You are on page 1of 12

Received: 5 December 2017 | Revised: 3 April 2018 | Accepted: 3 April 2018

DOI: 10.1111/gcb.14171

PRIMARY RESEARCH ARTICLE

Counterintuitive effects of global warming-induced wind


patterns on primary production in the Northern Humboldt
Current System

n
Rodrigo Mogollo | Paulo H. R. Calil

Laboratorio de Dina^mica e Modelagem


Oce^anica - DinaMO, Instituto de Abstract
Oceanografia- Universidade Federal do Rio It has been hypothesized that global warming will strengthen upwelling-favorable
Grande - FURG, Rio Grande, RS, Brazil
winds in the Northern Humboldt Current System (NHCS) as a consequence of the
Correspondence increase of the land–sea thermal gradient along the Peruvian coast. The effect of
n, Laborato
Rodrigo Mogollo rio de Din^amica
e Modelagem Oce^anica - DinaMO, Instituto strengthened winds in this region is assessed with the use of a coupled physical–bio-
de Oceanografia- Universidade Federal do geochemical model forced with projected and climatological winds. Strengthened
Rio Grande - FURG, Rio Grande, RS, Brazil.
Email: rodrigo.mogollon.aburto@gmail.com winds induce an increase in primary production of 2% per latitudinal degree from 9.5°S
to 5°S. In some important coastal upwelling sites primary production is reduced. This is
Funding information
Conselho Nacional de Desenvolvimento due to a complex balance between nutrient availability, nutrient use efficiency, as well
gico (CNPq) - Bolsa de
Cientıfico e Tecnolo as eddy- and wind-driven factors. Mesoscale activity induces a net offshore transport
Produtividade em Pesquisa, Grant/Award
Number: 306971/2016-0; Coordenacß~ao de of inorganic nutrients, thus reducing primary production in the coastal upwelling region.
Aperfeicßoamento de Pessoal de Nıvel Wind mixing, in general disadvantageous for primary producers, leads to shorter resi-
Superior (CAPES) scholarship
dence times in the southern and central coastal zones. Overall, instead of a propor-
tional enhancement in primary production due to increased winds, the NHCS becomes
only 5% more productive (+5 mol C m2 year1), 10% less limited by nutrients and
15% less efficient due to eddy-driven effects. It is found that regions with a initial
strong nutrient limitation are more efficient in terms of nutrient assimilation which
makes them more resilient in face of the acceleration of the upwelling circulation.

KEYWORDS
Bakun’s hypothesis, build up of atmospheric CO2, eddy-driven effects, global warming-induced
wind intensification, latitudinal variability of primary production, Northern Humboldt Current
System, response of biological production

1 | INTRODUCTION 2017; Montes, Schneider, Colas, Blanke, & Echevin, 2011; Sanchez,
Calienes, & Zuta, 2000). This intrinsic variability makes the HCS resi-
The Humboldt Current System (HCS) sustains high levels of biologi- lient to natural climate variability (Bakun et al., 2015).
cal productivity due to yearlong upwelling-favorable southeast trade Previous studies show that Eastern Boundary Upwelling Systems
winds, which induce the vertical advection of cold, nutrient-rich and (EBUS’s) are very sensitive to climatic perturbations particularly in
oxygen-poor waters from relatively shallow depths (Strub, Combes, terms of deoxygenation, acidification and sea level rise (Bakun et al.,
Shillington, & Pizarro, 2013; Strub & Mesıas, 1998). The low-latitude 2015; Doney et al., 2012). In addition, the acceleration of the upwel-
of the HCS allows a strong upwelling circulation due to larger off- ling circulation, which is a direct consequence of the increase in upwel-
shore Ekman transport but also makes it very sensitive to equatorial ling-favorable winds, is potentially an important consequence of
~ o Southern Oscillation (ENSO)
oceanic perturbations, such as El Nin climate change in these highly productive areas. This hypothesis was
 n & Calil,
(Colas, Capet, McWilliams, & Shchepetkin, 2008; Mogollo initially proposed by Bakun (1990), who anticipated that the increase

Glob Change Biol. 2018;1–12. wileyonlinelibrary.com/journal/gcb © 2018 John Wiley & Sons Ltd | 1

2 | MOGOLLON AND R CALIL

in anthropogenic greenhouse gases, such as CO2, would lead to the the adjacent Eastern South Pacific (ESP). The horizontal resolution of
intensification of coastal upwelling in EBUS’s, particularly in the HCS the model grid is 1/12° (~9 km) with 30 vertical sigma levels
rrez et al., 2011). This is due to an increase in the land–sea ther-
(Gutie stretched toward the surface (approximately 18 levels over the con-
mal gradient that has been occurring as a consequence of global tinental slope). The topography is derived from the General Bathy-
warming in several EBUS’s (Sydeman et al., 2014). The mechanism metric Chart of the Oceans (GEBCO_2014 Grid, version 20150318,
works as follows: In normal conditions, a thermal low pressure cell is (www.gebco.net), which has a grid spacing of 30 arc sec (Weatherall
developed at the continental mass and a high pressure zone at the et al., 2015). It was smoothed in order to reduce pressure gradient
adjacent ocean. This is primarily due to the increased heating on land computation errors (Marchesiello, McWilliams, & Shchepetkin, 2003).
during the day and cooling at night when compared to the adjacent The hydrodynamical model is the Regional Oceanic Modeling
ocean. However, in a global warming scenario, the accumulation of System (ROMS-AGRIF) (Shchepetkin & McWilliams, 2005, 2009).
anthropogenic greenhouse gases would lead to an inhibition of the ROMS is a split-explicit, free surface oceanic model that solves the
nighttime radiative cooling, resulting in an enhancement of the day- primitive equations, based on the Boussinesq approximation and
time heating. This would induce to an intensification of the continental hydrostatic vertical momentum balance and is vertically discretized
thermal lows adjacent to the upwelling regions. in terrain-following curvilinear coordinates.
The consequences of this intensification would be (i) an Regional oceanic modeling system was forced at the boundaries
increased cross-shore atmospheric pressure gradient, (ii) the subse- with a monthly climatology constructed for the period 1999–2009
quent intensification of the alongshore geostrophic wind that drives from an eddy-permitting global ocean reanalysis, namely the CMCC
an offshore Ekman transport at the surface layer and (iii) the general Global Ocean Reanalysis System (C-GLORS) V.4 (Storto, Masina, &
acceleration of the coastal upwelling circulation. Bakun’s hypothesis Navarra, 2016), which was forced by ERA-Interim atmospheric vari-
tends to be a spring-summer phenomenon in the subtropics, while in ables with data assimilation of temperature and salinity profiles as
Peru it becomes a year-round phenomenon due to its proximity to well as sea level anomalies, sea ice concentration, sea surface tem-
the Equator (Bakun & Weeks, 2008). perature and mixed layer depth. It has a horizontal resolution of
Lachkar and Gruber (2013) explored the response of biological 0.25°.
production and air–sea CO2 fluxes to the upwelling intensification in At the surface, ROMS is forced with net shortwave radiation
two EBUS’s, namely the California and the Canary Current System extracted from j-OFURO V.3 (Kubota, Iwasaka, Kizu, Konda, & Kut-
using a NPZD-type ecosystem model. They found that a doubling of suwada, 2002) (obtained from http://dtsv.scc.u-tokai.ac.jp/j-ofuro/)
the wind stress magnitude doubles primary production in the south- as a monthly climatology at 0.25° resolution computed from 2002 to
ern California and central/northern Canary CS, while in the central/ 2007. For the computation of the kinematic surface net heat sensi-
northern California and southern Canary CS the increase in primary tivity to the sea surface temperature the air temperature, specific
production in response to the same wind variation is more modest. humidity and wind speed from j-OFURO database were used. The
In this work we assess the effect of intensifying winds on pri- evaporation and precipitation monthly climatology rates come from
mary production in the NHCS and the associated biological response the Woods Hole Oceanographic Institute’s objectively analyzed air–
in a climatological sense. Our methodology follows that of Lachkar sea heat fluxes (WHOI OAFlux V.3) available at http://oaflux.
and Gruber (2013) in that we simulate the upwelling intensification whoi.edu, and from the Tropical Rainfall Measuring Mission
trend using an idealized experiment which consists in varying the (TRMM_TMI V.3) available at www.remss.com/missions/tmi at 1°
wind forcing. The work is organized as follows. In Section Materials and 0.25° horizontal resolution, respectively, both computed from
and Methods, the physical–biogeochemical model and the numerical 1998 to 2006. The monthly wind stress climatology comes from the
experiments are described. The strengthening of the wind forcing is QuickSCAT satellite scatterometer (Liu, Tang, & Polito, 1998; Risien
detailed and the methods used for all the analyses are provided. In & Chelton, 2008) for the period 1999–2009. In order to account for
Results, we describe the effect of the induced wind trend on biologi- air-sea feedbacks, a relaxation to the sea surface temperature
cal production. We also assessed other eddy and wind-driven factors extracted from the Group for High Resolution Sea Surface Tempera-
that may negatively affect the latitudinal variability of primary pro- ture (GHRSST)—Operational Sea Surface Temperature and Sea Ice
duction. In Discussions we evaluate the Bakun’s hypothesis and the Analysis (OSTIA) of 5 km resolution (Donlon et al., 2012), was
overall biological response. An evaluation of our model experiment imposed following the parameterization of Barnier, Siefridt, and
against available observations may be found in the supporting infor- Marchesiello (1995). The sea surface salinity involved in the restoring
mation (from Appendices S1 to S4). terms was obtained from the World Ocean Atlas 2013 V.2 of 0.25°
horizontal resolution (Zweng et al., 2013).
A monthly climatology of the net heat fluxes, computed from
2 | MATERIALS AND METHODS
1999 to 2009, was extracted from The European Centre for Med-
ium-Range Weather Forecasts (ECMWF) Ocean Analysis System
2.1 | Model setup
ORAS3 (Balmaseda, Vidard, & Anderson, 2008). In spite of its
The domain spans the region between 70°W to 100°W and from lower spatial resolution of 1° when compared to COADS (0.5°) (Da
4°N to 26°S (Figure 1), encompassing the whole of the NHCS and Silva, Young, & Levitus, 1994), it does not lead to a warm bias at the

