Subscriber access provided by Bethel University
Environmental Measurements Methods
Potential of Nitrogen/Argon Analysis in Surface Waters in the Examination of Areal Nitrogen Deficits Caused by Nitrogen Fixation Oliver Schmale, Mattis Karle, Michael Glockzin, and Bernd Schneider Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06665 • Publication Date (Web): 22 May 2019 Downloaded from http://pubs.acs.org on May 30, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 25
Environmental Science & Technology
1
Potential of Nitrogen/Argon Analysis in Surface Waters in the Examination of
2
Areal Nitrogen Deficits Caused by Nitrogen Fixation
3
Oliver Schmale*, Mattis Karle, Michael Glockzin, and Bernd Schneider
4
*Correspondence to:
5
Oliver Schmale (oliver.schmale@io-warnemuende)
6
Leibniz Institute of Baltic Sea Research (IOW), Seestrasse 15, D-18119 Rostock
7
1
ACS Paragon Plus Environment
Environmental Science & Technology
Page 2 of 25
8
ABSTRACT
9
In marine systems, the loss of nitrogen caused by denitrification in oxygen-deficient zones is
10
balanced by nitrogen fixation mediated by cyanobacteria, which may form extensive blooms in
11
surface waters. In this study, by determining the concentration ratio of nitrogen (N 2) and argon
12
(Ar) in air equilibrated with surface water we were able to detect changes in the N2 concentration
13
attributable to N2 fixation. For this purpose, surface water was pumped continuously into a spray-
14
type equilibrator while the air in the equilibrator’s headspace was analyzed by mass
15
spectrometry. After laboratory tests and model analysis to evaluate the sensitivity of our N2/Ar
16
approach, feasibility studies were conducted in the central Baltic Sea in the summer of 2015,
17
during the development of a cyanobacterial bloom. Our results showed that N2 deficits
18
accumulated during periods of low wind and increasing surface water temperatures. A
19
comparison of our results with the N2 deficits calculated from changes in the partial pressure of
20
carbon dioxide in surface water indicated a similar trend. By demonstrating the ability of the
21
N2/Ar approach to resolve N2 deficits in surface water caused by N2 fixation, our study
22
contributes to assessments of the N2 fixation efficiency of cyanobacterial blooms.
23
24 25 2
ACS Paragon Plus Environment
Page 3 of 25
Environmental Science & Technology
26
INTRODUCTION
27
The concentrations of dissolved gases in seawater are controlled by physical and biological
28
processes. The physical processes include temperature-, salinity-, and wind-driven gas exchange
29
with the atmosphere, bubble injection, and the interaction between different water bodies through
30
mixing. In marine biogeochemistry, the inert character of argon (Ar) is used to separate these
31
physical effects from the biological activities that affect gas concentrations. Because oxygen (O2)
32
and Ar have similar physicochemical properties, dissolved O2/Ar ratios can be employed to
33
estimate net community production and to resolve regional phenomena in the surface water
34
linked to photosynthesis and respiration.1, 2 Biological processes involved in the marine nitrogen
35
cycle and associated with changes in the concentration of molecular nitrogen (N2) include
36
denitrification, ammonium oxidation (anammox reaction), and N2 fixation.3 Analogous to
37
oxygen-related processes, the molecular nitrogen (N2) excess resulting from water column
38
denitrification and the anammox reaction in oxygen-deficient zones was estimated by measuring
39
N2/Ar ratios .4, 5
40
In the ocean, the loss of dissolved inorganic nitrogen (DIN) compounds due to denitrification and
41
anammox reactions is mainly balanced by N2 fixation, i.e., the conversion of molecular into
42
organic nitrogen, in reactions carried out by diazotrophic organisms. Most of the N2 fixation in
43
the central Baltic Sea and Gulf of Finland can be attributed to the cyanobacteria Nodularia
44
spumigena, Aphanizomenon spec. and Dolichospermum spec. 6, which form annually recurring
45
blooms from July to the beginning of August. There is general concern that eutrophication and
46
climate change will favor the proliferation of these harmful blooms.7 Consequently, an
47
understanding of their development is essential for the ecological management of the Baltic Sea
48
and other coastal areas. 3
ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 25
49
Among the methods used to estimate the contribution of N2 fixation, and therefore the activity of
50
bloom-forming cyanobacteria, to the nitrogen (DIN) budget of the Baltic Sea are: (i) the
51
incubation of discrete water samples with 15N-labeled molecular nitrogen to obtain instantaneous
52
fixation rates 6, 8, (ii) the establishment of basin-wide mass balances for total nitrogen 9, 10, (iii) the
53
deployment of measuring devices by voluntary observing ships (VOS) to obtain high-resolution
54
data on the carbon dioxide partial pressure (pCO2) in surface water, thus enabling determinations
55
of net community production and estimates of the nitrogen demand.11, 12
56
In this study, we demonstrate that the small changes in the N2 concentration arising from N2
57
fixation can be detected by a novel method based on mass spectroscopy of the N2/Ar ratio in air
58
equilibrated with surface water. To this end, a continuous-mode equilibrator system coupled to a
59
mass spectrometer (Equi-MS) was developed and its performance optimized and evaluated in
60
laboratory experiments. Then, using a simple physico-chemical model, we assessed the potential
61
and limitations of the N2/Ar approach for studies of the biogeochemical processes that decrease
62
or increase the N2 concentration. Finally, the feasibility of our N2/Ar approach in the detection of
63
N2 deficits in surface water caused by N2 fixation was tested in a field study in the central Baltic
64
Sea. The results were compared with those of an established approach for the quantification of N2
65
fixation capacities.
