1 0.6
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 407

Table A2 (continued)
Parameter Notation Units Value
Zooplankton
Half-saturation constant for the grazing of Zoo on POP in KPP 200
dependence to ratio POP/Zoo
Maximum specific rate of grazing of Zoo on Bae KBaeZ d− 1 0.6
Half-saturation constant for the grazing of Zoo on Bae for KBae 1.5
Bae/Zoo ratio
Maximum specific rate of grazing of Zoo on Bhe KBheZ d− 1 1.02
Half-saturation constant for the grazing of Zoo on Bhe for KBhe 1.1
Bhe/Zoo ratio
Maximum specific rate of grazing of Zoo on Baa KBaaZ d− 1 0.78
Half-saturation constant for the grazing of Zoo on Baa for KBaa 1.5
Baa/Zoo ratio
Maximum specific rate of grazing of Zoo on Bha KBhaZ d− 1 0.6
Half-saturation constant for the grazing of Zoo on Bha for KBha 1
Bha/Zoo ratio
Maximum specific rate of mortality of Zoo KZP d− 1 0.001 for: H2S ≤ 20 μM
0.9 for: H2S N 20 μM
Food absorbency for zooplankton Uz 0.7
Ratio between dissolved and particulate excretes of Hz 0.6
zooplankton
Aerobic heterotrophic bacteria
Maximum specific growth rate of Bhe Kmax
Bhe μM− 1 2
Half-saturation constant for the dependence of maximum KN
Bhe μM 0.5
specific growth rate of Bhe on POM and DOM content.
Maximum specific rate of mortality of Bhe KMort
Bhe d− 1 0.03 for: O2 N 1 μM
0.99 for: O2 ≤ 1 μM
Aerobic autotrophic bacteria
Maximum specific growth rate of Bae Kmax
Bae μM− 1 1
Half-saturation constant for the dependence of maximum KN
Bae μM 0.05
specific growth rate of Bae on NH4
Half-saturation constants for the dependence of maximum KPBae μM 0.3
specific growth rate of Bae on PO4
Maximum specific rate of mortality of Bae KMort
Bae d− 1 0.01 for: O2 N 1 μM
0.99 for: O2 ≤ 1 μM
Anaerobic heterotrophic bacteria
Maximum specific growth rate of Bha Kmax
Bha μM− 1 2
Half-saturation constant for the dependence of maximum KN
Bha μM 6
specific growth rate of Bha on POM and DOM
Maximum specific rate of mortality of Bha KMort
Bha d− 1 0.01
Anaerobic autotrophic bacteria
Maximum specific growth rate of Baa Kmax
Baa μM− 1 6.5
Half-saturation constants for the dependence of maximum KN
Baa μM 3
specific growth rate of Baa on NH4
Half-saturation constants for the dependence of maximum KPBaa μM 3
specific growth rate of Baa on PO4
Maximum specific rate of mortality of Baa KMort
Baa d− 1 0.001 for: H2S ≤ 16 μM
0.99 for: H2S N 16 μM
Sinking
Rate of sinking of Phy WPhy m d− 1 0.5
Rate of sinking of Zoo WZoo m d− 1 1.0
Rate of sinking of bacteria (Bhe,Bae,Bha,Baa) WBact m d− 1 0.5
Rate of sinking of detritus (POP, PON) WD m d− 1 6
Rate of accelerated sinking of particles with settled Mn Wmax
Me m d− 1 18
hydroxides
408 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

Table A3
Rates of biogeochemical production/consumption of the model compartments Ri