MOGOLLON AND R CALIL | 3

8 °N

Domain of simulation

0° Ecuador

Galápagos
Islands
200 Km
Punta Falsa
Domain of analysis of
8 °S
the wind intensification Chimbote Peru
Latitude

Callao
Pisco
Counterintuitive Zones
San Juan
16 °S

24 ° S
Chile

105 °W 100 °W 95 °W 90 °W 85 °W 80 °W 75 °W 70 °W 65 °W
Longitude

F I G U R E 1 Domain of simulation (black dashed line). The domain of analysis related with the wind intensification is shown as a green area
within the first 200 km nearshore. Red squares depict the counterintuitive zones

equatorial band and over the northern coastal zone. Nevertheless,


2.2 | Wind intensification
we retained the maximum and minimum values of the net heat
fluxes from the Modern Era Retrospective-Analysis for Research and In order to mimic Bakun’s hypothesis by comparing the climatologi-
Applications (MERRA2) (Bosilovich et al., 2015) during the cold cal wind forcing with a future scenario of strengthened winds that
phase of the climatological year in order to better reproduce the would be induced by global warming, we estimate the magnitude of
SST seasonal cycle. the strengthening of the upwelling-favorable trade winds in the HCS
Regional oceanic modeling system is coupled with the PISCES during a decade (1999–2009). This quantitative analysis was per-
biogeochemical model (Pelagic Interaction Scheme for Carbon and formed using the Cross Calibrated Multi Platform (CCMP) V.2
Ecosystem Studies) which simulates the marine biological productiv- (Wentz et al., 2015), from 1999 to 2009 (same period for the
ity, carbon and main nutrients cycling such as nitrate, phosphate, boundary conditions and QuickSCAT climatologies). CCMP combines
silicate and iron. The reader is referred to Aumont, Maier-Reimer, the remote sensing data with in situ measurements to produce a
Blain, and Monfray (2003); Aumont and Bopp (2006) for more consistent gap-free, long-term ocean surface wind product. CCMP
detailed information about the model. The nutrients and the oxy- (not shown) captures the cross-shore wind gradient in the coastal
gen fields were extracted from the World Ocean Atlas 2013 V.2 area.
(Garcia et al., 2014) for the upper ocean and from a global model Figure 2 shows a time series of the monthly zonal and meridional
output (NEMO-PISCES) for the ocean interior (Aumont et al., wind stress components from CCMP V2. Blue crosses are original
2003). The dissolved inorganic carbon and the total alkalinity were raw monthly data, the black continuous line denotes the filtered data
both obtained from Global Ocean Data Analysis Project (GLODAP (moving average seasonal low pass filter). The red line is least-
rez, 2016; Lauvset, Key, & Perez,
v2.2016b) (Key, Lauvset, & Pe squares fitted linear trend, and the green line is the long-term mean
2016). The surface forcing for PISCES includes atmospheric iron of each series. This analysis was performed over a 200 km-wide
dust input following Tegen and Fung (1995). The parameterization nearshore band along a nearly straight Peruvian coastline (orientation
of the biogeochemical model follows Echevin et al. (2014). Atmo- slope of about 120°) from 6°S to 15°S, as seen in Figure 1.
spheric pCO2 was set to a constant modern value of 397 ppm for The slope for each fitted line is 6.13 9 105 and
4 2 1
all experiments. It was estimated from the GLOBALVIEW-CO2 1.20 9 10 Nm month , for the zonal and meridional compo-
database (View-CO, 1979). nents, respectively (significance level p < .00001). The negative sign

4 | MOGOLLON AND R CALIL

of the zonal slope means that the zonal component of the wind is to temporal trends in ship sizes and associated anemometer heights,
strengthening westward, while the positive sign of the meridional trends in relative frequencies of measured winds and nonhomo-
slope means that this component is strengthening northward. During geneities in recording and archiving practices (Bakun et al., 2010).
almost a decade the zonal and meridional component were Within our spatio-temporal domain of analysis, we found that the
enhanced (with respect to their long-term mean) by, approximately, magnitude of the averaged meridional component of the wind stress
38.5% and 33.5% respectively. is actually more than twice (  2.3 times) the zonal one, which is
Narayan, Paul, Mulitza, and Schulz (2010) found a slope of reflected in a total wind intensification of 34.2% during a decade.
4 2 1
0.4 9 10 Nm month for the meridional wind component Since there exists low confidence in common trends in upwelling-
using COADS (four times coarser than CCMP horizontal resolution) favorable winds among EBUS’s (Narayan et al., 2010; Stocker, 2014),
from 1960 to 2000. This value is a third of the slope estimated in we may not discard the role of multidecadal climate variability in the
our period of analysis. The scale of the coastal intensification (tens observed trends. Thus, these findings, related with the wind intensi-
of kilometers) could partially explain this difference (Bakun, Field, fication percentages, should not be viewed as predictions, but rather
Redondo-Rodriguez, & Weeks, 2010). On the other hand, Bakun as reference values in order to perform a sensitivity study. Next, the
(1990) found a slope of 4.5 9 103 N m2 month1 from a 30 year numerical experiments and the rounded percentage values for the
linear trend (from mid 50s to mid 80s) of the alongshore wind stress idealized wind strengthening are detailed.
estimated between 4.5°S and 14.5°S within the coastal area. This
value is larger than the one used in this paper, basically because we
2.3 | Model experiment configuration
calculated the trend from the meridional rather than the alongshore
component of the wind stress, from 6°S to 15°S and over a wider A climatological simulation was run for 25 years starting from rest
area (200 km). The difference could also be explained by the differ- only with the hydrodynamical model. Statistical equilibrium is
ent period chosen for the analysis. Furthermore, uncertainty may reached after 10 years of model run (not shown). We kept the phys-
arise from an artificial long-term increasing trend on the winds due ical state of the end of year 25 to use it as the initial conditions in
which PISCES was coupled, starting with the biogeochemical clima-
tological conditions of January. After a spin-up of 5 years, the bio-
geochemical tracers reach a nearly repeating annual cycle. The
Meridional component
0.08 Ty : 1.20e–04 , 0.039 model solution used in this study is an average of the last 3 years of
simulation, from years 6–8, in order to remove the interannual vari-
0.07
ability. Hereafter, we refer to this solution as the “modern
0.06 winds”experiment. In order to reproduce the increase of upwelling-
favorable winds in the HCS, another simulation was performed in
0.05 which the zonal and meridional wind stress components were
increased by 40% and 35%, respectively (see Section Wind Intensifi-
0.04
)