66
67
MATERIAL AND METHODS
68
Fundamentals and limitations of the N2/Ar method
4
ACS Paragon Plus Environment
Page 5 of 25
Environmental Science & Technology
69
The N2/Ar method was applied to estimate N2 fixation by measuring the surface water
70
concentration of N2. As noted above, this same method is used to estimate the N2 excess caused
71
by denitrification (e.g. 4). Given the similar physico-chemical properties of N2 and Ar, the N2/Ar
72
concentration ratios in seawater correspond within narrow limits to saturation with atmospheric
73
N2 and Ar (denoted as (N2/Ar)sat), at the respective temperature and salinity as long as no N2
74
transformations occur
75
concentration will be reflected in a shift of the N2/Ar ratio and can be determined by multiplying
76
the difference in the N2/Ar ratio by the Ar concentration 14, as shown in Eq. (1):
77
∆𝑁2 = ((𝐴𝑟2 ) − (𝐴𝑟2 )
78
where the element symbols represent the concentrations of those elements.
79
However, the use of (N2/Ar)sat as a non-biotic reference for the actual measured N2/Ar ratio is an
80
approximation and may be associated with considerable uncertainties if the N2 concentration is
81
expected to change by only a few percent, as in the case of N2 fixation. One reason for these
82
problems is due to the seasonality of the surface water temperature in combination with the
83
different temperature coefficients of the solubility constants for N2 and Ar. The solubility of N2
84
and Ar decreases by 1.99% and 2.16% per Kelvin, respectively.15 Consequently, (N2/Ar)sat values
85
run through a minimum/maximum corresponding to the seasonality of the sea surface
86
temperature (SST, Fig. 1). However, the seasonality of the real N2/Ar ratio lags behind that of
87
(N2/Ar)sat because gas exchange and re-equilibration with atmospheric N2 and Ar do not occur
88
spontaneously but are delayed with respect to the temperature change. Furthermore, Ar gas
89
exchange is faster than N2 gas exchange because Ar has a higher diffusivity (lower Schmidt
90
number).
𝑁
𝑁
𝑠𝑎𝑡
13
. Hence, any biogeochemically induced changes in the N2 (ΔN2)
(1)
) 𝐴𝑟
5
ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 25
91
To estimate possible deviations between the N2/Ar ratios of the “real” abiotic concentrations and
92
those of (N2/Ar)sat, N2/Ar ratios were simulated by a simple model that takes into account the
93
different scenarios in the Baltic Sea. A sinus-shaped SST seasonality (0–20°C) of the water
94
column with a defined mixed layer depth (zmix) was considered. The initial concentrations of N2
95
and Ar at the start of the calculations were set to saturation values. SST was then changed at time
96
sat steps of 1 day and the difference between the “new” ct+1 and the previous ct was used to calculate
97
the exchange with the atmosphere (F) of either N2 or Ar for each time step, as shown in Eqs. (2)
98
and (3):
99
𝐹 = 𝑘660 (660)
𝑆𝑐
−0.5
𝑠𝑎𝑡 ) (𝑐𝑡 − 𝑐𝑡+1
(2)
100
with:
101
𝑘660 = 0.24 u2
102
where k660 is the gas exchange transfer velocity at a Schmidt number of 660 [cm h-1]; Sc is the
103
Schmidt number which is reciprocally proportional to the diffusivity of the considered gas in
104
seawater; and u is the wind speed [m s-1].16,17
105
Dividing the flux (F) by the mixed layer depth yields the changes in N2 and Ar concentrations in
106
units of mmol m-3 from t to t+1.
107
For the determination of csat we first calculated the solubility constant as a function of
108
temperature and salinity15 which was then multiplied with the atmospheric partial pressures of the
109
respective gases. For the calculation of the latter, we used standard atmospheric pressure as well
110
as the N2 and Ar mole fractions in dry air and assumed water vapor saturation at the sea-air
111
interface.
(3)
6
ACS Paragon Plus Environment
Page 7 of 25
Environmental Science & Technology
112
N2/Ar seasonality was simulated for a scenario in which conditions favor mid-summer production
113
fueled by N2 fixation: low wind speed (u = 4 m s−1), a shallow surface mixed layer (zmix = 5 m),
114
and a temperature maximum by the end of July of 20°C.
115
Three seasonal cycles were simulated for a spin-up of the model calculations. Deviations from a
116
steady state were already negligible after the second cycle. The seasonality of (N2/Ar)sat (Fig. 1,
117
blue curve) showed minimum and maximum values of 36.9 and 38.1, respectively, coinciding
118
with the minimum and maximum surface water temperature. In contrast, both the amplitude and
119
the phase of the simulated “real” N2/Ar ((N2/Ar)real; Fig. 1, red curve) were slightly different. The
120
difference (N2/Ar)real – (N2/Ar)sat corresponds to a bias in the calculated ΔN2 (δ(ΔN2), Fig. 1,
121
black curve). The asymmetric annual cycle of δ(ΔN2) was due to the superposition of the two
122
effects: different temperature coefficients of the solubility constants and different transfer
123
velocities (Schmidt numbers) for N2 and Ar.
124 125
Figure 1. Simulation of the N2/Ar ratios for a sinus-shaped seasonality of temperature:
126
Spontaneous equilibration by gas exchange (red line) and delayed equilibration due to slow gas
127
exchange (green line). The difference between the N2/Ar ratios at equilibrium (N2/Ar, saturation)
128
and under simulated real conditions (N2/Ar, real) corresponds to a bias of the calculated ΔN2
129
(δ(ΔN2), black line) for the performed scenario (Tmax=20°C, zmix =5 m, u = 4 m s-1). The blue dots 7
ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 25
130
in the third period of the simulations indicate the start (15.6.) and termination (15.8.) of the time
131
period when the bloom of cyanobacteria (i.e., nitrogen fixation) are most common in the Baltic
132
Sea.
