You are on page 1of 13

RESEARCH ARTICLE Role of Atlantification in Enhanced Primary Productivity in

10.1029/2023EF003709
the Barents Sea
Key Points:
Kyung-Min Noh1,2 , Ji-Hoon Oh1 , Hyung-Gyu Lim3,4,5, Hajoon Song2 , and Jong-Seong Kug1,6
• T he extents of Atlantic water expand
to the central Arctic basin due to 1
Division of Environmental Science and Engineering, Pohang University of Science and Technology (POSTECH), Pohang,
increased heat transport through
Republic of Korea, 2Department of Atmospheric Sciences, Yonsei University, Seoul, Republic of Korea, 3Princeton
the Atlantic gateways in a warming
climate University/Atmospheric and Oceanic Sciences Program, Princeton, NJ, USA, 4Scripps Institution of Oceanography,
• The intensified Atlantification University of California San Diego, San Diego, CA, USA, 5Korea Institute Ocean Science and Technology, Busan, Republic
drives the enhancement of spring of Korea, 6Institute for Convergence Research and Education in Advanced Technology, Yonsei University, Seoul, South Korea
productivity in the Barents Sea
• Notable divergence in physical
manifestations of Atlantification
among CMIP models engenders
Abstract Recent changes in the Arctic sea-ice are strongly influenced by the recent increase in heat
significant challenges in projecting transport from vigorous Atlantic inflows, so-called Atlantification. This Atlantification can induce physical
future Arctic productivity and ecological changes near the Atlantic gateway. Here, we used the observational data sets and 26 Earth
system models to estimate Atlantic water intrusion, and firstly suggest the impact of Atlantification on marine
Supporting Information: productivity in the Barents Sea in a warming climate, especially on boreal spring. In a warming climate, the
Supporting Information may be found in heat transport across the Barents Sea Opening (BSO) is projected to be enhanced (45.5 ± 34.9 TW) by the
the online version of this article. end of the 21st century compared to the present climate. This poleward intrusion of the Atlantic water is likely
to increase productivity with the largest increase in spring (70%). In a warming climate, the productivity is
Correspondence to: enhanced by Atlantification-induced changes in physical states—ocean temperature, circulations, stratification,
J.-S. Kug, and sea-ice. Based on inter-model analyses, we estimated that the Atlantification can explain approximately
jskug1@gmail.com
26% of the productivity changes in the Barents Sea. Thus, Atlantification is critical for future changes in
biological productivity and physical states over the Arctic Ocean.
Citation:
Noh, K.-M., Oh, J.-H., Lim, H.-G.,
Song, H., & Kug, J.-S. (2024). Role of
Plain Language Summary Human-induced greenhouse gases are causing the sea-ice in the
Atlantification in enhanced primary Arctic Ocean to decrease. This is making the edges of the sea-ice retreat poleward to the central Arctic. The
productivity in the Barents Sea. Earth's Atlantic water, which is warm, salty, and nutrient-rich, is also expanding northwards. This is causing the Arctic
Future, 12, e2023EF003709. https://doi.
water to become more like Atlantic water, which is called “Atlantification.” Atlantification-induced physical
org/10.1029/2023EF003709
manifestations and their future changes have been relatively well understood, while influence of the marine
Received 24 APR 2023 productivity still has large uncertainties. In this study, we analyzed the 26 state-of-the-art Earth system models
Accepted 14 DEC 2023 (ESMs) and estimated the future changes caused by the Atlantification, and related productivity changes based
on diverse responses of model projections in the Barents Sea where the largest productivity change exists over
Author Contributions: the Arctic Ocean. We find that the Atlantification-induced changes in temperature, sea-ice, and vertical mixing,
Conceptualization: Kyung-Min Noh, can enhance the level of productivity in the Barents Sea, especially in boreal spring. Our results suggest that
Ji-Hoon Oh, Hyung-Gyu Lim, Hajoon
understanding the interactions between the Atlantic water and Arctic systems such as ocean, cryosphere, and
Song, Jong-Seong Kug
Data curation: Kyung-Min Noh biology is critical to projecting future Arctic productivity.
Formal analysis: Kyung-Min Noh
Funding acquisition: Jong-Seong Kug
Investigation: Kyung-Min Noh, Ji-Hoon 1. Introduction
Oh, Hyung-Gyu Lim
Methodology: Kyung-Min Noh, The Arctic has warmed rapidly in recent decades (Meredith et al., 2019), and the rate of warming in the Arctic is
Ji-Hoon Oh
expected to be more than double the global average warming rate (Screen & Simmonds, 2010; Serreze et al., 2009).
This rapid Arctic warming manifests the other components of the Arctic Ocean from physical states—ocean
© 2024 Pohang University of Science temperature, circulation, stratification, and sea ice—to marine ecosystems (Ardyna & Arrigo, 2020). Specifi-
and Technology. Earth's Future published
by Wiley Periodicals LLC on behalf of cally, the sea-ice extent has rapidly retreated from the North Atlantic to the central Arctic Ocean; thus, the Barents
American Geophysical Union. Sea is expected to be ice-free in 2060–2080 (Onarheim & Arthun, 2017). Observations and climate models
This is an open access article under revealed that the Barents Sea has the most pronounced warming and sea-ice reduction (Jansen et al., 2020). This
the terms of the Creative Commons
Attribution-NonCommercial-NoDerivs sea-ice reduction triggers the albedo feedback, which contributes to the amplified warming in the Arctic (Pithan
License, which permits use and & Mauritsen, 2014) and higher sensitivity to the warming in the Barents Sea than in the other Arctic sub-regions
distribution in any medium, provided the (Smedsrud et al., 2013). The influence of warming in the Barents Sea is not only limited to the Arctic climate
original work is properly cited, the use is
non-commercial and no modifications or but expanded to the mid-latitude weather, implying the importance of the Barents Sea (Cohen et al., 2014; Jang
adaptations are made. et al., 2021; Kim et al., 2022; Kug et al., 2015).