cation). All other surface forcing and boundary conditions were kept
0.03 as in the modern winds experiment. ROMS/PISCES was run with
(

this perturbed wind product for 8 years. For consistency, we dis-


0.02
carded the first 5 years of simulation. Therefore, the model solution
raw for the “increased winds”experiment is an average from years 6–8 as
0.01 filtered
Zonal component trend in the case of the modern winds run.
0 Tx : –6.13e–05 , –0.017 mean

–0.01 2.4 | Production terms


–0.02 Primary Production (PP) is limited by the amount of photosyntheti-
cally active radiation (PAR), nutrient concentration, temperature,
–0.03
chlorophyll-to-carbon ratio and phytoplankton biomass in the follow-
–0.04 ing manner:
2 0 13
–0.05 ð ÞPAR
a Chl
C
PP ¼ Phytoplankton  4leppley @1  e leppley NL A5  NL (1)
99

00

01

02

03

04

05

06

07

08

09
19

20

20

20

20

20

20

20

20

20

20

The first factor that affects the variability of PP is the Nutrient


F I G U R E 2 Time series of the monthly zonal (Tx) and meridional Limitation term (0 ≤ NL ≤ 1), which is a nondimensional factor
(Ty) wind stress components from CCMP V.2 [N/m2]. The slope and parameterized using the Michaelis–Menten relation. The lower NL,
the intercept of the fitted lines are shown in the insets
the more limited by nutrients phytoplankton growth is. In PISCES,

MOGOLLON AND R CALIL | 5

phytoplankton is modeled by two compartments using the size of nitrogen-dependent, that is, constrained by a constant Redfield ratio.
the cells as a criterion. The main difference between them is the Since we are interested in quantifying the horizontal eddy advection
dependence on silicate by diatoms. Thus, NL is the minimum among of positive anomalies of TIN, we discarded the analysis of the nega-
all the limitation terms computed for all nutrients. The second factor tive ones. This procedure allows to associate positive values of hori-
is the specific growth rate (GR in day1), which is a term that gives zontal eddy fluxes with an onshore advection (positive u0 ) of positive
the combined influence of the physics on phytoplankton growth (the anomalies of TIN, and negative values with an offshore advection
term in brackets in Equation 1). (negative u0 ) of positive anomalies of TIN. In other words, we do not
leppley is the temperature-dependent maximum growth rate include the effect of an onshore flow of negative anomalies of TIN
(in day1). The term in parenthesis is a nondimensional factor which would result in a negative horizontal eddy flux. A more
which gives the combined influence of light (PAR) as well as the detailed inspection shows that this effect is much smaller than the
chlorophyll-to-carbon ratio, where a is the initial slope of the offshore advection of positive TIN anomalies and its removal does
photosynthesis-irradiance curve. not change our conclusions.
Pennington et al. (2006) defined the boundaries of the Peruvian
biogeochemical province as a coastal zone whose width is, approxi-
2.6 | Residence time
mately, of 250 km ranging from 4°S to 15°S. We estimated the
annual mean phytoplankton abundance between the simulated dia- We computed the theoretical Residence Time (hereafter RT) as the
toms and nanophytoplankton concentrations within these latitudes product of the layer depth scale and the habitat width, divided by
over the 200 km-wide nearshore area (extent of the productive the offshore volume transport. Here, the layer depth scale is given
zone). The large phytoplankton cells, that is, diatoms, represent by the latitude-dependant surface Ekman layer depth (dEkman) using a
almost 70% of the total phytoplankton community in our simula- coastal domain averaged eddy viscosity (parameterized as function
tions. This group is known to be dominant nearshore in Peruvian of the wind speed). The habitat width is determined by the first
waters (Iriarte & Gonzalez, 2004). This finding serves to weight the baroclinic Rossby radius of deformation (kRossby) extracted from
vertically averaged variables involved in the analyses (e.g., the limita- Chelton, Deszoeke, Schlax, El Naggar, and Siwertz (1998), and the
tions terms), related with both diatoms and nanophytoplankton. This offshore volume transport is estimated by the wind-driven cross-
is done in order to represent both phytoplankton types as an aver- shore Ekman transport (Mx). Hence, RT is the time required to
aged response of the phytoplankton community in the coastal area. entirely replace the volume of the upwelled water which travels a
The analyses (latitudinal distributions) were performed over the distance of one Rossby radius of deformation offshore.
300 km-wide nearshore band from 5°S to 17°S. This latitudinal
kRossby  dEkman
range was chosen following three criteria, namely, (i) to be away RT ¼ (2)
Mx
from the model boundaries, (ii) to focus on the most productive
region of the Peruvian biogeochemical province and (ii) to consider
only the latitudinal band in which the temporal quantitative analysis 3 | RESULTS
of the strengthening of the winds was performed (Section Wind
Intensification). The annual mean PP was vertically integrated within 3.1 | Effect of the induced wind trend on biological
the euphotic zone (ZEU), which is defined as the downward light production
penetration depth of 1% of the surface value of PAR. PP is the total
sum of the production associated to nano and diatoms. The NL and Figure 3 shows the annual mean latitudinal distribution of PP and

GR factors were both averaged in relation to their weights of phyto- the limitation factors (NL and GR). Blue continuous lines are the

plankton abundance and vertically averaged within the ZEU. results associated to the modern winds experiment while the red
dotted lines are the results for the increased winds experiment. The
percentage variations between both experiments are shown on the
2.5 | Eddy fluxes of nutrients right panel. On average, intensified winds induce more nutrient sup-

In order to quantify the eddy-driven leaking of the coastal inventory ply and an increase in PP, as seen in Figure 3a. This effect, however,

of nutrients, we performed a Reynolds decomposition of the hori- is not uniform in the coastal upwelling zone. An equatorward trend

zontal advection term, where the flux is decomposed into a mean of PP intensification of, approximately, 2% per latitudinal degree, is

and a time-varying (eddy)–induced flux, following the methodology found starting from Chimbote. The maximum relative PP increase of,

of Gruber et al. (2011). The eddy-induced horizontal flux is u0 :TIN0 , approximately, 15% occurs at southern Callao and southern San

where TIN 0
is the time-varying total inorganic nitrogen (in Juan. In general, PP is increased by, approximately, only
2 1
mmol N m3) and u0 is the time-varying cross-shore component of 5 mol C m year .