133
Positive δ(ΔN2), which cause an underestimate of ΔN2, were the largest during spring when the
134
temperature increase was the strongest. For the period from mid-June to mid-August when
135
cyanobacteria blooms are common in the central Baltic, δ(ΔN2) ranged between 0.2 and 1.1 µmol
136
L-1. The simulations presented in Fig. 1 for a whole year was based on a mid-summer scenario
137
regarding zmix and u. Nonetheless, the results for a typical winter scenario, i.e., higher wind speed
138
and deeper mixing, showed a similar pattern for δ(ΔN2), but with somewhat larger amplitudes.
139
Our exemplary and simplified simulations of the influence of seasonal surface water temperature
140
variations on the N2/Ar ratios provided a sensitivity test of the N2/Ar method. The results are not
141
intended for practical use; rather, they indicate the possibility of errors in the determination of N 2
142
concentration changes caused by biogeochemical processes such as N2 fixation and
143
denitrification. In reality, the wind speed and mixed layer depth are subject to short-term
144
fluctuations, and the seasonality of the surface water temperature is not a continuous function of
145
time, but frequently pulse-like. To account for such conditions, repeated measurements on the
146
basis of a few days may provide a realistic N2/Ar reference ratio for the start of the N2 fixation by
147
cyanobacteria. This could be facilitated by the deployment of an automated equi-MS system on a
148
VOS line (see section “Comparison with the pCO2 approach”).
149
Our simulations did not account for the impact of air bubble injection on the determination of
150
ΔN2. Previous studies have demonstrated that the injection of air bubbles may generate a
151
significant surface water N2 and Ar disequilibrium that, depending on the bubble size, may also
152
affect the N2/Ar ratio (e.g. 18, 19). Small bubbles (bubble diameter < 150 m) may collapse by the 8
ACS Paragon Plus Environment
Page 9 of 25
Environmental Science & Technology
153
complete dissolution of the entrapped air, thereby transferring the atmospheric N2/Ar signature
154
into the surface water. This process would favor the enrichment of the less-soluble gas species
155
and thus result in an increase of ΔN2 that counteracts the N2 loss by N2 fixation. The effect of the
156
injection of large air bubbles on the N2/Ar ratio is far more complex since these bubbles rise back
157
to the sea surface and the entrapped air is incompletely dissolved in the seawater. Modeling the
158
effect of bubbles on the N2/Ar ratio of seawater must therefore include several parameters (e.g.,
159
bubble size spectrum, penetration depth, rising speed, transfer velocity, surfactants), which was
160
beyond the scope of this study. However, to assess the possible effect of bubbles on our field
161
measurements and on N2 fixation studies in the Baltic Sea in general, the meteorological and
162
hydrographic conditions that allow for the massive occurrence of cyanobacteria in the Baltic Sea
163
must be taken into account. Studies have shown that cyanobacterial blooms occurring during
164
June–August are episodic events triggered by the low-wind conditions (about < 4 m/s), which
165
typically last for 1–3 weeks.20, 21 The shallow thermal stratification (5–10 m) that develops in the
166
calm waters increases the exposure of the cyanobacteria to solar radiation
167
conditions, the generation of bubbles can be reasonably excluded and a possible effect on N2
168
fixation studies neglected. Furthermore, the development of extreme temperature gradients at the
169
thermocline (up to 10°C over a depth interval of only 10 m) stabilizes the stratification, such that
170
mixing (entrainment) with deeper water masses that may have been subjected to denitrification is
171
widely prevented.
22
. Under these
172 173
Measurement Procedure
174
The concentrations of N2 and Ar in surface water were measured by coupling an equilibrator
175
system with a mass spectrometer (Equi-MS system), as shown in Fig. 2. 9
ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 25
176 177
Figure 2. Diagram of the Equi-MS system.
178
The approach is based on the equilibrium of atmospheric gases between the water and gas phases
179
within the headspace of the spray-type equilibrator (E1, total cell volume 2.2 L, water volume
180
500 mL). Seawater enters the equilibrator at the top, through two nozzles, at a flow rate of 5.5 L
181
min-1 and leaves the chamber through a siphon at the bottom. Temperature is measured using a
182
temperature probe (PT100, precision of 0.1°C) inserted into the cell. To achieve equilibrium
183
between the headspace and the seawater, gas (air) is circulating in a closed loop (gas volume in
184
the tubing 0.5 L) through the equilibrator at a flow rate of 600 ml min-1 (membrane pump, P1,
185
KNF Neuberger). A Peltier cooling trap (PK1, Gröger & Obst) reduces the water content in the
186
gas stream and the subsequent water guard protects the mass spectrometer from water leakage.
187
From this pre-dried gas stream a constant flow of 50 ml min-1 is directed into the sampling loop
188
of the mass spectrometer (adjusted with valve V1). For gas analyses, a three-way valve (V2)
189
opens and an external membrane pump (P2) transports the gas along the inlet of the mass
190
spectrometer (flow rate 15 ml min-1). The dead volume between the three-way valve and
191
membrane pump is flushed for 1 min before the gas analyses are conducted for an additional
192
minute, leading to a total removal of 30 ml of gas from the equilibrator gas cycle (< 2% of the
193
total gas volume in the Equi-MS system). The removal of the gas sample is balanced by pre-
194
equilibrated air from the headspace of an auxiliary bubble-type equilibrator (E2, total cell volume
10
ACS Paragon Plus Environment
Page 11 of 25
Environmental Science & Technology
23
195
0.35 L, water volume 0.23 L,
). The auxiliary equilibrator is open to the atmosphere and the
196
headspace gas in the chamber is circulated by an external pump (P3) through the water column.