Phosphate (PO4) RPO4 = Sp(GrowthPhy(KFN − 1.) − Chemos − ChemosA + KZNZoo) + PhosPOP + PhosDOP +
Coprecip
Dissolved organic phosphorus (DOP) RDOP = Sp(ExcrPhy + Grazing(1. − Uz)Hz + 0.7MortBact − (Hetero + HeteroA) − DON / (DON +
PON)(Hetero + HeteroA)) + AutolisP − PhosDOP
Particulate organic phosphorus (POP) RPOP = Sp(MortPhy + MortZoo + 0.3MortBact + Grazing(1. − Uz)(1. − Hz) − GrazPOP − PON /
(DON + PON) ⁎ (Hetero + HeteroA)) − AutolisP − PhosPOP
Particulate organic nitrogen (PON) RPON = Sn(MortPhy + MortZoo + 0.3MortBact + Grazing(1. − Uz)(1. − Hz) − GrazPOP − PON /
(DON + PON)(Hetero + HeteroA))−AutolisN − AmmonPON
Dissolved organic phosphorus (DON) RDON = Sn(ExcrPhy + Grazing(1. − Uz)Hz + 0.7MortBact − (Hetero + HeteroA)) + AutolisN −
AmmonDON
Ammonia (NH4) RNH4 = Sn ⁎ (GrowthPhy(KFN − 1.)(LimNH4 / LimN) − Chemos − ChemosA + KZN ⁎ Zoo)
AmmonPON + AmmonDON − Nitrif1 + Nfixation
Nitrite (NO2) RNO2 = Sn ⁎ (GrowthPhy ⁎ (KFN − 1.) ⁎ (LimNO3 / LimN) ⁎ (NO2 / (NO2 + NO3))) + Nitrif1−Nitrif2 +
Denitr1 − Denitr2 − sulfido2
Nitrate (NO3) RNO3 = Sn ⁎ (GrowthPhy ⁎ (KFN − 1.) ⁎ (LimNO3 / LimN) ⁎ (NO3 / (NO2 + NO3))) + Nitrif2−Denitr1−
1.25sulfido − ox− fe_nox
Oxygen (O2) RO2 = OkP Sp GrowthPhy + 2.Sn GrowthPhy(NO3 / (NO2 + NO3))(LimNO3 / LimN) + 0.5Sn
GrowthPhy (NO2 / (NO2 + NO3))(LimNO3 / LimN) − Destr_OM − OkP Sp(GrowthPhy KFN +
KZN Zoo) − 1.5Nitrif1 − 0.5Nitrif2 − 0.5hs_ox − 1.s0_ox − 2.s23_ox − 1.mn_ox − fe_ox
Hydrogen sulphide (H2S) RH2S = − hs_ox + s23_rd − 0.5fe_rd − mn_rd − sulfido − sulfido2 + 0.5Disprop
Elemental sulphur (S0) RS0 = hs_ox − s0_ox + 1.mn_rd − Disprop
Thiosulfate (S2O3) RS2O3 = s0_ox − s23_ox + s4_rd − s23_rd + 0.5Disprop
Sulfate (SO4) RSO4 = s23_ox − s4_rd + sulfido + sulfido2
Bivalent manganese (Mn(II)) RMn2 = mn_rd2 − mn_ox + 0.5 ⁎ fe_mnox
Quadrivalent manganese (Mn(IV)) RMn4 = mn_ox2 − mn_rd − 0.5 ⁎ fe_mnox
Trivalent manganese (Mn(III)) RMn3 = mn_ox − mn_ox2 + mn_rd − mn_rd2
Bivalent iron (Fe(II)) RFe2 = fe_rd − fe_ox − fe_mnox − 5. ⁎ fe_nox
Trivalent iron (Fe(III)) RFe3 = fe_ox + fe_mnox + 5. ⁎ fe_nox − fe_rd
Phytoplankton (Phy) RPhy = GrowthPhy(1 − KFN) − MortPhy − ExcrPhy − GrasPhy
Zooplankton (Zoo) RZoo = Grazing ⁎ Uz − MortZoo − KZN ⁎ Zoo
Aerobic heterotrophic bacteria (Bhe) RBhe = CBhe − MortBhe − GrazBhe,
Aerobic autotrophic bacteria (Bae) RBae = CBae − MortBae − GrazBae,
Anaerobic heterotrophic bacteria (Bha) RBha = CBha − MortBha − GrazBha,
Anaerobic autotrophic bacteria (Baa) RBaa = CBaa − MortBaa − GrazBaa,
Where Sn = 0.016, Sp = 0.001 are ratios between N and P content and the wet weight, OkP= 106 is the O2:P molar ratio.

References Fuiman, L.A. (Eds.), Advances in Marine Biology, vol. 48.


Elsevier Academic Press, Amsterdam, p. 640.