the velocity (in m/s). Both anomalies were calculated relative to the While PP increases in the whole domain in response to increased

annual mean. Only TIN (the sum of nitrate and ammonium) was winds, two important regions experience a decrease in PP, which is

considered in this analysis because the nitrogen pool undergoes evidenced by a negative relative percentage variation of around

nitrogen fixation and denitrification. In contrast, phosphate is 3%, as seen in Figure 3b. Note that another minimum occurs at

6 | MOGOLLON AND R CALIL

southern Chimbote, where the relative percentage variation is zero. coast, the lower NL values may result from an increased eddy-driven
We investigate in detail the response of these counterintuitive zones export of nutrients rather than a lack of nutrients. Next, we explore
in order to understand their resilience to increased upwelling-favor- this possibility.
able winds and assess their overall importance in the behavior of this Under increased winds GR decreases, on average, 15% (see Fig-
important EBUS to climate change. ure 3e,f). The maximum GR is found within the counterintuitive
As seen in Figure 3c,d, NL increases, on average, 10% within the zones, but since we demonstrated that these key regions exhibit rel-
domain (as NL increases, it yields a lower nutrient limitation status). ative strong NL status, GR is ultimately controlled by temperature
The maximum relative increase of around 20% occurs in the north- (limNL?0 GR  f(T)).
ernmost area and between Pisco–Callao. Note that the regions that The dependence of the Eppley relation (Equation 1) on tempera-
exhibit a counterintuitive response have the lowest NL values, mean- ture explains the negative variation in GR. Enhanced upwelling sup-
ing that phytoplankton growth is more limited by the availability of plies colder waters to the surface. Modeled temperature (averaged
nutrients in these areas than the rest of the domain. However, since over the ZEU) decreases 5% (about 0.85°C) under the increased
these regions are important upwelling centers along the Peruvian winds scenario. The correlation coefficient between temperature and

(a) (b)
50
15
Primary production
PP (mol C/m2/year )

% Relative variation
40 10 Counterintuitive
zones
5
30

0
20 modern increased
–5
–17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5 –17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5
(c) (d)
0.8
20
Nutrient limitation
% Relative variation

factor Counterintuitive
15
0.6 zones
NL

10

0.4 5
modern increased
0
–17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5 –17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5
(e) (f)
0
Counterintuitive zones
0.5 modern increased
% Relative variation

–10
GR (day –1)

0.4

0.3 –20
Specific growth
rate San Pisco Callao Chimbote Pta. Falsa
0.2 –30 Juan
–17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5 –17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5
Latitude Latitude

F I G U R E 3 Annual means latitudinal distributions of: (a) Primary production (PP) vertically integrated within ZEU. (c) Nutrient limitation (NL)
term vertically averaged within ZEU (e) Specific growth rate (GR) vertically averaged within ZEU. Panels b, d and f, are the percentage relative
variation of PP, NL and GR between modern (blue line) and increased (red dotted line) winds experiments, respectively. All the quantities were
averaged within the first 300 km nearshore band. Red dashed circles depict the counterintuitive zones

MOGOLLON AND R CALIL | 7

PP for the modern winds experiment is 0.73, while for increased important upwelling centers. Positive horizontal eddy fluxes are
winds it is 0.83. The simulated GR beyond our latitudinal limits of found over the shelf at Chiclayo (7°S) and Chimbote (9.4°S). How-
analysis showed larger rates near the Equator and southern San Juan ever, negative horizontal eddy fluxes are found from 70 km from the
(not shown). This is due to an interplay between warmer waters and coast in the upper 50 m for Chimbote. This demonstrates that an
well lit areas, rather than the temperature-dependent shift that offshore export of nitrogen (in the form of nitrate + ammonium) is
occurs along the Peruvian coast under the increased winds scenario. occurring, removing inorganic nutrients at a rate of
0.35 mmol N m2 s1. Callao and Pisco exhibit a similar pattern
with export rates of 0.20 and 0.25 mmol N m2 s1, respec-
3.2 | Eddy-driven effects on primary production
tively. We can now relate the level of mesoscale activity (red line in
If we assume that nutrient availability and nutrient use efficiency Figure 4a) with the four panels in Figure 4b. The lowest EKE values
(i.e., NL and GR) are equally important in controlling PP, we obtain a are found in the northern domain, where positive eddy-induced
5% deficit (since GR decreases 15% and NL increases 10%) that may fluxes of TIN are found. The largest EKE peak occurs between Chim-
be attributed to other physical factors under the increased winds bote and Callao at the central domain of analysis, where high rates
scenario. These factors negatively influence PP in the NHCS. This is of negative eddy-induced flux of TIN are found. In spite of Pisco
because the strengthening of the winds may also induce other non- having low EKE values, when compared to the whole coastal
linear effects on primary production, arising as by-products of the domain, it exhibits a nutrient-leaking pattern similar to Chimbote or
induced trend. In order to assess this influence and understand the Callao. This response could be associated with other eddy-driven
latitudinal variability of PP (see Figure S2), we focus in two mecha- factors, such as the change in orientation of the Peruvian coastline,
nisms. The first is the eddy-induced horizontal advection of nutrients bathymetric features or wind mixing energy production. The latter is
and the second is the local effect of wind mixing, as it induces a tur- assessed in the next section.
bulent habitat and deepens the mixed layer impacting PP.

3.4 | Mixing and residence times


3.3 | Offshore leaking of nutrients
The more equatorial location of the NHCS yields stronger Ekman
Our model results show that PP is higher in the northern coastal transport for similar wind speeds when compared to other EBUS’s.
domain. This is supported by satellite observations as seen in Fig- For example, Bakun and Weeks (2008) found that the wind speed in
ure S2, which shows a decrease in production rates from, approxi- the California CS (Cape Mendocino) needs to be twice the value
mately, 9°S to 14°S. This latitudinal band coincides with regions of observed along the Peruvian coast (Chimbote) in order to produce a
high Eddy Kinetic Energy (EKE), which is a measurement of the similar rate of upwelling. However, winds also induce turbulent mix-
mesoscale eddy activity. Figure 4a shows the annual mean latitudinal ing, which is roughly proportional to the third power of the wind
distribution of EKE (red line) derived from surface geostrophic cur- speed (Niiler, 1977). Large wind-induced turbulence is disadvanta-
rents computed from sea level anomalies from AVISO (1999–2009) geous for primary production as it tends to deepen the mixed layer
(http://www.aviso.altimetry.fr/duacs/). The satellite-derived EKE was and dilute phytoplankton in the water column.
horizontally averaged within the first 300 km-wide nearshore band. In order to assess the importance of the wind mixing on PP, we
A region with moderate EKE (10 cm2/s2) is found at the south- follow the methodology of Bakun and Weeks (2008), who defined
ernmost region between San Juan and Pisco (  16°S). A peak of the wind mixing index as the cube of the wind speed (w3). In Fig-
EKE is found in the central NHCS between Callao (12.1°S) and ure 4a, the black dotted line denotes the annual mean latitudinal dis-
Chimbote (9.4°S) with a mean EKE of, approximately, 20 cm2/s2. tribution of the wind mixing index. The southern domain has larger
The difference in the level of mesoscale activity in these two key values of w3 than the northern domain, with two relative maxima.
regions may induce different responses in PP. This hypothesis was The largest w3 is located between San Juan (15.7°S) and Pisco
initially proposed by Gruber et al. (2011), who suggested that eddies (14°S). This was expected due to the proximity of the eastern part
suppress production in highly productive EBUS’s. The mechanism with the South Pacific subtropical anticyclonic gyre. The other peak
consists in a leaking of the coastal nutrient inventory by horizontal is located north of Callao. Note that the latitudinal distribution of
advection of nutrients into the oligotrophic ocean, eventually leading the mixed layer depth is negatively correlated with the wind mixing.
n
to a reduction of biological production in the coastal zone. Mogollo In the southern domain, the mixed layer is almost twice deeper than
~ a,
and Calil (2017) demonstrated that the cold ENSO phase, La Nin in the northern domain, allowing a deeper penetration of wind-dri-
induces an increase in EKE in the southern NHCS, which leads to ven turbulence. Thus, the counterintuitive zones are disadvantageous
substantial offshore nitrogen advection (in the form of nitrate) at a for primary producers along the southern and central Peruvian coast.
rate of 0.5 kg N m2 day1 at the surface. Furthermore, under modern winds, the correlation between PP and
In order to quantify the efficiency of this mechanism, Figure 4b the wind mixing index is 0.86, which suggests that PP is strongly
shows vertical cross sections of the horizontal eddy fluxes of TIN (in modulated by changes in the wind forcing. Note that the counterin-
mmol N m2 s1) at 7°S, 9.4°S, 12.1°S and 14°S. These locations tuitive zones, where we found the largest levels of wind-driven mix-
were chosen in order to illustrate the latitudinal differences among ing when compared to the rest of the domain, exhibit lower primary