197
The mole fractions of N2, Ar and H2O are determined using a commercially available quadrupole
198
mass spectrometer (QMS, GAM 200, InProcess Instruments). The gas flow from the gas inlet
199
into the vacuum system is reduced to approximately 1 cm3STP min-1 by a 3.2-m-long capillary
200
(inner diameter of 50 m). A high-vacuum of ~6×10-6 mbar is maintained by a mini-diaphragm
201
roughing pump (KNF Neuberger) and a HiPace 80 turbo pump (Pfeiffer Vacuum). The gas
202
species are ionized by electron impact ionization (ionization energy of 70 eV). The produced ions
203
are sorted by a quadrupole mass analyzer and detected by an electron-multiplier. N2, Ar, and H2O
204
are quantified from the signal intensities at mass-to-charge ratios (m/z) of 29, 40, and 18
205
respectively. In contrast to the detection of N2 at m/z 28, the use of m/z 29 shows less mass
206
interference with other gas species and avoids the large difference of signal intensity between m/z
207
28 and m/z 40. Within a measurement cycle of 6 s, each gas species is measured for 1 s, with the
208
remaining time used to define the baseline of the detector signal. This cycle is repeated ten times
209
to derive a solid database leading to a total run time for the gas analyses of 1 min. Before each
210
sequence of gas analyses, the mass spectrometer is calibrated against atmospheric air. An external
211
pump (P4) delivers the gas sample, which is dried by passing through a phosphorus pentoxide
212
(P4O10, Sicapent® with indicator, Merck) and silica gel (SiO2, silica gel with indicator, Roth)
213
column and directed by a multiposition valve (V3, VICI Valco Instruments) and the three-way
214
valve to the gas inlet of the mass spectrometer. The Quadstar 32-bit software (InProcess
215
Instruments) is used to operate the mass spectrometer and analyze the data.
216
The partial pressure of a gas i in the headspace of the equilibrator cell (piequi) is calculated from
217
the mole fraction of the gas component in dry air (xidry), the headspace pressure (ptotal) measured 11
ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 25
218
outside the equilibrator cell, and the partial pressure of water (pH2O), which is assumed to be at
219
saturation level for the given temperature and salinity.24 The relationship is expressed by Eq. (4).
220
𝑝𝑖𝑒𝑞𝑢𝑖 = 𝑥𝑖
221
The concentration of the dissolved gas ci (mol kg-1) is then calculated according to Henry’s law
222
using the Bunsen coefficient and the molar gas volume (Vmi), as shown in Eq. (5).
223
𝑐𝑖 =
224
Since the concentration changes of N2 in surface water attributable to N2 fixation are expected to
225
be small, an evaluation of the changes in the concentrations of N2 and Ar requires a high degree
226
of precision and accuracy. We therefore conducted a series of laboratory tests in which the Equi-
227
MS system was coupled to a circulation thermostat (Huber CC-K15 with Pilot ONE, Peter Huber
228
Kältemaschinenbau) filled with 15 L of tap water. To achieve stable equilibrium conditions
229
between the tap water and laboratory air, the water temperature in the thermostat was fixed at 20
230
±0.02°C and laboratory air was permanently introduced into the water reservoir through glass frit.
231
From this reservoir, water was continuously circulated through the equilibrator cell by an external
232
pump with a flow rate of 5.5 L min-1. To determine N2 concentration changes the Equi-MS
233
measurement cycles were conducted in the following order: (i) calibration of the MS with outside
234
air, as described above; (ii) measurement of the mole fractions of N2, Ar, and H2O in laboratory
235
air; and (iii) measurement of these gases in the headspace of the equilibrator cell. The cycle was
236
repeated every 60 min and a record of seven consecutive cycles was used to define the standard
237
deviation (SD) and detection limits for N2 and Ar gas measurements in the equilibrator cell
238
(Table 1). The values of these indicators were then applied to define the SDs of the N2/Ar ratios
239
and N2. In addition, the measured concentrations of N2 and Ar in tap water (Cmeas) were
𝑑𝑟𝑦
(4)
∙ (𝑝𝑡𝑜𝑡𝑎𝑙 − 𝑝𝐻2𝑂 )
𝛽𝑖 ∙𝑝𝑖
(5)
𝑉𝑚𝑖
12
ACS Paragon Plus Environment
Page 13 of 25
Environmental Science & Technology
240
compared with the saturation concentrations of these gases (csat, calculated after 15), as shown in
241
Table 1.
242
Table 1. Performance indicators of the Equi-MS system obtained by repeated measurements of
243
the gases in ambient air and in the headspace of the equilibrator cell (absolute standard deviation,
244
aSD). The detection limit is defined as twice the aSD. Gas
aSD
Detection
aSD
limit
29
(N2)
Detection
aSD
limit
N2/Ar
(ppm)
(ppm)
(mol L-1)
(mol L-1)
182
364
0.13
0.26
(mol L-1)
0.05 40
Ar
12
24
0.02
aSD N2
cmeas
csat
tap water
tap water
(mol L-1)
(mol L-1)
540.1
549.1
13.9
14.1
0.7
0.04
245 246
Another important feature of the Equi-MS system is the equilibration time (τ), which is the time
247
required for the establishment of a “new” equilibrium in response to changing water masses.
248
Using our system onboard a moving ship, τ together with the ship’s speed determines the
249
temporal resolution of the measurements and may be interpreted as the “footprint” of the
250
measured change in the N2/Ar ratio. To calculate τ, we performed “step” experiments, in which
251
the Equi-MS system was successively connected to two different water sources. The system was
252
first supplied with tap water previously equilibrated with ambient air in a temperature-controlled
253
bath (T = 20 ± 0.02°C). After a stable N2/Ar signal was obtained from the MS, the water supply
254
was switched to the original tap water, which was not at equilibrium with the ambient air at T =
255
14°C and had a lower equilibrium N2/Ar ratio, (N2/Ar)eq. Due to gas exchange, a progressive
256
decrease in the measured headspace N2/Ar ratio was registered, with (N2/Ar)eq reached after
13
ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 25
257
about one hour (Fig. 3, inset). The temporal development of (N2/Ar)t in the headspace was
258
expressed by a simple exponential function:
259
(N2/Ar)eq – (N2/Ar)t = [(N2/Ar)eq – (N2/Ar)t=0] ∙ exp – t/
260
Eq. (6) implies that the gas exchange of N2 and Ar occurs with identical equilibration times. This
261
is not exactly the case because the exchange rate of Ar is slightly faster than that of N2 (lower
262
Schmidt number); nonetheless, the effect on the calculation of τ for the ratio N2/Ar is slight. For
263
the determination of τ, we plotted ln[(N2/Ar)eq-(N2/Ar)t] vs. time (t, Fig. 3). According to Eq. (6),
264
the reciprocal slope of the respective regression line gives the e-fold equilibration time (τ). A
265
value of τ = 12 min was obtained, which means that it takes 12 min until 63 % of an initial
266
(arbitrary) disequilibrium is compensated by gas exchange.