Ayzatulllin, T.A., Leonov, A.V., 1975. Kinetika i mekhanizm Dalsgaard, T., Canfield, D.E., Peterson, J., Thamdrup, B., Acuna-
okislitslnoi transformatsii soedineniy rastvoryonnoy sery v Gonzales, J., 2003. N production by anamox in the anoxic water
morskoi vode (Kinetics and mechanism of the oxidizing column of Golfo Dulce, Costa Rica. Nature 422, 606–608.
transformation of reduced sulfur compounds in the sea water). Davies, G., 1969. Some aspects of chemistry of Mn(III) in aqueous
Okeanologiya 15, 1026–1033 (in Russian). solutions. Coordination Chemistry Reviews 4, 199–224.
Bauer, S., 2003. Structure and function of nitrifying bacterial Detmer, A.E., Giesenhagen, H.C., Trenkel, V., Auf dem Venne, H.,
communities in the eastern Gotland basin (central Baltic Sea). Jochem, F.J., 1993. Phototrophic and heterotrophic pico- and
Ph.D.Thesis, Mathematisch-Naturwissenschaftliche Fakultat. Uni- nanoplankton in anoxic waters of the Central Baltic Sea. Marine
versitat Rostock, Rostock. Ecology. Progress Series 99, 197–203.
Boudreau, B.P., 1996. A method-of-lines code for carbon and nutrient Enoksson, V., 1986. Nitrification rates in the Baltic Sea: comparison of
diagenesis in aquatic sediments. Computers and Geosciences 22, tree isotope techniques. Applied and Environmental Microbiology
479–496. 51, 244–250.
Brettar, I., Rheinheimer, G., 1991. Denitrification in the Central Baltic: Erdogan, S., Yemenicioglu, S., Tugrul, S., 2003. Distribution of
evidence for H2S-oxidation as motor of denitrification at the oxic- dissolved and particulate forms of iron and manganese in
anoxic interface. Marine Ecology. Progress Series 77, 157–169. the Black Sea. In: Yilmaz, A. (Ed.), Oceanography of the
Burchard, H., Bolding, K., Villareal, M.R., 1999. GOTM, a general Eastern Mediterranean and Black Sea. Tubitak Publishers, Ankara,
ocean turbulence model. Theory, applications and test cases. pp. 447–451.
European Commission Report EUR 18745 EN: 103. Fasham, M.J., Ducklow, H.W., McKelvie, S.M., 1990. A nitrogen-
Canfield, D.E., Thamdrup, B., Kristensen, E., 2005. Aquatic based model of plankton dynamics in the oceanic mixed layer.
geomicrobiology. In: Southward, A.J., Tyler, P.A., Young, C.M., Journal of Marine Research 48, 591–639.
E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410 409

Fenchel, T., King, G.M., Blackburn, T.H., 1998. Bacterial Biogeo- Oguz, T., Ducklow, H., Shushkina, E.A., Malonotte-Rizzoli, P., Tugrul,
chemistry: The Ecophysiology of Mineral Cycling. Academic S., Lebedeva, L.P., 1998. Simulation of upper layer biochemical
Press, San Diego, p. 307. structure in the Black Sea. In: Ivanov, L., Oguz, T. (Eds.), NATO
Fennel, W., Neumann, T., 2004. Introduction to the modeling of TU-Black Sea Project. Ecosystem Modeling as a Tool for the Black
marine ecosystems. In: Halpne, D. (Ed.), Elsevier Oceanographic Sea. Symp. on Sci. Res., vol. 2. Kluwer Academic Publishers,
Series, 72, p. 298. Amsterdam. Norwell, pp. 257–299.
Fonselius, S.H., 1974. Phosphorus in the Black Sea. In: Degens, E.J., Overmann, J., Manske, A.K., 2005. Anoxygenic phototrophic bacteria
Koss, D.A. (Eds.), The Black Sea — Geology, Chemistry and in the Blcak Sea chemocline. In: Neretin, L.N. (Ed.), Past and
Biology. Amer. Ass. of Petrol. Geologists, Tusla, pp. 144–150. Present Water Column Anoxia. NATO Sciences Series. Springer,
Gargett, A.E., 1984. Vertical eddy diffusivity in the ocean interior. Dordrecht, pp. 501–522.