8 | MOGOLLON AND R CALIL

(a) (b)
w3 EKE RT MLD
Chiclayo 7°S
Chimbote 9.4°S
0 0

–50 –50

Depth (m)
–16 –100 –100
35 250 25
–18 –150 –150

Eddy kinetic energy (cm2/s2 )


Wind mixing index (m3/s3)
Residence time (days)

Mixed layer depth (m)


30
20 –20 –200 –200
200 200 150 100 50 200 150 100 50
25
–22
–0.2 –0.1 0 0.1 0.2
15
20 –24
150
–26 mmol N/m2/s
15 10 Pisco 14°S Callao 12.1°S
0 0
100 –28
10 –50 –50

Depth (m)
5 –30
–17 –16 –15 –14 –13 –12 –11 –10 –9 –8 –7 –6 –5
–100 –100
Latitude
–150 –150

–200 –200
200 150 100 50 200 150 100 50
Distance (km) Distance (km)

F I G U R E 4 (a) Annual mean latitudinal distribution of residence time (RT: blue dash-dotted line), wind mixing index (w3: black dotted line),
eddy kinetic energy (EKE: red continuous line) and mixed layer depth (MLD: green dashed line). (b) Cross-shore vertical sections of the
horizontal eddy-induced fluxes of total inorganic nitrogen (TIN) [mmol N m2 s1] at four different locations, as indicated. Positive (negative)
values mean eddy-induced onshore (offshore) advection of positive TIN anomalies

production rates. The key to understand the inverse relationship these features on the overall biogeochemistry of this EBUS deserves
between wind-driven mixing and PP, relies on the efficiency of nutri- further attention and should be addressed in a future study. Further-
ent recycling and the effective retainment within the productive more, as the largest w3 index is found at 14°S, the cluster released
upper layers. The lack of exposure to intense turbulent mixing would at this latitude exhibits a different vertical pattern, in which a sub-
be beneficial to primary producers in the NHCS. To have an insight stantial amount of offshore downwelling occurs, as evidenced by a
on these two mechanisms, which were initially proposed by Bakun sharply deepening of the trajectory in the southern drifters during
and Weeks (2008), we estimated the theoretical residence time the end of the month. This is in agreement with Gruber et al.
along the Peruvian coast (see Section Residence Time). Longer resi- (2011), who suggested that eddy-induced subduction and offshore
dence times are usually associated with lower turbulent activity, transport remove inorganic and organic nutrients from the nearshore
hence, we may use the residence time as a proxy of these two and potentially enrich the offshore region. Accordingly, there exists a
mechanisms at work. latitudinal trend in which the eddy-induced subduction effect weak-
Figure 4a shows the annual mean latitudinal distribution of RT in ens equatoward. As a consequence, the northern domain is less
days (blue dashed line). RT and w3 are inversely correlated (0.9). impacted by this mechanism when compared to the southern
The larger the wind mixing, the shorter the residence time. This domain. This helps to explain the latitudinal variability of PP (see Fig-
explains the higher PP rates in the northern portion of the domain, ure S2). The width of the shelf, which may be another factor that
as opposed to lower values in the central or southern domain (see impacts particle residence times and, hence, biological production,
Figure S2), where the counterintuitive zones are located. remains unexplored. However, based on our results, we speculate
A Lagrangian experiment helps illustrate how the mechanism that a wider continental shelf would lead to longer residence times,
works. It consisted in the deployment of four clusters of floats at thus favoring primary production.
four fixed locations and released at 55 m depth (average depth of
the coastal upwelling) within the productive coastal area during
3.5 | Temporal variability
August (stronger winds during austral winter). This parallel experi-
ment is detailed in the Supporting Information section The temporal variability for all variables previously discussed is sum-
(Appendix S5). The residence times of the modeled drifters agree marized in Figure 5a,b, where results are shown as monthly aver-
remarkably well with the theoretical one, although the latter repre- aged time series for the modern winds experiment. We decided not
sents the annual mean. Results show that the central portion of the to show the time series for the increased winds experiment because
NHCS is more affected by meso and submesoscale structures, such both solutions exhibit almost the same temporal pattern with some
as filaments, that effectively advect coastal waters into the open slight differences in the amplitude (PP and production terms). How-
ocean, which is evidenced by the scattered pattern in the central ever, we detailed these important differences as follows: Under the
NHCS (see Figure S5). Although characterizing the dynamics of the increased winds scenario, PP increases from winter to summer. Dur-
upwelling filaments is beyond the scope of this work, the impact of ing fall and early winter (AMJJ) no significant differences are

MOGOLLON AND R CALIL | 9

observed (only a decay in productivity between 1% and 2% under during late-winter/early-spring (not shown). As a consequence, no
increased winds). During summertime (JFM), PP increases almost significantly temporal variability is related to the well-lit zone. How-
13% (from 31 to 35 mol C m2 year1) under increased winds. A ever, the seasonal cycle of PAR (not shown) follows GR, with a maxi-
yearlong increase of, approximately, 10% of the NL factor is mum value of 260 W/m2 during summer and a minimum value of
obtained in both scenarios. Similarly, the GR is about 15% lower 125 W/m2 during winter.
under increased winds than in the modern winds experiment Our results are in agreement with Echevin et al. (2008), who
throughout the year. demonstrated that the surface chlorophyll concentration is highest
Wind mixing is larger and residence times are shorter during fall in austral summer and decreases during austral winter, in phase
and winter. The EKE is maximum in March, July and November. Our opposition with the coastal upwelling intensity. They demonstrated
results suggest that in the NHCS there exists a coupling between that the seasonal cycle of surface chlorophyll is mainly controlled
the eddy-driven export and wind mixing during late fall and early by the mixed layer depth, which deepens during winter and dilute
winter, that is, June and July, as evidenced in Figure 5b (black the surface chlorophyll concentration by entraining waters from
marked and red continuous lines). In contrast, during the rest of the below the mixed layer. In fact, our simulated mixed layer gets, on
year, they are both negatively correlated. This could partially explain average, 18% deeper along the year under the increased winds sce-
the so-called “Peruvian paradox” (Echevin, Aumont, Ledesma, & nario (not shown). Thus, we suggest that the global warming-
Flores, 2008; Pennington et al., 2006), which consists in an offset induced strengthening of the upwelling-favorable winds will ulti-
between the stronger upwelling period (austral winter) with the mately lead to a deepening of the mixed layer with strong impact
lower rates of primary production. This opposite relationship is clear on biogeochemical processes, such as entrainment/recycling rates
when we compare the time series of PP and w3 (from Figure 5a,b, of nutrients. This deepening would mitigate the effects of the
respectively). During winter, the NL status decreases (higher NL val- acceleration of the upwelling circulation by preventing abrupt
ues) due to larger nutrient concentrations in the euphotic zone. changes in production rates, thus yielding a more resilient biological
However, the enhanced coastal upwelling also leads to a decrease in response in the NHCS. The absence of the counteractive mecha-
the nutrient use efficiency partially due to a cooler environment nisms would probably result in a proportional increase in primary
(lower GR rates). Thus, the combined opposite effect of these two production to the wind perturbation, as found in other EBUS’s.
production terms partially control the seasonal cycle of PP in the One possible explanation for the resilience of the NHCS is that the
NHCS. In addition, the euphotic layer fluctuates around 40  3 m deepening of the mixed layer buffers an increase in primary pro-
depth, with a maximum depth during summer and a minimum one duction under increased winds.