(6)
267 268
Figure 3. The N2/Ar ratio as a function of time (inset) and the logarithmic presentation of the
269
difference between N2/Ar at any time t and at equilibrium as a function of time. The reciprocal
270
slope of the regression line corresponds to the e-fold equilibration time ().
271 272
Field measurements
14
ACS Paragon Plus Environment
Page 15 of 25
Environmental Science & Technology
273
To demonstrate that the Equi-MS system is capable of detecting N2 loss due to N2 fixation, we
274
participated in Meteor cruise M117, conducted in order to study a cyanobacterial bloom in the
275
Baltic Sea.25 Surface water was permanently pumped from 3 m below the sea surface to the Equi-
276
MS system at a flow rate of 5.5 L min−1. Sea surface temperature, salinity and meteorological
277
data (e.g., wind speed) were recorded using a thermosalinograph (SBE 21 SeaCAT, Sea-Bird
278
Scientific) and the ship’s weather station.
279
The experimental set-up was the same as that used in our laboratory-based measurements. N2/Ar
280
ratios in the equilibrator’s headspace were measured at hourly intervals. At an e-fold
281
equilibration time of about 12 min, this time difference guaranteed that the N2/Ar ratio in the
282
headspace was at equilibrium with that of the gasses dissolved in seawater. The equilibration time
283
τ also determines the spatial resolution (footprint) of the N2 concentration measurements. A value
284
of three times the equilibration time (95% equilibrium) and a cruising speed of 10 knots yielded a
285
spatial resolution of about 5 nautical miles, which is sufficient for large-scale records of N2
286
fixation. However, in highly dynamic surface water regimes, the presence of strong temperature
287
variations may complicate the choice of an appropriate temperature for the calculation of the
288
reference N2/Ar ratio. This problem can be mitigated by reducing the equilibration time, which is
289
easily done by reducing the volume of the equilibrator’s headspace/tubing, a value directly
290
proportional to τ. 26
291
Another problem related to the circumstances of our cruise were measurements made while the
292
ship remained at a station to conduct other research activities. Due to the deployment of samplers
293
and various instruments as well as the frequent re-positioning of the ship, the thermal
294
stratification of the upper surface water was disrupted. This may have caused erroneous
295
measurements because cyanobacterial blooms develop preferably in the shallow surface layer that 15
ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 25
296
forms under the calm conditions typical of mid-summer. To eliminate potential artifacts from our
297
dataset, we rejected data obtained at a ship’s speed of < 3 knots. Nevertheless, if sampling-related
298
perturbations are avoided, then N2/Ar data acquisition at a ship speed < 3 knots is possible and is
299
suggested for investigations in areas subject to complex oceanographic conditions (e.g., filaments
300
in upwelling regions).
301 302
RESULTS AND DISCUSSION
303
Surface water N2 deficits detected using the N2/Ar approach
304
Our field investigations focused on a time slot of the research cruise that was characterized by a
305
general decrease in wind speed and an increase in SST, which occurred between July 27 and
306
August 12 (Fig. 4a). The cruise can be subdivided into three main parts in terms of its
307
oceanographic/meteorological features and the impact of those features on the surface water N2
308
pattern.
309 16
ACS Paragon Plus Environment
Page 17 of 25
Environmental Science & Technology
310
Figure 4. (a) Color-coded sea surface temperature (SST), wind speed, and (b) N2 values
311
measured along a ship track in the central Baltic Sea in summer 2015. In (b) data points in the
312
eastern Gotland Sea that belong to Part 3 are framed by black lines to better distinguish these data
313
from the overlapping data points belonging to Part 1. The black line in (b) indicates the VOS
314
Finnmaid track used to calculate N2 values from surface water measurements of the CO2 partial
315
pressure.
316
Part 1. Between July 27 and August 4, relatively strong wind speeds of up to 13 m s -1 were
317
detected and N2 values of around zero reflected atmospheric equilibrium (Fig. 4b). However, an
318
oversaturation of N2 in the surface water at some measuring points was indicated by the positive
319
N2 values. We assume that the increase in N2 values observed in Part 1 may have been due to
320
wind-induced breaking waves and the concomitant entrainment of atmospheric bubbles and their
321
dissolution in the surface water.13,
322
atmospheric gas composition (N2/Ar = 83), which was quite different from the surface water ratio
323
of N2/Ar in equilibrium with the atmosphere.
324
Part 2. Around August 3, satellite-based SST measurements showed relatively cold temperatures
325
along the eastern coast of Sweden (Fig. 5a). These could be attributed to the upwelling of cold
326
deep-water masses and they were well-reflected in the surface water temperatures recorded
327
during the research cruise between the southern tip of Gotland and east of Öland (drop in the SST
328
from 15.8°C to 11.4°C, Fig. 4a). The temperature pattern correlated with a positive shift in N2
329
(up to 5.3 mol kg-1 near Öland, Fig. 4b) and suggested that denitrification in suboxic and anoxic
330
zones near the sediment28 contributed to the N2 content in the surface water, due to an upward
331
transport of bottom water to the sea surface. The upwelling of deep cold water and the subsequent
27
These air bubbles carried the N2/Ar signature of the
17
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 25
332
warming at the sea surface would further support the positive shift in N2 due to a mechanism
333
comparable with the warming of surface water during spring (Fig. 1).
