You are on page 1of 13

21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023].

See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, G01020, doi:10.1029/2011JG001849, 2012

Carbon and oxygen cycles: Sensitivity to changes in environmental


forcing in a coastal upwelling system
L. Bianucci1,2 and K. L. Denman1,3,4
Received 31 August 2011; revised 6 December 2011; accepted 10 December 2011; published 22 February 2012.
[1] Biogeochemical cycles in the coastal ocean are changing and will continue to change
in response to a changing climate. Effects on the oxygen and carbon cycles are particularly
important, as either episodic or permanent shifts toward lower oxygen and/or higher
inorganic carbon conditions can impact coastal ecosystems negatively. Here we study the
sensitivity of these cycles to changes that may occur in the coastal ocean, focusing on a
summer wind-driven upwelling region off southern Vancouver Island shelf. We use a quasi
2-D configuration of the Regional Ocean Modeling System (ROMS) to perform six
sensitivity experiments. Results indicate that carbon and oxygen cycles in this region may
be significantly affected by an altered upwelling season, a shallower offshore Oxygen
Minimum Zone, and a carbon-enriched environment. Combinations of these scenarios
suggest a potentially increasing risk for the development of coastal hypoxia and corrosive
conditions in the region.
Citation: Bianucci, L., and K. L. Denman (2012), Carbon and oxygen cycles: Sensitivity to changes in environmental forcing in
a coastal upwelling system, J. Geophys. Res., 117, G01020, doi:10.1029/2011JG001849.

1. Introduction [3] Oxygen Minimum Zones (OMZ) are permanently


hypoxic regions of the open ocean (O2 < 60 mmol m3),
[2] Future climate change will affect every aspect of the
typically along the continental margins of the eastern Pacific,
ocean, from its physics to its chemistry and biology. In
eastern Atlantic, and Indian oceans [Helly and Levin, 2004;
particular, biogeochemical cycles will not only be altered
Stramma et al., 2008]. Time series of O2 show a persistent
directly by the rising partial pressure of carbon dioxide
decline of concentrations in different OMZs as well as a
(pCO2) in the atmosphere, but also by indirect effects, such
shoaling of their upper boundaries [Whitney et al., 2007;
as higher temperatures and changes in wind strength and
Stramma et al., 2008]. These changes could impact shallower
patterns [e.g., Bakun, 1990; Falkowski et al., 2000]. For
waters in regions where OMZs affect outer continental
instance, the increase in sea surface temperatures and the shelves and upper slopes [Helly and Levin, 2004], especially
strengthening of stratification as global climate warms is
where wind-driven upwelling brings deep waters closer to or
expected to be sufficient to intensify existing hypoxia and
onto the shelves (e.g., western North America). The shoaling
generate hypoxia in new areas [Rabalais et al., 2010]. In
of the OMZ in the northeast Pacific [Whitney et al., 2007]
addition, if the biotic carbon to nitrogen ratio (C:N) increa-
may be a cause of recently observed hypoxic events in the
ses under elevated CO2 conditions (as observed in a meso-
coastal waters off Oregon and California [Grantham et al.,
cosm experiment [Riebesell et al., 2007]), anthropogenic
2004; Bograd et al., 2008; Chan et al., 2008], which in
CO2 emissions may extend tropical oxygen (O2) deficient
some cases turned the shelves into “dead zones” due to the
zones [Oschlies et al., 2008]. Models project a decline in O2
widespread mortality of benthic animals.
concentrations in the oceans during the 21st century
[4] Cycling of carbon in the ocean is also expected to
[Frölicher et al., 2009] and over the next 100,000 years change as pCO2 increases in the atmosphere and climate
[Shaffer et al., 2009]. Recent warming in the 1990s has
changes. A decline in pH and an increase in surface pCO2
already produced an estimated global oceanic O2 outgassing
have already been observed at several oceanic time series
of 0.3  0.4 1014 mol-O2 yr1 [Keeling and Garcia, 2002]. sites: stations ALOHA in the central North Pacific [Dore
et al., 2009] and BATS, off Bermuda in the North Atlan-
1 tic (both time series longer than 20 years [Bates, 2007]), as
School of Earth and Ocean Sciences, University of Victoria, Victoria,
British Columbia, Canada. well as station ESTOC (100 km north off Gran Canaria Island
2
Now at Department of Oceanography, Dalhousie University, Halifax, in the North Atlantic [Santana-Casiano et al., 2007]), which
Nova Scotia, Canada.
3
has a 10 yearlong record. These stations show that surface
Canadian Centre for Climate Modeling and Analysis, Environment ocean pCO2 has increased at rates indistinguishable from the
Canada, Victoria, British Columbia, Canada.
4
Now at VENUS Project, University of Victoria, Victoria, British
atmospheric increase (1.5 to 1.9 matm yr1 [Bindoff et al.,
Columbia, Canada. 2007]). On the west coast of the US, model simulations indi-
cate a pH decrease of 0.1 since pre-industrial times [Hauri
Copyright 2012 by the American Geophysical Union. et al., 2009]. Moreover, corrosive waters have been observed
0148-0227/12/2011JG001849

G01020 1 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

represents summer upwelling over the southern Vancouver


Island shelf. Six experiments (details in section 2) investi-
gate the model sensitivity to changes in strength of upwell-
ing, depth of the OMZ, and inorganic carbon concentration.
We introduce to our analysis the concept of the Respiration
Index [Brewer and Peltzer, 2009a, section 3], which simul-
taneously considers the effects of changes in both O2 and
pCO2 on marine organisms. This index is calculated in the
different simulations and compared with the traditional
definitions of hypoxia.