Journal of Marine Research 42, 359–393. Pakhomova, S.V., 2005. Rastvoryonnye formy zheleza i margantsa v
Gregoire, M., Beckers, J.-M., Nihoul, J.C.J., Stanev, E., 1997. Coupled morskoi vode, osadkakh i na granitse voda-dno (Dissolved forms
hydrodynamic ecosystem model of the Black Sea at basin scale. In: of iron and manganese in marine water, sediments and the water-
Ozsoy, E., Mikaelyan, A. (Eds.), Sensitivity to Change: Black Sea, bottom boundary). Ph.D.Thesis, SIO RAS, Moscow, unpublished.
Baltic Sea and North Sea. Kluwer, Netherlands, pp. 487–499. (in Russian).
Hannig, M., Lavik, G., Kuypers, M., Wobken, D., Jurgens, K., 2006. Pimenov, N.G., Neretin, L.N., 2006. Composition and activities of
Distribution of denitrification and anammox activity in the water microbial communities, involved in carbon, sulfur, nitrogen and
column of the central Baltic Sea. 9th Int. Estuarine Biogeochemistry manganese cycling in the oxic/anoxic interface of the Black Sea.
Symposium. Estuaries and Enclosed Seas under Changing Environ- In: Neretin, L.N. (Ed.), Past and Present Water Column Anoxia.
mental Conditions, May 7–11. IOW, Warnemuende, Germany, p. 49. NATO Sciences Series. Springer, Dordrecht, pp. 501–522.
HELCOM, 2002. Environment of the Baltic Sea area 1994–1998. Pshenin, L.N., 1963. Distribution and ecology of Azotobacter in the Black
Baltic Sea Environment Proceedings 82B, 216. Sea. In: Oppenheimer, C.H. (Ed.), Symposium on Marine Microbiol-
Jorgensen, B.B., Fossing, H., Wirsen, C.O., Jannasch, H.W., 1991. ogy. Charles C Thomas, Publisher, Springfield, Ill, pp. 383–391.
Sulfate oxidation in the anoxic Black Sea chemocline. Deep-Sea Richards, F.A., 1965. Anoxic basins and fjords. In: Riley, J.P., Skirrow,
Research. Part 2. Topical Studies in Oceanography 38, 1083–1104. G. (Eds.), Chemical Oceanography, vol. 1. Academic Press, NY,
Konovalov, S.K., Murray, J.W., Luther, G.W., Tebo, B.M., 2006. pp. 611–645.
Processes controlling the Redox budget for oxic/anoxic water Samodurov, A.S., Ivanov, L.I., 1998. Processes of ventilation of the
column of the Black Sea. Deep-Sea Research. Part 2. Topical Black Sea related to water exchange through the bosphorus. In:
Studies in Oceanography 53, 1817–1841. Ivanov, L., Oguz, T. (Eds.), NATO TU-Black Sea Project.
Kostka, J.E., Luther III, G.W., Nealson, K.H., 1995. Chemical and Ecosystem Modeling as a Tool for the Black Sea. Symposium on
biological reduction of Mn(III)-pyrophosphate complexes: poten- Scientific Results, vol. 2/47(2). P.II. Kluwer Academic Publishers,
tial importance of dissolved Mn(III) as an environmental oxidant. Netherland, pp. 221–236.
Geochimica et Cosmochimica Acta 59, 885–894. Savchuk, O., Wulff, F., 1996. Biogeochemical transformation of nitrogen
Kuypers, M.M.M., Sliekers, A.O., Lavik, G., Schmid, M., Jorgensen, B.B., and phosphorus in the marine environment. Coupling Hydrodynamic
Kuenen, J.G., Sinnenghe Damste, J.S., Strous, M., Jetten, M.S.M., and Biogeochemical Processes in Models for the Baltic Proper. Systems
2003. Anaerobic ammonium oxidation by anammox bacteria in the Ecology Contributions, vol. 2. Stockholm University, Stockholm.
Black Sea. Nature 422, 608–611. Savenko, A.V., 1995. Precipitation of phosphate with iron hydroxide
Lewis, B.L., Landing, W.M., 1991. The biogeochemistry of manganese forming by mixing of submarine hydrothermal solutions and the
and iron in the Black Sea. Deep-Sea Research. Part 2. Topical sea water (on the base of experimental data). Geochemistry
Studies in Oceanography 38, 773–803. International 9, 1383–1389.