(a)
0.6 0.6 –10
35 PP NL GR MLD
–15 Mixed layer depth (m)
Growth rate (day –1)

PP (mol C/m2/year)

0.5
Nutrient limitation

0.55
–20
30
0.4
0.5 –25

0.3 25 –30
0.45
–35
0.2
20 0.4 –40
J F M A M J J A S O N D

(b)
35 250 16
Eddy kinetic energy (cm 2 /s 2)

w3 EKE RT
Wind mixing index (m3/s 3)
Residence time (days)

30 14
200
12
25
10
150
20
8
15 100 6

10 4
J F M A M J J A S O N D
Months

F I G U R E 5 Monthly time series of: (a) Primary production, growth rate, nutrient limitation and mixed layer depth for the modern winds
experiment. (b) Residence time, wind mixing index and eddy kinetic energy. The residence time (RT) was averaged within the first 100 km
nearshore band. The rest of the quantities were averaged within the first 300 km nearshore band and between 5°S and 17°S

10 | MOGOLLON AND R CALIL

“predation hell”, because fish production usually depends more on


4 | DISCUSSION
recruitment success than on food availability. Furthermore, Salvatteci
et al. (2018) suggest that a long-term increase in coastal upwelling
4.1 | Bakun’s hypothesis
due to the warming climate, at least up to the start of the 21st cen-
In order to have an insight of the trends of atmospheric CO2 and tury, is favorable for fisheries productivity in the HCS. Our results
evaluate Bakun’s hypothesis along the Peruvian upwelling system, a show a moderate increase in primary production and suggest that its
complementary analysis was performed by using NOAA’s Car- spatiotemporal variability is a result of a delicate interplay between
bonTracker, version CT2016 (Peters et al., 2007), which is a global nutrient availability and nutrient use efficiency. This is reflected in
model of atmospheric carbon dioxide designed to keep track of the the opposite relationship between NL and GR which tend to balance
CO2 uptake and release at the Earth’s surface since 2000 at 1° reso- each other out.
lution. Detailed results can be found in the Supplementary Informa- Previous studies have suggested that PP will decrease with the
tion (Appendix S6). Overall, in agreement with Bakun’s hypothesis, a increased stratification associated with future surface warming, given
warming trend and a weakening of the surface pressure were the inverse relationship between ocean temperatures and ecosystem
obtained. The air temperature increased 0.45°C associated with a productivity in observations (Behrenfeld et al., 2006; Sarmiento
slight decay on air surface pressure over the decade of analysis et al., 2004). For instance, Brochier et al. (2013) found that the
(2000–2009). A positive trend of atmospheric pCO2 is obtained expected surface global warming will increase stratification, with
because the source terms (fossil fuels and fires) are larger than the strong impacts in the circulation and mesoscale activity in the Hum-
sink terms (biosphere and the ocean). Using Least Squares, a slope boldt CS. Both physical changes would in turn negatively affect PP
of 0.163 ppm per month is found which translates to an increase of, due to the impact on nutrients and oxygen contents and due to the
approximately, 2 ppm of pCO2 in the atmosphere per year at the strong reduction of the offshore extent of the productive area. The
coastal Peruvian upwelling region mainly due to anthropogenic latter appears that is not fully compensated by the increase in reten-
emissions. tion rates on the continental shelf. As a consequence, they predict a
future reduction in small pelagic fish recruitment in the Humboldt
CS. In sharp contrast, Rykaczewski and Dunne (2010) obtained a
4.2 | Biological response
counterintuitive increased nutrient supply (attributed to nitrate
Previous studies in other EBUS’s found that regions with low initial enrichment of deep source waters) which results from long-term
nutrient limitation status (high NL values) tend to be more resilient warming and stratification in the North Pacific California Current
to wind changes, while regions with high initial nutrient limitation Ecosystem, demonstrating that local responses to changing climate
(low NL values), increase their nutrient limitation status and hence can differ significantly from the general global response. In our work,
PP substantially (Lachkar & Gruber, 2013). However, this general which consisted in a single factor tuning methodology (modern and
behavior does not explain the biological response in the NHCS. We increased winds scenarios), the surface warming effect (and associ-
found that regions with low initial NL values (such as the counterin- ated increased water column stratification) is neglected because we
tuitive zones) do not significantly increase their limitation status kept the same heat fluxes in both control and perturbed simulation.
under an increased winds scenario. Moreover, the northern domain, The combined effect of wind intensification and stratification in the
which is characterized by higher NL or weak limitation status, sub- Humboldt CS remains open to debate, and we speculate that the
stantially increased its limitation status under increased winds sce- overall biological response may depend on the specific characteris-
nario. As a consequence, our findings suggest that the tics of each ecosystem. Our study focuses specifically on the effects
counterintuitive zones are the most resilient (in terms of the biologi- of the increase of upwelling-favorable winds that may be induced by
cal response) to the global warming-induced upwelling intensification global warming. Other modeling studies are necessary to address not
trend. only the isolated impacts of winds or heat fluxes on PP but also on
Overall, as the highest GR are found in the southern and central how the inherently coupled ocean–atmosphere system and its asso-
portion of the domain, it appears that the nutrients are being more ciated feedbacks will impact ocean biogeochemistry.
efficiently used in these regions (particularly within the counterintu-
itive zones) than over the northern domain. Despite their relatively
ACKNOWLEDGEMENTS
lower rates of primary production, it agrees with Bakun et al. (2015).
They suggest that the increased intensity of the upwelling-favorable n acknowledges support from the Coordenacßa~o de Aper-
R. Mogollo
winds could lead to less phytoplankton production within the pri- feicßoamento de Pessoal de Nıvel Superior (CAPES) scholarship.
mary upwelling zone due to deeper wind-driven mixing and P.H.R. Calil acknowledges support from the Conselho Nacional de
increased light limitation. However, this moderate primary produc-  gico (CNPq) - Bolsa de Produ-
Desenvolvimento Cientıfico e Tecnolo
tion (which would in turn sustain zooplankton and upper trophic tividade em Pesquisa (Process: 306971/2016-0). We thank M. Bosi-
level communities) does not necessarily lead to smaller fish produc- lovich from the MERRA team for technical support. The ARGO data
tion, according to Bakun and Weeks (2008), who emphasizes the were collected and made freely available by the International Argo
notion that in the ocean, “food heaven” generally coincides with Program and the national programs that contribute to it. The Argo