2. Design of Sensitivity Experiments


[6] We have developed a quasi 2-D application of the
Regional Ocean Modeling System (ROMS, www.myroms.
org [Haidvogel et al., 2008]) to model upwelling on the
Vancouver Island shelf in response to time variable wind-
forcing. The domain is a transect perpendicular to the iso-
baths (i.e., cross-shore distance versus depth) at 49°N
(Figure 1) and the small alongshore dimension (5 km) has
uniform conditions. A summary of the model configuration
Figure 1. Map of western North America showing the is provided in Table 1. This quasi 2-D approach precludes
location of the southern Vancouver Island shelf. The inset the inclusion of an alongshore pressure gradient that varies
shows the transect that the quasi 2-D model represents with cross-shelf distance and depth, preventing the modeling
(black line) and the meteorological buoy 46202 that pro- of a dynamic VICC [Masson and Cummins, 1999]. Hence,
vided winds to force the model (triangle). NCEP data used to model the VICC, water properties in the nearshore 6 km
to force surface net heat and shortwave fluxes are represen- are restored toward a fixed vertical profile representative of
tative of a region of 1.9° latitude  2.4° longitude centered the core of the VICC. The restoring is strongest at the
on the solid black circle. inshore boundary and decreases as exp(x2) toward off-
shore, where x is the distance from the inshore boundary (for
on shelves off western North America [Feely et al., 2008]. more details, see Bianucci et al. [2011, Appendix B]. The
There, the undersaturation horizon of aragonite (the less model configuration and evaluation, as well as the bio-
stable form of calcium carbonate, CaCO3, found in corals for logical and sediment modules coupled to the physical
example) was observed at depths between 40 and 120 m, model, are described in detail elsewhere [Bianucci, 2010;
even reaching the surface along one transect off northern Bianucci et al., 2011]. Vertical profiles represented cor-
California. rectly the observed vertical structure of temperature, nitrate,
[5] In contrast to other coastal regions of western North and other variables. Moreover, the statistical properties of the
America, wide spread hypoxic events and corrosive waters model and observations were in agreement. Modeled time
have not yet been observed along the west coast of Van- series of sea surface height captured some of the surface
couver Island. This region represents the northern limit of dynamics observed at a meteorological buoy in the region.
the California Current System (CCS) in Pacific Canada. The The model also represented the observed cross-shore gradi-
outflow from the Juan de Fuca Strait, relatively fresh and ent of DIC [Ianson et al., 2003], such that VICC waters are
nutrient-rich due to intense tidal mixing in the Strait, gen- enriched in DIC with respect to shelf waters at comparable
erates a buoyancy-driven coastal current, known as the depths.
Vancouver Island Coastal Current (VICC) [Freeland et al., [7] Coastal upwelling ecosystems are intrinsically 3-D
1984; Thomson et al., 1989; Crawford and Dewey, 1989]. systems, where mesoscale phenomena play an important
This O2-rich current and the relatively wide shelf protect the role [e.g., Gruber et al., 2006; Lathuilière et al., 2010].
shallower waters from developing hypoxia and aragonite However, a 2-D model is able to represent locally forced
undersaturation [Bianucci et al., 2011]. Moreover, wind- upwelling and facilitates extensive sensitivity analysis.
driven upwelling only occurs during summer [Strub et al.,
1987a, 1987b], such that the influence of carbon-rich and Table 1. Summary of Model Configuration
O2-depleted deep waters from offshore is limited to that
season (winds are downwelling-favorable during winter). Configuration
However, it is uncertain if hypoxia and undersaturation may Model dimensions Cross-shore (Lx) = 185 km.
develop in the future, given potential alterations to envi- Alongshore (Ly) = 5 km.
ronmental forcing with climate change. Projections for the Vertical (H) = from 40 to 1500 m.
Grid resolution Cross-shore (dx) = 953 m
21st century from an earth system model have predicted a Alongshore (dy) = 1.67 km.
decrease in O2 concentrations and increase in ocean acidi- Vertical (dz) = from 1.3 to 72.8 m
fication in the entire CCS [Rykaczewski and Dunne, 2010]. (30 s-layers).
Therefore, we use a coastal circulation model to investigate Time step dt = 320 sec.
Spin-up time 50 days.
the sensitivity of the carbon and O2 cycles to different for- Total length of experiments 125 days.
cings in our region of interest. A quasi 2-D configuration

2 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Table 2. Description of Model Sensitivity Experiments


Experiment Description and Characteristic Valuesa
1 Control experiment with 1993 late spring and summer forcing from 27 May 1993 (day 0) to 29 September 1993 (day 125).
2 Like 1, but with 1993 wind stress during spin-up (day 0 to 50) and 2002 wind stress (increased upwelling) after day 50 (16 July).
Mean t y = 0.028 N m2 (0.018 N m2)
3 Like 1, but with 2002 wind stress from day 0.
Mean t y = 0.023 N m2 (0.018 N m2)
4 Like 1, but initial conditions for O2 with a shallower OMZ and lower O2 concentrations.
Min. initial O2 at 800 m = 1.3 mmol m3 (10.6 mmol m3)
OMZ hypoxic threshold depth = 280 m (380 m)
5 Like 4, but with 2002 wind stress from day 0.
Min. initial O2 at 800 m = 1.3 mmol m3 (10.6 mmol m3)
OMZ hypoxic threshold depth = 280 m (380 m)
Mean t y = 0.023 N m2 (0.018 N m2)
6 Like 1, but with higher pCO2atm and DIC initial conditions (expected 2000 to 2050 changes from CanESM 1.1 A2 simulation).
CanESM 1.1 is the Canadian Earth System Model version 1.1 [Arora et al., 2009; Christian et al., 2010].
Max. initial DIC = 2291 mmol m3 (2273 mmol m3)
pCO2atm = 513 ppmv (370 ppmv)
a
Values in parentheses represent conditions in control experiment 1.

Previous studies on the Oregon shelf have successfully used summer upwelling due to increased atmospheric pCO2 is
2-D models [Allen et al., 1995; Federiuk and Allen, 1995; predicted from a regional climate model with high resolution
Spitz et al., 2003]. Here, a set of model experiments helps to over the coastal region of California [Snyder et al., 2003]. Off
determine the sensitivity of the carbon and O2 cycles to the Vancouver Island shelf, an ensemble of 18 climate models
different forcing and changing conditions. By changing one predicts increased upwelling summer winds in the 21st century
aspect of the model at a time, we compare the differences [Merryfield et al., 2009]. Therefore, we test the sensitivity of
with respect to a control simulation (experiment 1). The the O2 and carbon cycles in the model to increased upwelling.
sensitivity experiments are described below and summarized The summer of 2002 experienced higher than normal
in Table 2. Surface incoming shortwave radiation and net upwelling-favorable winds off Vancouver Island. Mean
heat fluxes are specified in all experiments from NCEP alongshore wind stress was 28 % greater in 2002 than in 1993
reanalysis as daily values for the period 27 May to 29 for the period 27 May to 29 September (0.018 N m2 in
September 1993, at 48.57°N, 125.62°W. Wind stress is cal- 1993 versus 0.023 N m2 in 2002; see Figure 2a). For the
culated from observed hourly winds at meteorological period 16 July to 29 September (the period of analysis fol-
buoy 46206 (48.83°N, 126.00°W, Figure 1) following Smith lowing 50 days of spin-up), mean alongshore wind stress
[1988], then filtered with a 6-hour low-pass Fast Fourier almost tripled in 2002 (0.026 N m2) relative to 1993
Transform (FFT) filter. Atmospheric pCO2 concentration (0.009 N m2). The simulations with stronger upwelling are
(pCO2atm) is set as a constant boundary condition (370 ppmv, forced with 2002 wind stress in two ways: experiment 2 has
except in experiment 6). To provide sufficient time for the same spin-up as the control experiment (first 50 days with
spin-up, since sediments take more than a month to 1993 wind stress), while experiment 3 is forced from the start
approach equilibrium [Bianucci et al., 2011], the experi- with 2002 winds. The first approach allows a comparison of
ments start on 27 May 1993 (day 0) and analyses begin on results after a common forcing during the 50 day spin-up
16 July (day 50). period; the second allows examining the effect of using dif-
2.1. Control Experiment (Experiment 1) ferent forcing during spin-up.
[8] Wind and surface heat forcing corresponds to late 2.3. Shallower OMZ Experiments (Experiments 4
spring and summer 1993, a year representing a “normal and 5)
upwelling summer”: upwelling indices for July and August [10] Although not all species have the same tolerance to
(38 and 26 m3 s1 per meter of coastline, respectively) low O2 concentrations [Vaquer-Sunyer and Duarte, 2008],
were close to the climatological monthly averages (34 and hypoxia is commonly defined as waters with O2 <
22 m3 s1 per meter of coastline; upwelling indices from 60 mmol m3 (= 60 mM  60 mmol kg1  1.4 mL L1)
the Environmental Research Division, Pacific Fisheries [Gray et al., 2002; Whitney et al., 2007; Stramma et al.,
Environmental Laboratory). Observed deep ocean summer 2008]. The hypoxic threshold at Ocean Station Papa
profiles from the study region are used to create average (OSP) has shoaled roughly by 100 m (from 400 to 300 m
depth profiles to initialize scalar properties (i.e., horizon- depth) between 1956 and 2006 [Whitney et al., 2007]. The
tally uniform distributions), and initial velocities are set to southern CCS has experienced a shoaling of up to 90 m in
zero. the period 1984–2006 [Bograd et al., 2008]. Moreover, O2
concentrations in the OMZ are decreasing with observed
2.2. Stronger Upwelling Experiments (Experiments 2 rates of 0.18 mmol-O2 m3 yr1 at a depth of 800 m at
and 3) OSP and 0.15 mmol-O2 m3 yr1 at a depth of 500 m in
[9] Increasing trends in upwelling intensity have been the southern CCS region [Whitney et al., 2007; Bograd et al.,
observed in some major coastal upwelling systems of the 2008]. To analyze the effects of a shallower and more intense
world during the 20th Century [e.g., Bakun, 1990; Schwing OMZ on the southern Vancouver Island shelf system
and Mendelssohn, 1997; McGregor et al., 2007]. Intensified (experiment 4), we modify the initial O2 field such that the