Morgan, J.A., Quinby, H.L., Ducklow, H.W., 2006. Bacterial Savenko, A.V., Baturin, G.N., 1996. Experimental study of the
abundance and production in the Western Black Sea. Deep-Sea sorption of phosphorus on manganese dioxide. Geochemistry
Research. Part 2. Topical Studies in Oceanography 53, 1945–1960. International 5, 472–474.
Murray, J.W., Codispoti, L.A., Friederich, G.E., 1995. The suboxic Schneider, B., Nausch, G., Kubsch, H., Peterson, I., 2002. Accumu-
zone in the Black Sea. In: Huang, C.P., O'Melia, R., Morgan, J.J. lation of total CO2 during stagnation in the Baltic deep water and
(Eds.), Aquatic Chemistry: Interfacial and Interspecies Processes. its relationship to nutrient and oxygen concentrations. Marine
ACS Advances in Chemistry Series, vol. 224, pp. 157–176. Chemistry 77, 277–291.
Naqvi, S.W.A., 2006. Oxygen deficiency in the North Indian Ocean. Shaffer, G., 1986. Phosphorus pumps and shuttles in the Black Sea,
Gayana 70, 53–58 (suplemento). Letters to Nature. Nature 321, 515–517.
Nealson, K.N., Stahl, D.A., 1997. Microorganisms and biogeochemical Sorokin, Yu.I., 2002. The Black Sea. Ecology and Oceanography.
cycles: what can be learn from layered microbial communities? In: Backhuys Publishers, Leiden, p. 875.
Banfield, J.F., Nealson, K.N. (Eds.), Geomicrobiology: Inter- Sorokin, Yu.I., Sorokin, D.Yu., Avdeev, V.A., 1991. Aktivnost' mikroflory
actions Between Microbes and Minerals. Reviews in Mineralogy, i okislitel'nye protsessy sernogo tsykla v tolshche vody Chernogo
vol. 35. Mineralogical Society of America, Washington D.C., morya (Microbial activity and sulfur cycle oxidation processes in the
pp. 5–34. Black Sea water column). In: Vinogradov, M. (Ed.), Izmenchivost'
Nealson, K.H., Myers, C.R., Wimpee, B.B., 1991. Isolation and Ekosystemy Chernogo Morya (Estesstvennye i Antropogennyye
identification of manganese reducing bacteria and estimates of Faktory). Nauka, Moscow, pp. 173–188 (in Russian).
microbial Mn (IV)-reducing potential in the Black Sea. Deep-Sea Stal, L.J., Walsby, A.E., 2000. Photosynthesis and nitrogen fixation in
Research. Part 2. Topical Studies in Oceanography 38, 907–920. a cyanobacterial bloom in the Baltic Sea. European Journal of
Neretin, L., Pohl, C., Jost, G., Leipe, T., Pollehne, F., 2003. Manganese Phycology 35, 97–108.
cycling at the oxic/anoxic interface in the Gotland deep, Baltic Sea. Stokozov, N.A., 2004. Long-lived radionuclides 137Cs and 90Sr in the
Marine Chemistry 82, 125–143. Black Sea after the Chernobyl NPP accident and their use as a
410 E.V. Yakushev et al. / Marine Chemistry 107 (2007) 388–410

tracers of water exchange processes. Ph.D.Thesis., MHI, Sebas- Environmental Degradation of the Black Sea: Challenges and
topol, unpublished. (in Russian). Remedies. Kluwer Academic Publishers, Dordrecht, pp. 93–108.
Tebo, B.M., 1991. Manganese (II) oxidation in the suboxic zone of the Yakushev, E.V., Debolskaya, E.I., 2000. Particulate manganese as a
Black Sea. Deep-Sea Research. Part 2. Topical Studies in main factor of oxidation of hydrogen sulfide in redox zone of the
Oceanography 38, 883–906. Black Sea. Proc. Konstantin Fedorov Memorial Symp. Oceanic
Trouwborst, R.E., Brian, G.C., Tebo, B.M., Glazer, B.T., Luther III, G.W., Fronts and Related Phenomena 18–22 May, 1998. Pushkin, Saint-
2006. Soluble Mn(III) in suboxic zones. Science 313 (5795), Petersburg, Russia. IOC Workshop Report, vol. 159. Kluwer
1955–1957. Academic Publishers, Dordrecht, pp. 592–597.