334 335
Figure 5. Color-coded SSTs from (a) August 3 and (b) August 7 (data from AVHRR Pathfinder
336
Version 5.2 obtained from the US National Oceanographic Data Center and the Group for High-
337
Resolution Sea Surface Temperature and reanalyzed by the Integrated Climate Data Center). 29 (c)
338
True-color image from August 7 indicating the accumulation of cyanobacteria at the sea surface
339
(data from MODIS Aqua, NASA Goddard Space Flight Center, Ocean Ecology Laboratory,
340
Ocean Biology Processing Group).
341
Part 3. In the following part of the cruise (August 5–12), the continuous decline of the wind
342
speed and the associated breakdown of the upwelling cells were reflected in the increase in SST
343
(Figs. 4a and 5b). In addition, the N2 values declined (Fig. 4b) and correlated negatively with
344
the SST (Fig. 6). This clearly indicated persistent N2 fixation by cyanobacteria. Schneider et al.12
345
showed that temperature increases correlate strongly with N2 fixation activity and concluded that
346
the exposure of cyanobacteria to solar radiation is the most important factor regulating the
347
intensity of N2 fixation in the Baltic Sea. This conclusion was supported by the true-color satellite
348
images obtained for this study, which indicated that the calm weather conditions during the cruise
349
reduced the turbulence in the mixed layer and favored the accumulation of cyanobacteria at the 18
ACS Paragon Plus Environment
Page 19 of 25
Environmental Science & Technology
350
sea surface (Fig. 5c). The images also showed that cyanobacteria were not homogenously spread
351
over the sea surface but had accumulated in patches.
352 353
Figure 6. Relationship between SST and N2 for the time span July 30 to August 13.
354 355
Comparison with the pCO2 approach
356
A confirmation of the surface water N2 deficits detected using the N2/Ar method was obtained
357
from automated measurements of the surface water pCO2 made from the VOS Finnmaid.12
358
Between July 30 and August 12, the ship crossed the eastern Gotland Sea and the entrance to the
359
Gulf of Finland seven times. This period of time, corresponding to Part III of our cruise, was
360
characterized by high N2 values in the surface water. From the high-resolution time series data
361
for the mean pCO2 (Surface Ocean CO2 Atlas, SOCAT v4 data, http://www.socat.info/), we used
362
the data from the northern Gotland Sea (Fig. 4b) to characterize the temporal development of N2
363
fixation in the northern part of the central Baltic Sea. However, since our N2/Ar measurements
364
were performed at different times and at different places, a direct comparison of the two datasets
365
is only conditionally possible.
19
ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 25
366
At the beginning of August, the total CO2 concentration (CT), derived from the pCO2
367
measurements
368
investigations (Fig. 7, inset). Since the DIN pool (Swedish National Monitoring Program,
369
https://sharkweb.smhi.se) was entirely exhausted, the drop in CT (~40 µmol L-1) was attributed to
370
biomass production fueled by N2 fixation. Dividing the organic carbon equivalent by the Redfield
371
C/N ratio (6.625
372
(blue dots/line) as ΔN2 together with the range of variability that corresponded to the spatial
373
variability of CT within the considered area in the northern Gotland Sea.
30
, started to decrease and reached a minimum by August 13, at the end of our
31
) yielded the amount of “fixed” nitrogen.12 The latter is presented in Fig. 6
374 375
Figure 7. ΔN2 values for the individual measurements made between July 30 and August 13
376
obtained from N2/Ar measurements (red dots) and as inferred from continuous pCO2 records
377
obtained from a cargo ship (blue dots/line). Open blue circles indicate the range of the spatial
378
variability. The shadowed green area in the inset displays the mean total CO2 decrease (obtained
379
from the pCO2 data) during the period of N2 fixation in the northern Gotland Sea (CT = blue line,
380
SST = red line).
381
The uncertainties and problems associated with the choice of the reference N2/Ar ratio for the
382
determination of ΔN2 were discussed above. Here we applied the (N2/Ar)sat value from the
383
beginning of N2 fixation activity, on August 4, for the following fixation period (Part III). The 20
ACS Paragon Plus Environment
Page 21 of 25
Environmental Science & Technology
384
constant (N2/Ar)sat implied that the gas exchange of N2 and Ar was negligible, an assumption
385
justified by the strong reduction in gas exchange related to the low wind speed. The general trend
386
in the mean CO2-derived ΔN2 was widely reproduced by the temporal change in ΔN2 determined
387
from our N2/Ar method (Fig. 7, red dots). However, this method resulted in a large scatter of the
388
ΔN2 values, perhaps due to the patchiness of the cyanobacterial distribution (Fig. 5) and thus of
389
their N2 fixation activity. In this regard, it must be taken into account that the N 2/Ar
390
measurements were made while the ship was cruising at a speed of about 10 knots. Thus,
391
measurements performed in the course of a single day may refer to water masses separated from
392
each other by 100 miles or even more. Hence, the variability of ΔN2 reflects not only small scale
393
patchiness but may also be caused by regional gradients of the N2 fixation activity.