3 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Figure 2. (a) Alongshore wind stress for experiments 1 (1993, black) and 3 (2002, red). Experiment 2 is
forced by 1993 winds up to day 50 (vertical dashed grey line) and 2002 winds afterwards. (b) Initial
O2 profiles (used over the whole domain) for experiments 1 (black) and 4 (red), which coincide in the
upper 130 m. The OMZ hypoxic threshold depth is defined as the depth where open ocean waters
have O2 = 60 mmol m3 (hypoxic threshold, vertical dashed lined). (c) Profiles of decadal means
for dissolved inorganic carbon (DIC) centered on years 2000 and 2050 from the Canadian Earth Sys-
tem Model (CanESM 1.1). (d) Profile of the DIC difference between decadal means, which is added to
initial conditions in experiment 6.

hypoxic threshold is 100 m shallower than in the control predicted changes in DIC and pCO2atm over the period 2000
simulation. Moreover, the minimum concentration is to 2050 from a global climate model (Figures 2c and 2d). We
changed to 1.3 mmol-O2 m3, 9 mmol-O2 m3 lower than use the Canadian Earth System Model (CanESM 1.1), with
the minimun in the control experiment (Figure 2b). This O2 emission scenario A2 [Arora et al., 2009; Christian et al.,
decrease would be achieved in 50 years at the observed rate at 2010] at the model location closest to the Vancouver Island
800 m at OSP. In experiment 5, in addition to the shallower shelf (50°N, 130°W). DIC concentrations in the VICC were
OMZ, the model is forced with the stronger upwelling winds assumed to increase by the same amount as those over the
from 2002. shelf. pCO2atm, which is set as a constant boundary condi-
tion, was increased by 143 ppmv relative to the control
2.4. Higher Carbon Scenario Experiment experiment (from 370 to 513 ppmv).
(Experiment 6)
[11] As CO2 increases in the atmosphere due to anthropo-
3. Respiration Index (RI) and Aragonite
genic activities, roughly one third of emissions are absorbed
by the ocean [Sabine et al., 2004; Sabine and Feely, 2007].
Saturation State (WA)
Experiment 6 examines the effect of increased pCO2atm and [12] Recently, Brewer and Peltzer [2009a] argued that
ocean dissolved inorganic carbon (DIC) on biogeochemical elevated pCO2 may impose a physiological strain on higher
cycles in the model. The increments correspond to the animals, and that the use of an O2 limit alone to define a

4 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Figure 3. Across-shelf distribution of temporally and vertically averaged (a, c) O2 and (b, d) pCO2 for
experiments 1 (blue) and 2 (red). Temporal average is for days 50 to 125, after the spin-up period. Vertical
averages are shown for the upper 30 m of the water column (Figures 3a and 3b) and from 30 m to the sea-
floor (Figures 3c and 3d). (e, f) The bathymetry, with a magenta vertical line indicating the edge of the
shelf break. The dashed grey line in Figure 3b indicates atmospheric pCO2 (370 ppmv).

dead zone implicitly assumes that pCO2 levels are low and seawater is supersaturated (undersaturated) with respect to
inversely proportional to O2. They defined a Respiration aragonite and favors calcification (dissolution). We will
Index that is linearly related to available energy in basic oxic use WA as well as pCO2 to describe the carbon state of the
respiration (RI = log(pO2/pCO2)). The RI reflects the ther- system.
modynamic energy yield of aerobic respiration as the con- [15] WA is calculated from modeled temperature (T),
centration ratio of substrate and product changes. This salinity (S), DIC, and total alkalinity (TA) using the
concept suggests that, as atmospheric pCO2 rises and more CO2SYS software [Lewis and Wallace, 1998]. When com-
carbon is absorbed by the ocean, dead zones could expand paring changes in WA between two simulations, we calculate
even if O2 levels were not affected [Brewer and Peltzer, the total change as
2009a]. RI = 1 can be considered as a general boundary
for aerobic stress, although Brewer and Peltzer [2009a] DWATotal ¼ ðWAn  WA1 Þ=WA1  100% ð1Þ
pointed out that the actual limits will be species-dependent.
[13] Some of the assumptions behind this index aroused where the subscripts 1 and n indicate the control experiment
controversy. For instance, the calculation assumes a closed and any of the sensitivity simulations, respectively. To
thermodynamic system, while living organisms are essen- evaluate the role that DIC alone plays in DWATotal, we com-
tially open systems [Seibel et al., 2009]. These authors also pute DWADIC =(W∗A  WA1)/WA1  100%, where W∗A is calcu-
were concerned about the use of environmental gas partial lated with T, S, and TA from experiment 1 and DIC from
pressures, arguing that the intracellular concentrations are experiment n. DWADIC allows us to quantify the effect of
regulated independently by kinetic and physiological changing only DIC on WA1. The same procedure can be
mechanisms. Despite these criticisms, to which Brewer and performed for every variable (e.g., computation of DWAT,
Peltzer [2009b] responded, we use the RI in the context of DWATA, etc) or for combinations of variables (e.g., the com-
our modeling study to evaluate the combined effect of O2 bined role of TA and DIC leads to DWATA+DIC).
and CO2 as a threshold for habitable shelf environments,
compared with the more typical O2 thresholds. 4. Results
[14] Aragonite is the less stable form of CaCO3 in the
ocean and its degree of saturation can be approximated as 4.1. Effect of Increased Upwelling
WA  [CO2 2 2
3 ]/[CO3 ]sat where [CO3 ] and [CO3 ]sat
2
[16] Temporal and vertical averages of O2 and pCO2
represent the concentrations of the carbonate ions in across the shelf are compared for the control experiment
ambient seawater and at saturation. When WA > 1 (<1), (experiment 1) and experiment 2, which has increased