MOGOLLON AND R CALIL | 11

Program is part of the Global Ocean Observing System. Argo DOI Chelton, D. B., Deszoeke, R. A., Schlax, M. G., El Naggar, K., & Siwertz,
(https://doi.org/10.17882/42182). CCMP Version-2.0 vector wind N. (1998). Geographical variability of the first baroclinic rossby radius
of deformation. Journal of Physical Oceanography, 28(3), 433–460.
analyses are produced by Remote Sensing Systems (www.remss.c
https://doi.org/10.1175/1520-0485(1998)028&lt;0433:
om). The altimeter products were produced by Ssalto/Duacs and dis- GVOTFB&gt;2.0.CO;2
tributed by Aviso, with support from Cnes. CarbonTracker CT2016 Colas, F., Capet, X., McWilliams, J., & Shchepetkin, A. (2008). 1997–1998
results provided by NOAA ESRL, Boulder, Colorado, USA. El Nin ~o off Peru: A numerical study. Progress in Oceanography, 79(2–
4), 138–155. https://doi.org/10.1016/j.pocean.2008.10.015
Da Silva, A., Young, C., & Levitus, S. (1994). Atlas of surface marine data
CONFLICT OF INTEREST 1994, vol. 1, algorithms and procedures, noaa atlas nesdis 6. US
Department of Commerce, NOAA, NESDIS, USA. page 74.
The authors have no conflict of interest. Doney, S. C., Ruckelshaus, M., Duffy, J. E., Barry, J. P., Chan, F., English,
C. A., . . . Talley, L. D. (2012). Climate change impacts on marine
ecosystems. Annual Review of Marine Science, 4(1), 11–37. PMID:
ORCID 22457967. https://doi.org/10.1146/annurev-marine-041911-111611
Donlon, C. J., Martin, M., Stark, J., Roberts-Jones, J., Fiedler, E., & Wim-
n
Rodrigo Mogollo http://orcid.org/0000-0002-8049-9520 mer, W. (2012). The operational sea surface temperature and sea ice
analysis (ostia) system. Remote Sensing of Environment, 116, 140–158.
Paulo H. R. Calil https://orcid.org/0000-0001-6361-1747
https://doi.org/10.1016/j.rse.2010.10.017
Echevin, V., Albert, A., Le vy, M., Graco, M., Aumont, O., Pietri, A., & Gar-
ric, G. (2014). Intraseasonal variability of nearshore productivity in
REFERENCES the northern humboldt current system: The role of coastal trapped
waves. Continental Shelf Research, 73, 14–30. https://doi.org/
Aumont, O., & Bopp, L. (2006). Globalizing results from ocean in situ iron 10.1016/j.csr.2013.11.015
fertilization studies. Global Biogeochemical Cycles, 20, GB2017. Echevin, V., Aumont, O., Ledesma, J., & Flores, G. (2008). The seasonal
https://doi.org/10.1029/2005GB002591 cycle of surface chlorophyll in the peruvian upwelling system: A mod-
Aumont, O., Maier-Reimer, E., Blain, S., & Monfray, P. (2003). An ecosys- elling study. Progress in Oceanography, 79(2), 167–176. https://doi.
tem model of the global ocean including Fe, Si, P colimitations. Global org/10.1016/j.pocean.2008.10.026
Biogeochemical Cycles, 17, 1060. https://doi.org/10.1029/2001GB Garcia, H., Locarnini, R., Boyer, T., Antonov, J., Baranova, O., Zweng, M.,
001745 . . . Johnson, D. (2014). World ocean atlas 2013, volume 3: Dissolved
Bakun, A. (1990). Global climate change and intensification of coastal oxygen, apparent oxygen utilization, and oxygen saturation. NOAA
ocean upwelling. Science, 247(4939), 198–201. https://doi.org/10. Atlas NESDIS, 75, 27.
1126/science.247.4939.198 Gruber, N., Lachkar, Z., Frenzel, H., Marchesiello, P., Mu €nnich, M., McWil-
Bakun, A., Black, B. A., Bograd, S. J., Garcia-Reyes, M., Miller, A. J., liams, J. C., . . . Plattner, G.-K. (2011). Eddy-induced reduction of bio-
Rykaczewski, R. R., & Sydeman, W. J. (2015). Anticipated effects of logical production in eastern boundary upwelling systems. Nature
climate change on coastal upwelling ecosystems. Current Climate geoscience, 4(11), 787–792. https://doi.org/10.1038/ngeo1273
Change Reports, 1(2), 85–93. https://doi.org/10.1007/s40641-015- Gutie rrez, D., Bouloubassi, I., Sifeddine, A., Purca, S., Goubanova, K.,
0008-4 Graco, M., . . . Salvatteci, R. (2011). Coastal cooling and increased
Bakun, A., Field, D. B., Redondo-Rodriguez, A., & Weeks, S. J. (2010). productivity in the main upwelling zone off peru since the mid-twen-
Greenhouse gas, upwelling favorable winds, and the future of coastal tieth century. Geophysical Research Letters, 38, L07603. https://doi.
ocean upwelling ecosystems. Global Change Biology, 16(4), 1213– org/10.1029/2010GL046324
1228. https://doi.org/10.1111/j.1365-2486.2009.02094.x Iriarte, J. L., & Gonzalez, H. E. (2004). Phytoplankton size structure during
Bakun, A., & Weeks, S. J. (2008). The marine ecosystem off peru: What and after the 1997/98 el nin ~o in a coastal upwelling area of the
are the secrets of its fishery productivity and what might its future northern humboldt current system. Marine ecology progress series,
hold? Progress in Oceanography, 79(2), 290–299. https://doi.org/10. 269, 83–90. https://doi.org/10.3354/meps269083
1016/j.pocean.2008.10.027 Key, R. M., Lauvset, S. K., & Pe rez, F. F. (2016). The global ocean data
Balmaseda, M. A., Vidard, A., & Anderson, D. L. (2008). The ecmwf ocean analysis project version 2 (glodapv2)-an internally consistent data
analysis system: Ora-s3. Monthly Weather Review, 136(8), 3018–3034. product for the world ocean. Earth System Science Data, 8(2), 297.
https://doi.org/10.1175/2008MWR2433.1 Kubota, M., Iwasaka, N., Kizu, S., Konda, M., & Kutsuwada, K. (2002).
Barnier, B., Siefridt, L., & Marchesiello, P. (1995). Thermal forcing for a Japanese ocean flux data sets with use of remote sensing observa-
global ocean circulation model using a three-year climatology of tions (j-ofuro). Journal of Oceanography, 58(1), 213–225. https://doi.
ecmwf analyses. Journal of Marine Systems, 6(4), 363–380. https:// org/10.1023/A:1015845321836
doi.org/10.1016/0924-7963(94)00034-9 Lachkar, Z., & Gruber, N. (2013). Response of biological production and
Behrenfeld, M. J., O’Malley, R. T., Siegel, D. A., McClain, C. R., Sarmiento, air–sea co2 fluxes to upwelling intensification in the california and
J. L., Feldman, G. C., . . . Boss, E. S. (2006). Climate-driven trends in canary current systems. Journal of Marine Systems, 109, 149–160.
contemporary ocean productivity. Nature, 444(7120), 752. https:// https://doi.org/10.1016/j.jmarsys.2012.04.003
doi.org/10.1038/nature05317 Lauvset, S. K., Key, R. M., & Perez, F. F. (2016). A new global interior
Bosilovich, M. G., Akella, S., Coy, L., Cullather, R., Draper, C., Gelaro, R., ocean mapped climatology: The 1x1 glodap version 2. Earth System
. . . Suarez, M. (2015). Merra-2: Initial evaluation of the climate. Series Science Data, 8(2), 325.
on Global Modeling and Data Assimilation, NASA/TM, 104606. Liu, W. T., Tang, W., & Polito, P. S. (1998). Nasa scatterometer provides
Brochier, T., Echevin, V., Tam, J., Chaigneau, A., Goubanova, K., & Ber- global ocean-surface wind fields with more structures than numerical
trand, A. (2013). Climate change scenarios experiments predict a weather prediction. Geophysical Research Letters, 25(6), 761–764.
future reduction in small pelagic fish recruitment in the humboldt https://doi.org/10.1029/98GL00544
current system. Global change biology, 19(6), 1841–1853. https://doi. Marchesiello, P., McWilliams, J. C., & Shchepetkin, A. (2003). Equilibrium
org/10.1111/gcb.12184 structure and dynamics of the california current system. Journal of