394 395
Assessment
396
The results of laboratory tests and a pilot study in the Baltic Sea showed that our N2/Ar approach
397
enables the detection of relatively small decreases in N2 concentrations, attributable to N2
398
fixation. Observations of enhanced N2 concentrations could be explained by upwelling events and
399
bubble entrapment, conditions at which the occurrence of cyanobacteria is unlikely. We consider
400
our first research campaign as a feasibility study for “underway” measurements of N2 fixation on
401
a research vessel. This approach will facilitate large-scale, direct detections of N2 fixation in the
402
Baltic Sea, which is believed to be of the same order of magnitude as the sum of riverine input
403
and the atmospheric deposition of DIN.32,
404
account for the regional variability and episodic character of N2 fixation in the Baltic Sea cannot
405
be reasonably made from sample-based studies such as
33
Reliable measurements based on estimates that
15
N incubations and calculations of the
21
ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 25
406
total N budget.32 We are aware of the uncertainties and limitations of our N2/Ar approach, which
407
are mainly associated with the choice of an appropriate reference N2/Ar ratio.
408
The development of an automated version of our MS-Equi system and its deployment on a VOS,
409
such as for the above-mentioned pCO2 measurements, will provide further information on the
410
temporal and spatial distribution of N2 fixation. An automated approach may also circumvent
411
some of the problems discussed above, because the generation of N2/Ar ratios as a function of
412
time includes the determination of an appropriate reference N2/Ar ratio. In addition, the
413
applicability of the N2/Ar approach should be validated through tests in different oceanographic
414
settings with similar biogeochemical characteristics. Like the Baltic Sea, other coastal regions are
415
progressively affected by warming, eutrophication and algal blooms. We are confident that, also
416
in those regions, the N2/Ar approach will support the ecological management of these systems.
417
418
AUTHOR INFORMATION
419
Corresponding Author
420
Oliver Schmale (oliver.schmale@io-warnemuende)
421
Author Contributions
422
All of the authors contributed to the writing of the manuscript and gave their approval to the final
423
version of the manuscript.
424
Funding Sources
425
This present work was supported by the Federal Ministry of Education and Research (BMBF)
426
through MOBALAB: Mobiles Analyselabor "Veränderungen der Ostsee". 22
ACS Paragon Plus Environment
Page 23 of 25
Environmental Science & Technology
427 428
ACKNOWLEDGMENTS
429
We thank Bernd Sadkowiak for the construction and set up of the equilibrator line, and Stefan
430
Otto for supporting the mass spectrometry work. We also thank Lars Umlauf, Herbert Siegel and
431
Monika Gerth for processing the satellite data, and the captain and crew of the RV Meteor M117
432
for technical support. We acknowledge the use of Rapid Response imagery from the Land,
433
Atmosphere Near real-time Capability for EOS (LANCE) system, operated by NASA's Earth
434
Science Data and Information System (ESDIS) and funded by NASA Headquarters. We thank
435
GHRSST and the US National Oceanographic Data Center for providing sea surface temperature
436
data, obtained in part with support by a grant from the NOAA Climate Data Record (CDR)
437
program for satellites. The work described herein was supported by the Federal Ministry of
438
Education and Research (BMBF) through MOBALAB: Mobiles Analyselabor "Veränderungen
439
der Ostsee".
440 441
REFERENCES
442 443 444 445 446 447 448 449 450 451 452 453 454 455 456
1. Kaiser, J.; Reuer, M. K.; Barnett, B.; Bender, M. L., Marine productivity estimates from continuous O2/Ar ratio measurements by membrane inlet mass spectrometry. Geophys. Res. Lett. 2005, 32, (19), L19605. 2. Tortell, P. D.; Bittig, H. C.; Körtzinger, A.; Jones, E. M.; Hoppema, M., Biological and physical controls on N2, O2, and CO2 distributions in contrasting Southern Ocean surface waters. Glob. Biogeochem. Cycles 2015, 29, (7), 994-1013. 3. Capone, D. G.; Bronk, D. A.; Mulholland, M. R.; Carpenter, E. J., Nitrogen in the marine environment. Elsevier: Amsterdam, 2008; p 1757. 4. Codispoti, L. A.; Brandes, J. A.; Christensen, J. P.; Devol, A. H.; Naqvi, S. W. A.; Paerl, H. W.; Yoshinari, T., The oceanic fixed nitrogen and nitrous oxide budgets : Moving targets as we enter the anthropocene? Scientia Marina 2001, 65, (2), 85-105. 5. Devol, A. H.; Uhlenhopp, A. G.; Naqvi, S. W. A.; Brandes, J. A.; Jayakumar, D. A.; Naik, H.; Gaurin, S.; Codispoti, L. A.; Yoshinari, T., Denitrification rates and excess nitrogen gas concentrations in the Arabian Sea oxygen deficient zone. Deep Sea Research Part I: Oceanographic Research Papers 2006, 53, (9), 1533-1547. 23
ACS Paragon Plus Environment
Environmental Science & Technology
457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503
Page 24 of 25