5 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

18 km) O2 is lower under stronger upwelling relative to the


control experiment 1. These changes respond to the position
of phytoplankton blooms in both simulations, since phyto-
plankton are advected offshore near the surface during
intense upwelling events (Figure 4). Regression analysis
predicts a 20 km displacement of the maximum primary
production from the inshore boundary for a mean alongshore
equatorward wind stress of 0.05 N m2 in the previous
9 days [Bianucci, 2010]. The bulk of the primary production
(which increases by 12%) occurs farther offshore in experi-
ment 2, over the mid- and outer shelf (between 10 and
40 km, Figure 4). Moreover, high primary production over
the mid-shelf leads to more organic matter reaching the
seafloor at those depths, contributing to enhanced exchange
of O2 and DIC with the sediments. Between 20 and 30 km
from the inshore boundary (i.e., at bottom depths between 90
and 115 m) O2 and DIC exchanges with the sediments
increase up to 50% (the increment is 8% on average).
[18] As the development of hypoxia and/or corrosive
conditions first occurs in the bottom waters over the shelf,
we focus on the spatial and temporal distributions of O2 and
WA in the near-bottom layer of the model (Figure 5). This
layer accurately represents the bottom boundary layer,
except within the shallowest 3 km of the domain where
restoring to the VICC is strongest [Bianucci et al., 2011].
With increased upwelling as in experiment 2, low O2 con-
centrations develop in the bottom layer inshore of the shelf
Figure 4. Temporal evolution (Hovmöller plots) for total break, reaching a minimum of 34 mmol-O2 m3 (Figure 5c).
water column primary production in experiments (a) 1 and In particular, the hypoxic boundary (60 mmol-O2 m3, bold
(b) 2. The dashed magenta line represents the location of black contour) migrates inshore across the shelf after the
the shelf break. On the right, alongshore wind stress for each intense upwelling-favorable wind event centered on day 80
experiment (upwelling/downwelling in red/blue) is shown; (see Figure 2a). The onset of hypoxia in shallower waters is
the horizontal dashed line denotes day 50, the end of spin- triggered by advection of low oxygen during that event (red
up period with common 1993 forcing. Primary production line in Figure 6). Once upwelling intensity decreases, bio-
units are g-C m2 d1. logical consumption (especially in the sediments; black line
in Figure 6) maintains the low O2 levels achieved by
advection. The enhanced advection of low O2 results partly
upwelling (Figure 3, blue and red lines, respectively). The from the stronger upwelling circulation and partly from the
upper 30 m of the water column experiences higher O2 lower O2 content of upwelled waters (the depth of upwelling
concentrations under intensified upwelling over most of the increases 40 m in experiment 2). Since advection and local
shelf (Figure 3a), while below 30 m depth these wind con- biological consumption generate the lowest O2 concentra-
ditions decrease O2 concentrations (Figure 3c). The O2 tions in bottom waters, vertical mixing generates a down-
changes are up to +7 and 11% in the upper and lower ward flux of O2 to the bottom layers from the overlying,
region of the water column, respectively. The inner shelf more oxygenated waters (blue line in Figure 6).
concentrations are similar in both simulations due to the [19] The wind-forcing used during the first 50 days of the
restoring to VICC properties. On average, pCO2 in the upper experiments with stronger upwelling affects the results. An
ocean is lower than pCO2atm (dashed grey line in Figure 3b) intense upwelling event between days 35 to 45 in 1993
over the mid- and outer shelf in both experiments, so air-sea (first week of July) advects the hypoxic threshold closer
CO2 fluxes are from the atmosphere to the ocean in those to the shelf break. Therefore, if 1993 winds are used during
regions (surface pCO2 has a similar pattern but lower values spin-up (experiment 2, Figure 5c), waters on the shelf
than the average over the upper 30 m). However, the high become more hypoxic than if 2002 winds are used from day 0
DIC concentrations specified in the VICC lead to pCO2 (experiment 3, Figure 5e). From day 50 to 125, hypoxia
concentrations higher than pCO2atm over the inner shelf. covers 43% of the near-bottom waters over the shelf in
CO2 outgassing occurs on the inner 4 km of the model experiment 2 compared with 18% in experiment 3. More-
domain in experiment 1 and within 6 km in experiment 2. over, the minimum O2 concentration is lower in experiment 2
[17] In experiment 2, pCO2 is higher in the bottom layers relative to experiment 3 (34 versus 41 mmol-O2 m3).
(Figure 3d) and lower in the upper 30 m over most of the Hence, the timing of the onset of the upwelling season is
model domain (Figure 3b), with changes reaching up to +4 another factor determining both the timing and magnitude of
and 7% in each case. However, there is a region of the hypoxia (and analogously, of events with lower aragonite
inner shelf (10–25 km from inshore boundary) where saturation state).
pCO2 is higher in the upper layer under intensified upwell- [20] Dissolved carbon concentrations in the near-bottom
ing; moreover, approximately in the same region (10– layer over the shelf respond to the same processes as

6 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Figure 5. Hovmöller plots for near-bottom (left) O2 and (right) WA for experiments (a, b) 1, (c, d) 2, (e) 3,
(f) 6, (g) 4 and (h) 5. Spin-up period (first 50 days) not shown. The bold black contours represent either the
hypoxic threshold (60 mmol-O2 m3) or the limit for aragonite dissolution (WA = 1). The bold yellow con-
tour is RI = 1 (area with RI < 1 is indicated) and the dashed magenta line represents the location of the
shelf break. The black dashed line in Figure 5c indicates the 90 m isobath; the dash-dotted yellow lines
in Figure 5g show the 130 and 90 m isobaths.

dissolved O2: DIC increases due to advection of high con- experiment 2, WA decreases in the near-bottom layer over the
centrations from offshore, while sediment remineralization shelf near day 90 (Figure 5d) relative to the control experi-
maintains high DIC over the mid-shelf during relaxation and ment 1 (Figure 5b). As WA depends on T, S, DIC and TA, we
periods with weak winds (DIC budget terms not shown). In evaluate the contribution of each variable to the total change

7 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Figures 8a and 8b) reaches the shelf break in experiment 4,