UNESCO, 1986. Progress on oceanographic tables and standards Yakushev, E.V., Mikhailovsky, G.E, 1995. Mathematical modeling of
1983–1986: work and recommendations of the UNESCO/SCOR/ the influence of marine biota on the carbon dioxide ocean–
ICES/IAPSO Joint Panel. Unesco Technical Papers in Marine atmosphere exchange in high latitudes. In: Jaehne, B., Monahan,
Science, ndeg., vol. 50, p. 59. E.C. (Eds.), Air–Water Gas Transfer, Selected Papers, Third Int.
Volkov, I.I., Rozanov, A.G., Demidova, T.P., 1992. Vosstanovlenniye Symp. July 24–27, Heidelberg University. AEON Verlag &
soedineniya sery i rastvorennogo margantsa v vode Chyornogo Studio, Hanau, pp. 37–48.
moray (Reduced inorganic sulfur species and dissolved manganese Yakushev, E.V., Neretin, L.N., 1997. One-dimensional modeling of
in the water of the Black Sea). In: Vinogradov, M.E. (Ed.), Zimnee nitrogen and sulfur cycles in the aphotic zones of the Black and
sostoyanie ekosystemy otkrytoi chasti Chernogo Morya. Shirshov Arabian Seas. Global Biogeochemical Cycles 11, 401–414.
Institute of Oceanology RAS, Moscow, pp. 38–50 (in Russian). Yakushev, E.V., Pollehne, F., Jost, G., Kuznetsov, I., Schneider, B.,
Ward, B.B., Kilpatrick, K.A., 1991. Nitrogen transformations in the oxic Umlauf, L., 2006a. Redox-Layer Model (ROLM): a tool for
layer of permanent anoxic basins: the Black Sea and the Cariaco analysis of the water column oxci/anoxic interface processes.
Trench. In: Izdar, E., Murray, J.W. (Eds.), Black Sea Oceanography. Meereswissenschaftliche Berichte, Marine Scince Report, vol. 68.
Kluwer Academic Publishers, Norwell, Mass, pp. 111–124. Instutut fur Ostseeforschung Warnemuende. 54 pp.
Wasmund, N., Nausch, G., Schneider, B., Nagel, K., Voss, M., 2005. Yakushev, E.V., Chasovnikov, V.K., Debolskaya, E.I., Egorov, A.V.,
Comparison of nitrogen fixation rates determined with different Makkaveev, P.N., Pakhomova, S.V., Podymov, O.I., Yakubenko,
methods: a study in the Baltic Proper. Marine Ecology. Progress V.G., 2006b. The northeastern Black Sea redox zone: hydro-
Series 297, 23–31. chemical structure and its temporal variability. Deep-Sea Research.
Webb, S.M., Dick, G.J., Bargar, J.R., Tebo, B.M., 2005. Evidence for Part 2. Topical Studies in Oceanography 53, 1764–1786.
the presence of Mn(III) intermediates in the bacterial oxidation of Yilmaz, A., Coban-Yildiz, Y., Telli-Karakoc, F., Bologa, A., 2006.
Mn(II). In: Fridovich, I. (Ed.), PNAS, vol. 102, pp. 5558–5563. Surface and mid-water sources of organic carbon by photo- and
Westrich, J.T., Berner, R.A., 1984. The role of sedimentary organic chemoautotrophic production in the Black Sea. Deep-Sea Re-
matter in bacterial sulfate reduction the G model tested. Limnology search. Part 2. Topical Studies in Oceanography 53, 1988–2004.
and Oceanography 29, 236–249. Zopfi, J., Ferdelman, T.G., Jorgensen, B.B., Teske, A., Thamdrup, B.,
Yakushev, E.V., 1992. Numerical modeling of transformation of 2001. Influence of water column dynamics on sulfide oxidation
nitrogen compounds in the redox zone of the Black Sea. and other major biogeochemical processes in the chemocline of
Oceanology 32, 173–177. Mariager Fjord (Denmark). Marine Chemistry 74, 29–51.
Yakushev, E.V., 1999. An approach to modeling anoxic conditions in
the Black Sea. In: Besiktepe, S., Unluata, U., Bologa, A. (Eds.),

S-ar putea să vă placă și