6. Wasmund, W.; Voss, M.; Lochte, K., Evidence of nitrogen fixation by non-heterocystous cyanobacteria in the Baltic Sea and re-calculation of a budget of nitrogen fixation. Mar. Ecol.Prog. Ser. 2001, 214, 1-14. 7. O’Neil, J. M.; Davis, T. W.; Burford, M. A.; Gobler, C. J., The rise of harmful cyanobacteria blooms: The potential roles of eutrophication and climate change. Harmful Algae 2012, 14, 313-334. 8. Montoya, J. P.; Voss, M.; Kahler, P.; Capone, D. G., A Simple, High-Precision, HighSensitivity Tracer Assay for N(inf2) Fixation. Appl. Environ. Microbiol. 1996, 62, (3), 986-993. 9. Eggert, A.; Schneider, B., A nitrogen source in spring in the surface mixed-layer of the Baltic Sea: Evidence from total nitrogen and total phosphorus data. Journal of Marine Systems 2015, 148, 39-47. 10. Larsson, U.; Hajdu, S.; Walve, J.; Elmgren, R., Baltic Sea nitrogen fixation estimated from the summer increase in upper mixed layer total nitrogen. Limnol. Oceanogr. 2001, 46, (4), 811-820. 11. Schneider, B.; Kaitala, S.; Raateoja, M.; Sadkowiak, B., A nitrogen fixation estimate for the Baltic Sea based on continuous pCO2 measurements on a cargo ship and total nitrogen data. Continental Shelf Research 2009, 29, (11), 1535-1540. 12. Schneider, B.; Gustafsson, E.; Sadkowiak, B., Control of the mid-summer net community production and nitrogen fixation in the central Baltic Sea: An approach based on pCO2 measurements on a cargo ship. Journal of Marine Systems 2014, 136, 1-9. 13. Hamme, R. C.; Emerson, S. R., Mechanisms controlling the global oceanic distribution of the inert gases argon, nitrogen and neon. Geophys. Res. Lett. 2002, 29, (23), 35-1-35-4. 14. Benson, B. B.; Parker, P. D. M., Nitrogen/argon and nitrogen isotope ratios in aerobic sea water. Deep Sea Research (1953) 1961, 7, (4), 237-253. 15. Hamme, R. C.; Emerson, S. R., The solubility of neon, nitrogen and argon in distilled water and seawater. Deep Sea Research Part I: Oceanographic Research Papers 2004, 51, (11), 1517-1528. 16. Wanninkhof, R., Relationship between wind speed and gas exchange over the ocean. J. Geophys. Res. 1992, 97, (C5), 7373-7382. 17. Wanninkhof, R.; Asher, W.; Ho, D. T.; Sweeney, C.; McGillis, W. R., Advances in quantifying air-sea gas exchange and environmental forcing. Annual Reviews Marine Science 2009, 1, 213-244. 18. Nicholson, D. P.; Emerson, S. R.; Khatiwala, S.; Hamme, R. C., An inverse approach to estimate bubble-mediated air-sea gas flux from inert gas measurements. Proceedings on the 6th International Symposium on Gas Transfer at Water Surfaces 2011, 223-237. 19. Tortell, P. D.; Bittig, H. C.; Körtzinger, A.; Jones, E. M.; Hoppema ,M., Biological and physical controls on N2, O2, and CO2 distributions in contrasting Southern Ocean surface waters. Global Biogeochem. Cycles 2015, 29, 994–1013. 20. Wasmund, N., Occurrence of cyanobacterial blooms in the Baltic Sea in relation to environmental conditions. Int. Revue ges. Hydrobiol. 1997, 82, 169-184. 21. Lips, I. and Lips, U., Abiotic factors influencing cyanobacterial bloom development in the Gulf of Finland (Baltic Sea). Hydrobiologia 2008, 614, 133-140. 22. Kahru, M.; Savchuk, O. P.; Elmgren, R., Satellite measurements of cyanobacterial bloom frequency in the Baltic Sea: interannual and spatial variability. Mar. Ecol. Prog. Ser. 2007, 343, 15-23. 23. Gülzow, W.; Rehder, G.; Schneider, B.; Schneider von Deimling, J.; Sadkowiak, B., A new method for continuous measurement of methane and carbon dioxide in surface waters using 24
ACS Paragon Plus Environment
Page 25 of 25
504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533
Environmental Science & Technology
off-axis integrated cavity output spectroscopy (ICOS): an example from the Baltic Sea. Limnology and Oceanography-Methods 2011, 9, 176-184. 24. Ambrose, D.; Lawrenson, I. J., The vapour pressure of water. The Journal of Chemical Thermodynamics 1972, 4, (5), 755-761. 25. Wurl, O.; Nausch, G.; Nausch, M.; van Pinxteren, M.; Kuss, J.; Loick-Wilde, N., Biochemical processes in upwelling zones of the Baltic Sea - Cruise No. M117 - July 23 - August 17, 2015 - Hamburg (Germany) - Rostock (Germany); Bremen, 2016; p 47. 26. Schneider, B.; Sadkowiak, B.; Wachholz, F., A new method for continuous measurements of O2 in surface water in combination with pCO2 measurements: Implications for gas phase equilibration. Mar. Chem. 2007, 103, 163-171. 27. Schudlich, R.; Emerson, S., Gas supersaturation in the surface ocean: The roles of heat flux, gas exchange, and bubbles. Deep Sea Research Part II: Topical Studies in Oceanography 1996, 43, (2), 569-589. 28. Asmala, E.; Carstensen, J.; Conley, D. J.; Slomp, C. P.; Stadmark, J.; Voss, M., Efficiency of the coastal filter: Nitrogen and phosphorus removal in the Baltic Sea. Limnol. Oceanogr. 2017, 62, (S1), S222-S238. 29. Casey, K. S.; Brandon, T. B.; Cornillon, P.; Evans, R., The Past, Present, and Future of the AVHRR Pathfinder SST Program. In Oceanography from Space: Revisited, Barale, V.; Gower, J. F. R.; Alberotanza, L., Eds. Springer Netherlands: Dordrecht, 2010; pp 273-287. 30. Schneider, B.; Müller, J. D., Biogeochemical transformations in the Baltic Sea Obervations through carbon dioxide glasses. Springer Oceanography 2018, ISSN 2365-7677. 31. Redfield, A. C.; Ketchum, B. H.; Richards, F. A., The influence of organisms on the composition of sea-water. In The sea: ideas and observations on progress in the study of the seas, Hill, M. N., Ed. Interscience: New York, 1963; pp 26-77. 32. Wasmund, N.; Nausch, G.; Schneider, B.; Nagel, K.; Voss, M., Comparison of nitrogen fixation rates determined with different methods: a study in the Baltic Proper. Mar. Ecol. Prog. Ser. 2005, 297, 23-31. 33. Rolff, C.; Almesjö, L.; Elmgren, R., Nitrogen fixation and abundance of the diazotrophic cyanobacterium Aphanizomenon sp. in the Baltic Proper. Mar. Ecol. Progr. Ser. 2007, 332, 107118.
25
ACS Paragon Plus Environment