penetrating up to depths of 135 m. During the 75 days of
analysis, the near-bottom hypoxic threshold in experiment 4
is either near the shelf break or over the outer shelf (bold
black contour in Figure 5g). The threshold penetrates
onshore after strong upwelling events (after days 40 and
108, see Figure 2a). The lower O2 concentrations with
respect to the control experiment 1 respond to two factors:
(1) the initial conditions, which are already lower for depths
greater than 130 m (Figure 2b), and (2) the enhanced
advection of lower O2 from offshore onto the mid- and outer
shelf (see O2 budget in Figure 9). Biological sinks (both in
the water column and the sediments) remain essentially
unchanged over the shelf between experiments 1 and 4
(Figure 9), which differ only in the initial profiles of O2
concentrations. Since the lower bottom O2 concentrations
increase the vertical gradient of O2, vertical mixing is
enhanced in experiment 4 and partly compensates the
decrease in O2 by advection.
[22] When the modified initial conditions with a shallower
OMZ are used in combination with stronger upwelling-
favorable winds (experiment 5), O2 concentrations decrease
Figure 6. (top) Time series of O2 fluxes from advection even more in the near-bottom layer over the shelf
(red), vertical mixing (blue), and biological O2 sinks (black, (Figure 5h). Most of the bottom waters over the shelf
remineralization within the sediments; gray, remineralization become hypoxic after the strong upwelling event around day
plus nitrification in the water column) for experiment 2 in 80, except for the area influenced by the O2-rich VICC. O2
the bottom 10 m of the water column at the 90 m isobath concentrations drop to 21 mmol-O2 m3 near the shelf
(dashed black line in Figure 5c). Positive (negative) fluxes break, the lowest over the shelf for all the experiments pre-
indicate a gain (loss) of O2. (bottom) Alongshore wind sented here. Concentrations lower than 20 mmol-O2 m3 (or
stress: upwelling occurs when t y < 0. 0.5 mL L1) are usually considered to represent “severe
hypoxia” [e.g., Monteiro et al., 2006; Chan et al., 2008;
Diaz and Rosenberg, 2008].
in WA between both simulations as explained in section 3 [23] In experiments 4 and 5 (Figures 5g and 5h), the
(Figure 7). The increase in DIC by upwelling is primarily region with RI ≤ 1 expands greatly compared with experi-
responsible for the drop in WA (Figure 7a), while the increase ments 1 to 3 (Figures 5a, 5c, and 5e; bold yellow contours
in TA cancels out about 40% of the DIC effect (Figure 7b). indicate the RI = 1 contour). The decrease in RI is mainly
Upwelling also brings colder waters from offshore that due to the lower O2 concentrations in the near-bottom layer
reduce the temperature near the bottom over the shelf by of the open ocean, since pCO2 does not change significantly
up to 3°C (the mean cooling over the near-bottom shelf is in those waters. In particular, the shoaling of the OMZ does
0.63°C), which tends to decrease WA by increasing not significantly affect the carbon cycle, e.g., differences
3 ]sat (section 3). The reduction in WA due to cooling
[CO2 between experiments 1 and 4 for pCO2 and WA are < 0.6%
(Figure 7c) is small compared with the change due to and 0.4%, respectively (not shown).
increasing DIC (Figure 7a). The changes in salinity due to
stronger upwelling do not significantly affect WA over the 4.3. Effect of Higher Inorganic Carbon
near-bottom of the shelf outside the VICC region (Figure 7c). [24] Higher DIC initial conditions (experiment 6, see
The combined effect of DIC and TA (Figure 7d) accounts for Figures 2c and 2d) increase pCO2 throughout the model
essentially all of the change in near-bottom WA between domain (Figure 10). The bold black contours in the averaged
experiments 1 and 2 (black bars in Figure 7). pCO2 vertical sections correspond to the atmospheric
pCO2atm values in experiments 1 (370 ppmv, Figure 10a)
4.2. Effect of a Shallower Offshore OMZ and 6 (513 ppmv, Figure 10b). Over most of the shelf, the
[21] On average (from day 50 to 125), the bottom waters surface ocean is on average undersaturated in both simula-
over the shelf have lower O2 in the experiment with a shal- tions and absorbs CO2 from the atmosphere. However,
lower OMZ (experiment 4) compared with the control surface VICC waters within 4 km from the inshore
experiment 1 (Figure 8). Most of the shelf has bottom con- boundary are oversaturated in the control experiment 1
centrations below 80 mmol-O2 m3 in experiment 4, while (releasing CO2 to the atmosphere), in contrast to experi-
in experiment 1 shelf waters are mainly above that level ment 6. Aragonite undersaturation expands considerably
(waters between 60 and 80 mmol-O2 m3 are shaded in over the shelf in the latter scenario. The average saturation
Figure 8). The upper layers of both simulations remain horizon (WA = 1) coincides with averaged pCO2 values of
similar, since their initialization is the same (Figure 2b) and 710 and 800 ppmv in experiments 1 and 6, respectively
the changes in deeper waters do not greatly modify biology (Figure 10). The sensitivity of WA to pCO2 varies between
in the upper ocean or air-sea O2 exchange. The mean hyp- these experiments due to the different effect of DIC on WA
oxic threshold (60 mmol-O2 m3, bold black line in and pCO2. We can approximate the former using the

8 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Figure 7. Histograms of WA change (DWA) in the near-bottom layer of experiment 2 relative to experi-
ment 1 due to the contributions of individual and combined variables: (a) dissolved inorganic carbon
(DIC), (b) total alkalinity (TA), (c) temperature (T) and salinity (S), and (d) the combination of TA and
DIC (TA + DIC). Black histograms are the same in all plots and show the total change in near-bottom
layer WA between experiments (DWATotal). The histograms do not include the spin-up period or the area with
strong VICC restoring on the inner 3 km.

carbonate ion concentration in seawater and at saturation


[Sarmiento and Gruber, 2006]: WA  [CO2 3 ]/[CO3 ]sat 
2
2
(TA-DIC)/[CO3 ]sat. Hence, an increase in DIC leads to lower
WA. However, pCO2 can be approximated by (2  DIC 
TA)2/(TA-DIC), such that higher DIC increases pCO2.
[25] In experiment 1, the saturation horizon in the open
ocean is below 250 m (not visible in Figure 10a). However,
in experiment 6, the saturation horizon shoals 250 m in the
open ocean with respect to experiment 1 and extends over
the shelf up to 30 m above the seafloor (Figure 10b).
Moreover, the saturation state in experiment 6 is below unity
(reaching down to 0.78) after day 50 in the near-bottom
layer over the whole shelf seaward of 8 km from the inshore
boundary (Figure 5f). There, the VICC carries waters with
lower DIC relative to near-bottom waters over the shelf
[Bianucci et al., 2011, and references therein]. However, in
the control experiment 1, near-bottom waters with WA ≤ 1 are
found only offshore from 45 km from the inshore bound-
ary (Figure 5b). Figure 8. Mean O2 vertical sections down to 250 m
[26] The area with RI < 1 expands modestly in experiment 6 (average for days 50 to 125) for experiments (a) 1 and
compared with the control experiment 1 (yellow bold contour (b) 4. Bold black contours indicate the hypoxic threshold
in Figure 5f). In waters deeper than 1000 m, the modified (60 mmol-O2 m3); the areas shaded in yellow emphasize
initial conditions do not show a large increase in DIC waters between 60 and 80 mmol-O2 m3. The dashed
(Figure 2f). Consequently, pCO2 increases only moderately in magenta line indicates the position of the shelf break.

9 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

the near-bottom waters of the slope (where RI < 1, at 60 km


from the inshore boundary). If RI is calculated using pO2 from
experiment 4 instead, the area with indices values <1 is almost
identical to that of experiment 4 (only slightly larger). The
combination of pO2 from experiment 4 and pCO2 from
experiment 6 is consistent with an scenario of low-O2 and
high-DIC, since T and S are the same in both simulations and
changes in TA are minimal.

5. Discussion and Conclusions


[27] The experiments described here provide insight into
the sensitivity of the carbon and O2 cycles to changes in
forcing or initial conditions in a seasonal upwelling region.
Both cycles respond to changes in upwelling in our model
experiments. However, only O2 concentrations are affected
by the shoaling of the OMZ and only the carbon cycle
responds to changes in the oceanic DIC inventory. None-
theless, changes in the depth of the OMZ and carbon content
of the ocean will not be independent of each other, as both
are expected to respond to the increase in atmospheric CO2
from human activities. Moreover, these changes will be
coupled to modifications in the cycling of other nutrients.
Given the complexity of the feedbacks between biogeo-
Figure 9. Source and sink fluxes of O2 in the bottom 10 m chemical cycles, the introduction of a single change per
of the water column at locations where the bottom depth is experiment allows a clearer understanding of the responses
(top) 90 m and (bottom) 130 m (dash-dotted yellow lines in to any given change.
Figure 5g). Fluxes are integrated from day 50 to 125. Experi- [28] This study extends that of Ianson and Allen [2002],
ments 1 and 4 shown in black and white bars, respectively. which described results from a compartment model. Our
Abbreviations on the horizontal axis represent advection quasi 2-D model [Bianucci et al., 2011] can resolve the
(horizontal plus vertical), vertical mixing, exchanges with spatiotemporal response to wind-forcing in both the vertical
the sediments, water column remineralization of semilabile and across-shelf directions. Moreover, the current work
dissolved organic matter and detritus, and water column focuses on sensitivities to climate-related change of inter-
nitrification. actions between the carbon system and O2, which was not
included in the model of Ianson and Allen [2002].