12 | MOGOLLON AND R CALIL

Physical Oceanography, 33(4), 753–783. https://doi.org/10.1175/ Stocker, T. (2014). Climate change 2013: the physical science basis:
1520-0485(2003)33&lt;753:ESADOT&gt;2.0.CO;2 Working Group I contribution to the Fifth assessment report of the
Mogollo  n, R., & Calil, P. H. R. (2017). On the effects of ENSO on ocean Intergovernmental Panel on Climate Change. Cambridge University
biogeochemistry in the Northern Humboldt Current System (NHCS): Press.
A modeling study. Journal of Marine Systems, 172, 137–159. https:// Storto, A., Masina, S., & Navarra, A. (2016). Evaluation of the cmcc eddy-
doi.org/10.1016/j.jmarsys.2017.03.011 permitting global ocean physical reanalysis system (c-glors, 1982–
Montes, I., Schneider, W., Colas, F., Blanke, B., & Echevin, V. (2011). 2012) and its assimilation components. Quarterly Journal of the Royal
Subsurface connections in the eastern tropical Pacific during La Meteorological Society, 142(695), 738–758. https://doi.org/10.1002/
Nin~a 1999–2001 and El Nin ~o 2002–2003. Journal of Geophysical qj.2673
Research, 116(C12), C12022. https://doi.org/10.1029/2011JC0 Strub, P. T., Combes, V., Shillington, F. A., & Pizarro, O. (2013). Currents
07624 and processes along the eastern boundaries. Ocean Circulation and
Narayan, N., Paul, A., Mulitza, S., & Schulz, M. (2010). Trends in coastal Climate: A 21st Century Perspective. pp. 339–384.
upwelling intensity during the late 20th century. Ocean Science, 6(3), Strub, P. T., & Mesıas, J. M. (1998). Coastal ocean circulation off western
815. https://doi.org/10.5194/os-6-815-2010 South America. In A. R. Robinson, & K. H. Brink (Eds.), The sea (Vol.
Niiler, P. (1977) One-dimensional models of the upper ocean, modelling 11, pp. 273–313). NewYork: John Wiley.
and prediction of the upper layers of the ocean. EB Kraus. pp. 143– Sydeman, W., Garcıa-Reyes, M., Schoeman, D., Rykaczewski, R., Thomp-
172. son, S., Black, B., & Bograd, S. (2014). Climate change and wind
Pennington, J. T., Mahoney, K. L., Kuwahara, V. S., Kolber, D. D., intensification in coastal upwelling ecosystems. Science, 345(6192),
Calienes, R., & Chavez, F. P. (2006). Primary production in the east- 77–80. https://doi.org/10.1126/science.1251635
ern tropical pacific: A review. Progress in Oceanography, 69(2), 285– Tegen, I., & Fung, I. (1995). Contribution to the atmospheric mineral
317. https://doi.org/10.1016/j.pocean.2006.03.012 aerosol load from land surface modification. Journal of Geophysical
Peters, W., Jacobson, A. R., Sweeney, C., Andrews, A. E., Conway, T. J., Research: Atmospheres, 100(D9), 18707–18726. https://doi.org/10.
Masarie, K., . . . Worthy, D. E. (2007). An atmospheric perspective on 1029/95JD02051
north american carbon dioxide exchange: Carbontracker. Proceedings View-CO, G. (1979). Cooperative global atmospheric data integration
of the National Academy of Sciences, 104(48), 18925–18930. https:// project, updated annually. Multi-laboratory compilation of synchro-
doi.org/10.1073/pnas.0708986104 nized and gap-filled atmospheric carbon dioxide records for the per-
Risien, C. M., & Chelton, D. B. (2008). A global climatology of surface iod, 2012:2013.
wind and wind stress fields from eight years of quikscat scatterome- Weatherall, P., Marks, K., Jakobsson, M., Schmitt, T., Tani, S., Arndt, J. E.,
ter data. Journal of Physical Oceanography, 38(11), 2379–2413. . . . Wigley, R. (2015). A new digital bathymetric model of the world’s
https://doi.org/10.1175/2008JPO3881.1 oceans. Earth and Space Science, 2(8), 331–345. https://doi.org/10.
Rykaczewski, R. R., & Dunne, J. P. (2010). Enhanced nutrient supply to 1002/2015EA000107
the california current ecosystem with global warming and increased Wentz, F., Scott, J., Hoffman, R., Leidner, M., Atlas, R., & Ardizzone, J.
stratification in an earth system model. Geophysical Research Letters, (2015). Remote sensing systems cross-calibrated multi-platform
37, L21606. https://doi.org/10.1029/2010GL045019 (ccmp) 6-hourly ocean vector wind analysis product on 0.25 deg grid,
Salvatteci, R., Field, D., Gutie rrez, D., Baumgartner, T., Ferreira, V., version 2.0. Remote Sensing Systems, Santa Rosa, CA Google Scho-
Ortlieb, L., . . . Bertrand, A. (2018). Multifarious anchovy and sardine lar.
regimes in the humboldt current system during the last 150 years. Zweng, M., Reagan, J., Antonov, J., Locarnini, R., Mishonov, A., Boyer, T.,
Global Change Biology, 24(3), 1055–1068. https://doi.org/10.1111/ & Biddle, M. M. (2013). World ocean atlas 2013, noaa atlas nesdis
gcb.13991 74 vol. 2: Salinity. NOAA.
Sanchez, G., Calienes, R., & Zuta, S. (2000). The 1997-98 el nin ~o and its
effects on the coastal marine ecosystem off peru. Reports of Califor-
nia Cooperative Oceanic Fisheries Investigations, 41, 62–86.
Sarmiento, J. L., Slater, R., Barber, R., Bopp, L., Doney, S. C., Hirst, A., . . . SUPPORTING INFORMATION
Soldatov, V. (2004). Response of ocean ecosystems to climate warm-
Additional supporting information may be found online in the
ing. Global Biogeochemical Cycles, 18, GB3003. https://doi.org/10.
1029/2003GB002134 Supporting Information section at the end of the article.
Shchepetkin, A. F., & McWilliams, J. C. (2005). The regional oceanic mod-
eling system (ROMS): A split-explicit, free-surface, topography-follow-
ing-coordinate oceanic model. Ocean Modelling, 9(4), 347–404.
https://doi.org/10.1016/j.ocemod.2004.08.002 n R, R Calil PH.
How to cite this article: Mogollo
Shchepetkin, A. F., & McWilliams, J. C. (2009). Correction and commen- Counterintuitive effects of global warming-induced wind
tary for “Ocean forecasting in terrain-following coordinates: Formula- patterns on primary production in the Northern Humboldt
tion and skill assessment of the regional ocean modeling system” by
Current System. Glob Change Biol. 2018;00:1–12.
Haidvogel et al., J. Comp. Phys. 227, pp. 3595–3624. Journal of Com-
putational Physics, 228(24), 8985–9000. https://doi.org/10.1016/j.jcp. https://doi.org/10.1111/gcb.14171
2009.09.002

You might also like