Figure 10. Mean pCO2 vertical sections down to 250 m (average for days 50 to 125) for experiments
(a) 1 and (b) 6. Bold black contours show where the pCO2 is equal to the surface atmospheric pCO2:
(Figure 10a) 370 ppmv; (Figure 10b) 513 ppmv. The bold white contour in Figure 10b shows the sat-
uration horizon (WA = 1), which lies below 250 m depth in Figure 10a. The dashed magenta line indi-
cates the position of the shelf break.

10 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

[29] Strong upwelling events (such as the one at day 80 [32] RI calculated for a hypothetical scenario with lower
in experiment 2) result in onshore advection of low O2 and O2 and higher DIC showed a higher sensitivity of this index
high DIC from offshore that triggers low O2 and WA events to O2 than to CO2 changes. This difference results from the
in the near-bottom layers over the shelf. The WA decrease is changes in the initial conditions in experiments 4 and 6,
partially compensated by the onshore advection of high TA since the fractional decrease of initial O2 is larger than the
(advection of cooler waters has a much smaller effect). In fractional increase of DIC (both changes aim to represent a
addition to the intensified onshore advection under increased projection 50 years into the future).
upwelling scenarios, O2 consumption and DIC production in [33] Under elevated inorganic carbon conditions (experi-
the sediments are enhanced over the mid-shelf due to off- ment 6), near-bottom waters over the shelf (beyond the area
shore transport of near-surface phytoplankton blooms. The of influence of the VICC) become corrosive after 50 days,
latter increases deposition of organic matter to the seafloor i.e., WA drops below unity. In the current model configu-
over the mid-shelf, enhancing the exchange of O2 and DIC ration, no biological process depends on DIC concentra-
between the water column and the sediments. Therefore, low tion (e.g., calcification is not included), so the increase of
O2 and WA values are maintained after the upwelling events DIC in the ocean does not affect the O2 or nitrogen cycles
by remineralization within the sediments. O2 concentrations directly. Decreased calcification in a carbon-rich ocean
in particular are demonstrated to be sensitive to intensified may affect photosynthesis and phytoplankton community
upwelling, since hypoxia develops in the near-bottom waters structure [Sikes et al., 1980; Paasche, 2001], providing a
on the shelf in both simulations with stronger upwelling potential link between O2 and carbon. Furthermore, the
(experiments 2 and 3). model is missing a negative feedback in the carbon cycle
[30] Differences between experiments 2 and 3 (which that could potentially reduce acidification: a reduction of
differ only in the wind conditions during the spin-up period calcification (increase of CaCO3 dissolution) reduces the
of 50 days) demonstrate the effect of the timing of the onset production (enhances the consumption) of CO2 in seawater
of the upwelling season. If summer upwelling is stronger through the reaction 2HCO 3 + Ca
2 +
⇌ CO2 + CaCO3 +
and starts earlier (experiment 2), O2-poor and DIC-rich near- H2O. Most published works agree on the reduction of
bottom waters are advected upwards closer to the shelf break calcification rates with higher pCO2, although Iglesias-
earlier in the season. Thus, subsequent upwelling events Rodriguez et al. [2008] reported the opposite and
further reduce O2 and WA in the near-bottom layers over the aroused some controversy [see Riebesell et al., 2008].
shelf. A regional climate model of the northern California Blooms of calcifiers (e.g., coccolithophorids) are not fre-
shelf, with 40 km horizontal resolution, projects a 1-month quent over the Vancouver Island shelf, although they have
delay of the onset of seasonal upwelling for increased been observed (D. Ianson, personal communication, 2010).
atmospheric CO2 (560 to 686 ppmv), as well as intensified [34] We need to understand how biogeochemical cycles
upwelling [Snyder et al., 2003]. In the context of the present will respond to climate change in the coastal ocean, since we
study, a delayed upwelling season would reduce the poten- depend on its resources. The present work contributes
tial for hypoxia and acidification on the Vancouver Island toward that goal by focusing on the sensitivities of biogeo-
shelf. In contrast, Barth et al. [2007] reported on the nega- chemical cycling to environmental factors, individually and
tive consequences of a one-month delay in the 2005 transi- collectively, that may change (in most cases, certainly will)
tion to upwelling-favorable wind stress on the California and as the anthropocene evolves. The perturbations analyzed
Oregon shelves (warm waters, low nutrient levels, low pri- here (increased upwelling, shallower OMZ, and higher DIC
mary productivity, and low recruitment of rocky intertidal and pCO2atm) consistently drive the system toward lower O2,
organisms). pH, and WA states. These results emphasize the potential
[31] The effect of a shallower OMZ (experiment 4) is negative impact that these perturbations may have on ben-
straightforward: it moves the hypoxic threshold closer to the thic ecosystems, along the lines of the consequences first
shelf break, so upwelling events are more likely to advect hypothesized by Bakun [1990] for increased upwelling.
O2-poor waters onto the shelf. Therefore, the combination of Therefore, this sensitivity analysis suggests that a region
stronger upwelling and a shallower OMZ (experiment 5) such as the Vancouver Island shelf could develop hypoxia
would further reduce O2 concentrations on the shelf, repre- and corrosive conditions (WA < 1) in the future. Full 3-D
senting a potential increasing stress on ecosystems. The biogeochemical modeling as well as continuous and simul-
present model represents the shelf off Vancouver Island: taneous observations of carbon and O2 are needed to assess
despite the considerable shelf width, the influence of a the current state of these cycles in the coastal ocean and how
shallower OMZ reaches the inner shelf (except for the region they vary in time and space.
influenced by the O2-rich VICC). The effect could be more
severe on narrower shelves, where upwelling can transport [35] Acknowledgments. We thank K. Fennel for her constructive
comments on an earlier version of this manuscript, as well as the valuable
offshore waters to even shallower depths onshore. The input from two anonymous reviewers. L.B. acknowledges several sources
expected overall decrease in O2 concentration in the sub- of graduate support: a UVic Fellowship, a Maritime Award Society of
surface open ocean would lead to an expansion of the region Canada Scholarship, and a Bob Wright Fellowship. Computing support
was provided by the Canadian Centre for Climate Modeling and Analysis,
with RI < 1, implying a larger area where aerobic marine life Environment Canada. K.D. acknowledges funding from an NSERC Dis-
would be under stress [Brewer and Peltzer, 2009a]. In none covery Grant.
of the experiments performed did RI reach values below
unity over the shelf (although O2 levels did decline below References
the 60 mmol-O2 m3 hypoxic threshold). These results
Allen, J., P. Newberger, and J. Federiuk (1995), Upwelling circulation on
suggest that RI is an indicator of more severe aerobic stress the Oregon continental shelf. Part I: Response to idealized forcing,
than the 60 mmol-O2 m3 threshold. J. Phys. Oceanogr., 25(8), 1843–1866.

11 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Arora, V., G. Boer, J. Christian, C. Curry, K. Denman, K. Zahariev, Haidvogel, D., et al. (2008), Ocean forecasting in terrain-following coordi-
G. Flato, J. Scinocca, W. Merryfield, and W. Lee (2009), The effect of nates: Formulation and skill assessment of the Regional Ocean Model-
terrestrial photosynthesis down regulation on the twentieth-century car- ing System, J. Comput. Phys., 227(7), 3595–3624, doi:10.1016/j.jcp.
bon budget simulated with the CCCma Earth System Model, J. Climate, 2007.06.016.
22, 6066–6088, doi:10.1175/2009JCLI3037.1. Hauri, C., N. Gruber, G. Plattner, S. Alin, R. Feely, B. Hales, and
Bakun, A. (1990), Global climate change and intensification of coastal P. Wheeler (2009), Ocean acidification in the California Current System,
ocean upwelling, Science, 247(4939), 198–201, doi:10.1126/science.247. Oceanography, 22(4), 60–71.
4939.198. Helly, J. J., and L. A. Levin (2004), Global distribution of naturally occur-
Barth, J. A., B. A. Menge, J. Lubchenco, F. Chan, J. M. Bane, A. R. ring marine hypoxia on continental margins, Deep Sea Res., Part I, 51(9),
Kirincich, M. A. McManus, K. J. Nielsen, S. D. Pierce, and L. Washburn 1159–1168, doi:10.1016/j.dsr.2004.03.009.
(2007), Delayed upwelling alters nearshore coastal ocean ecosystems in Ianson, D., and S. E. Allen (2002), A two-dimensional nitrogen and carbon
the northern California current, Proc. Natl. Acad. Sci. U.S.A., 104(10), flux model in a coastal upwelling region, Global Biogeochem. Cycles,
3719–3724, doi:10.1073/pnas.0700462104. 16(1), 1011, doi:10.1029/2001GB001451.
Bates, N. R. (2007), Interannual variability of the oceanic CO2 sink in Ianson, D., S. E. Allen, S. L. Harris, K. J. Orians, D. E. Varela, and C. S.
the subtropical gyre of the North Atlantic Ocean over the last 2 decades, Wong (2003), The inorganic carbon system in the coastal upwelling
J. Geophys. Res., 112, C09013, doi:10.1029/2006JC003759. region west of Vancouver Island, Canada, Deep Sea Res., Part I, 50(8),
Bianucci, L. (2010), Carbon, oxygen, and nitrogen cycles on the Vancouver 1023–1042, doi:10.1016/S0967-0637(03)00114-6.
Island shelf, PhD thesis, Univ. of Victoria, Victoria, B. C., Canada. Iglesias-Rodriguez, et al. (2008), Phytoplankton calcification in a high-CO2
Bianucci, L., K. L. Denman, and D. Ianson (2011), Low oxygen and high world, Science, 320(5874), 336–340, doi:10.1126/science.1154122.
inorganic carbon on the vancouver island shelf, J. Geophys. Res., 116, Keeling, R. F., and H. E. Garcia (2002), The change in oceanic O2 inven-
C07011, doi:10.1029/2010JC006720. tory associated with recent global warming, Proc. Natl. Acad. Sci. U.S.
Bindoff, N., et al. (2007), Observations: Oceanic climate change and sea A., 99(12), 7848–7853, doi:10.1073/pnas.122154899.
level, in Climate Change 2007: The Physical Science Basis. Contribu- Lathuilière, C., V. Echevin, M. Lévy, and G. Madec (2010), On the role of
tion of Working Group I to the Fourth Assessment Report of the Inter- the mesoscale circulation on an idealized coastal upwelling ecosystem,
governmental Panel on Climate Change, edited by S. Solomon et al., J. Geophys. Res., 115, C09018, doi:10.1029/2009JC005827.
pp. 385–432, Cambridge Univ. Press, Cambridge, U. K. Lewis, E., and D. Wallace (1998), Program developed for CO2 system cal-
Bograd, S. J., C. G. Castro, E. Di Lorenzo, D. M. Palacios, H. Bailey, culations, Tech. Rep. oRNL/CDIAC-105, Carbon Dioxide Inf. Anal.
W. Gilly, and F. P. Chavez (2008), Oxygen declines and the shoaling Cent., Oak Ridge Natl. Lab., U.S. Dep. of Energy, Oak Ridge, Tenn.
of the hypoxic boundary in the California Current, Geophys. Res. Lett., Masson, D., and P. F. Cummins (1999), Numerical simulations of a
35, L12607, doi:10.1029/2008GL034185. buoyancy-driven coastal countercurrent off Vancouver Island, J. Phys.
Brewer, P. G., and E. T. Peltzer (2009a), Limits to marine life, Science, 324 Oceanogr., 29(3), 418–435.
(5925), 347–348, doi:10.1126/science.1170756. McGregor, H. V., M. Dima, H. W. Fischer, and S. Mulitza (2007), Rapid
Brewer, P. G., and E. T. Peltzer (2009b), Response to B. Seibel et al’s. E- 20th-Century Increase in coastal upwelling off Northwest Africa, Science,
Letter on “Limits to marine life,” Science, 324, 347–348. 315(5812), 637–639, doi:10.1126/science.1134839.
Chan, F., J. A. Barth, J. Lubchenco, A. Kirincich, H. Weeks, W. T. Peterson, Merryfield, W. J., B. Pal, and M. G. G. Foreman (2009), Projected future
and B. A. Menge (2008), Emergence of anoxia in the California Current changes in surface marine winds off the west coast of Canada, J. Geo-
Large Marine Ecosystem, Science, 319(5865), 920–921, doi:10.1126/ phys. Res., 114, C06008, doi:10.1029/2008JC005123.
science.1149016. Monteiro, P. M. S., A. van der Plas, V. Mohrholz, E. Mabille, A. Pascall,
Christian, J. R., et al.(2010), The global carbon cycle in the Canadian Earth and W. Joubert (2006), Variability of natural hypoxia and methane in a
System Model (CanESM1): Preindustrial control simulation, J. Geophys. coastal upwelling system: Oceanic physics or shelf biology?, Geophys.
Res., 115, G03014, doi:10.1029/2008JG000920. Res. Lett., 33, L16614, doi:10.1029/2006GL026234.
Crawford, W., and R. Dewey (1989), Turbulence and mixing: Sources Oschlies, A., K. G. Schulz, U. Riebesell, and A. Schmittner (2008), Simu-
of nutrients on the Vancouver Island continental shelf, Atmos. Ocean, lated 21st century’s increase in oceanic suboxia by CO2-enhanced biotic
27(2), 428–442. carbon export, Global Biogeochem. Cycles, 22, GB4008, doi:10.1029/
Diaz, R. J., and R. Rosenberg (2008), Spreading dead zones and consequences 2007GB003147.
for marine ecosystems, Science, 321(5891), 926–929, doi:10.1126/science. Paasche, E. (2001), A review of the coccolithophorid Emiliania huxleyi
1156401. (Prymnesiophyceae), with particular reference to growth, coccolith for-
Dore, J. E., R. Lukas, D. W. Sadler, M. J. Church, and D. M. Karl (2009), mation, and calcification-photosynthesis interactions, Phycologia, 40(6),
Physical and biogeochemical modulation of ocean acidification in the cen- 503–529.
tral North Pacific, Proc. Natl. Acad. Sci. U.S.A., 106(30), 12,235–12,240, Rabalais, N., L. Levin, R. Turner, D. Gilbert, and J. Zhang (2010), Dynam-
doi:10.1073/pnas.0906044106. ics and distribution of natural and human-caused hypoxia, Biogeo-
Falkowski, P., et al. (2000), The Global Carbon Cycle: A test of our knowl- sciences, 7, 585–619, doi:10.5194/bg-7-585-2010.
edge of Earth as a system, Science, 290(5490), 291–296, doi:10.1126/ Riebesell, U., et al. (2007), Enhanced biological carbon consumption in a
science.290.5490.291. high co2 ocean, Nature, 450(7169), 545–548.
Federiuk, J., and J. Allen (1995), Upwelling circulation on the Oregon Riebesell, U., R. G. J. Bellerby, A. Engel, V. J. Fabry, D. A. Hutchins,
continental shelf. Part II: Simulations and comparisons with observations, T. B. H. Reusch, K. G. Schulz, and F. M. M. Morel (2008), Comment
J. Phys. Oceanogr., 25(8), 1867–1889. on “Phytoplankton calcification in a high-CO2 world,” Science, 322(5907),
Feely, R. A., C. L. Sabine, J. M. Hernandez-Ayon, D. Ianson, and B. Hales 1466, doi:10.1126/science.1161096.
(2008), Evidence for upwelling of corrosive “acidified” water onto the Rykaczewski, R. R., and J. P. Dunne (2010), Enhanced nutrient supply to
continental shelf, Science, 320(5882), 1490–1492, doi:10.1126/ the California Current Ecosystem with global warming and increased
science.1155676. stratification in an earth system model, Geophys. Res. Lett., 37,
Freeland, H., W. Crawford, and R. Thomson (1984), Currents along the L21606, doi:10.1029/2010GL045019.
Pacific coast of Canada, Atmos. Ocean, 22, 151–172. Sabine, C., and R. Feely (2007), The oceanic sink for carbon dioxide, in
Frölicher, T. L., F. Joos, G.-K. Plattner, M. Steinacher, and S. C. Doney Greenhouse Gas Sinks, edited by D. S. Reay et al., chap. 3, pp. 31–49,
(2009), Natural variability and anthropogenic trends in oceanic oxygen CABI Publ., Oxfordshire, U. K.
in a coupled carbon cycle-climate model ensemble, Global Biogeochem. Sabine, C. L., et al. (2004), The oceanic sink for anthropogenic CO2,
Cycles, 23, GB1003, doi:10.1029/2008GB003316. Science, 305(5682), 367–371, doi:10.1126/science.1097403.
Grantham, B. A., F. Chan, K. J. Nielsen, D. S. Fox, J. A. Barth, A. Huyer, Santana-Casiano, J. M., M. González-Dávila, M.-J. Rueda, O. Llinás, and
J. Lubchenco, and B. A. Menge (2004), Upwelling-driven nearshore E.-F. González-Dávila (2007), The interannual variability of oceanic
hypoxia signals ecosystem and oceanographic changes in the northeast CO2 parameters in the northeast Atlantic subtropical gyre at the
Pacific, Nature, 429(6993), 749–754, doi:10.1038/nature02605. ESTOC site, Global Biogeochem. Cycles, 21, GB1015, doi:10.1029/
Gray, J. S., R. S. sun Wu, and Y. Y. Or (2002), Effects of hypoxia and 2006GB002788.
organic enrichment on the coastal marine environment, Mar. Ecol. Prog. Sarmiento, J., and N. Gruber (2006), Ocean Biogeochemical Dynamics,
Ser., 238, 249–279, doi:10.3354/meps238249. 526 pp., Princeton Univ. Press, Princeton, N. J.
Gruber, N., H. Frenzel, S. C. Doney, P. Marchesiello, J. C. McWilliams, Schwing, F., and R. Mendelssohn (1997), Increased coastal upwelling in
J. R. Moisan, J. J. Oram, G.-K. Plattner, and K. D. Stolzenbach (2006), the California Current System, J. Geophys. Res., 102(C2), 3421–3438.
Eddy-resolving simulation of plankton ecosystem dynamics in the cali- Seibel, B. A., P. R. Girguis, and J. J. Childress (2009), Critique of the res-
fornia current system, Deep Sea Res., Part I, 53(9), 1483–1516, piration index of “Limits to marine life,” Science, 324, 347–348.
doi:10.1016/j.dsr.2006.06.005.

12 of 13
21562202g, 2012, G1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JG001849 by Cochrane Philippines, Wiley Online Library on [04/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
G01020 BIANUCCI AND DENMAN: CARBON AND O2 SENSITIVITY TO FORCING G01020

Shaffer, G., S. M. Olsen, and J. O. P. Pedersen (2009), Long-term ocean Strub, P. T., J. S. Allen, A. Huyer, R. L. Smith, and R. C. Beardsley
oxygen depletion in response to carbon dioxide emissions from fossil (1987b), Seasonal cycles of currents, temperatures, winds, and sea level
fuels, Nature Geosci., 2, 105–109, doi:10.1038/ngeo420. over the northeast pacific continental shelf: 35°N to 48°N, J. Geophys.
Sikes, C., R. Roer, and K. Wilbur (1980), Photosynthesis and coccolith for- Res., 92(C2), 1507–1526.
mation: Inorganic carbon sources and net inorganic reaction of deposi- Thomson, R. E., B. M. Hickey, and P. H. LeBlond (1989), The Vancouver
tion, Limnol. Oceanogr., 25(2), 248–261. Island coastal current: Fisheries barrier and conduit, in Effects of Ocean
Smith, S. D. (1988), Coefficients for sea surface wind stress, heat flux, and Variability on Recruitment and an Evaluation of Parameters Used in
wind profiles as a function of wind speed and temperature, J. Geophys. Stock Assessment Models, edited by G. McFarlane, pp. 265–296, Dep.
Res., 93(C12), 15,467–15,472. of Fish. and Oceans, Ottawa, Ont., Canada.
Snyder, M. A., L. C. Sloan, N. S. Diffenbaugh, and J. L. Bell (2003), Future Vaquer-Sunyer, R., and C. M. Duarte (2008), Thresholds of hypoxia
climate change and upwelling in the California Current, Geophys. Res. for marine biodiversity, Proc. Natl. Acad. Sci. U.S.A., 105(40),
Lett., 30(15), 1823, doi:10.1029/2003GL017647. 15,452–15,457, doi:10.1073/pnas.0803833105.
Spitz, Y. H., P. A. Newberger, and J. S. Allen (2003), Ecosystem response Whitney, F. A., H. J. Freeland, and M. Robert (2007), Persistently declining
to upwelling off the Oregon coast: Behavior of three nitrogen-based mod- oxygen levels in the interior waters of the eastern subarctic Pacific, Prog.
els, J. Geophys. Res., 108(C3), 3062, doi:10.1029/2001JC001181. Oceanogr., 75(2), 179–199, doi:10.1016/j.pocean.2007.08.007.
Stramma, L., G. C. Johnson, J. Sprintall, and V. Mohrholz (2008), Expand-
ing oxygen-minimum zones in the tropical oceans, Science, 320(5876), L. Bianucci, Department of Oceanography, Dalhousie University, 1355
655–658, doi:10.1126/science.1153847.
Oxford St., Halifax, NS B3H 4J1, Canada. (laura.bianucci@dal.ca)
Strub, P. T., J. S. Allen, A. Huyer, and R. L. Smith (1987a), Large-scale K. L. Denman, VENUS Project, University of Victoria, PO Box 1700,
structure of the spring transition in the coastal ocean off western north STN CSC, Victoria, BC V8W 2Y2, Canada. (denmank@uvic.ca)
america, J. Geophys. Res., 92(C2), 1527–1544.

13 of 13

You might also like