NOH ET AL. 1 of 13
Earth’s Future 10.1029/2023EF003709

Project Administration: Hajoon Song, Although the atmospheric process affects the temperature changes in the Barents Sea (Asbjornsen et al., 2020;
Jong-Seong Kug
Woods & Caballero, 2016), the Barents Sea is mostly influenced by vigorous inflows of the Atlantic water in
Resources: Jong-Seong Kug
Software: Kyung-Min Noh the Arctic Ocean (Docquier & Koenigk, 2021), characterized by the warm (>0℃), salty, and nutrient-abundant
Supervision: Jong-Seong Kug water in the subsurface layer (Aagaard et al., 1981). Atlantic water, supplied to the Arctic Ocean through Atlantic
Visualization: Kyung-Min Noh, Ji-Hoon
gateways, gradually expands to the central Arctic Ocean and the Arctic Ocean becomes more like the Atlantic
Oh
Writing – original draft: Kyung-Min Ocean in terms of its physical and biological characteristics (Ingvaldsen et al., 2021), interpreted as “Atlantifi-
Noh cation.” The recent increased heat transport from the Atlantic gateways accelerates the Atlantification (Arthun
Writing – review & editing: Kyung-
et al., 2012; Tsubouchi et al., 2021). Previous studies have suggested that heat transport across the Barents Sea
Min Noh, Ji-Hoon Oh, Hyung-Gyu Lim,
Hajoon Song, Jong-Seong Kug Opening (BSO) plays an important role in the sea-ice decline in the Barents Sea (Li et al., 2017) and is a good
indicator of Arctic winter sea-ice variability (Arthun et al., 2019). The intrusion of the Atlantic water also modi-
fies the other physical states in the Barents Sea, such as increased ocean temperature and enhanced vertical
mixing (Shu et al., 2021). Evidence of the Atlantification is even observed in the Eurasian Basin (Polyakov
et al., 2017). In addition, these changes are accelerated by the positive feedback between vertical mixing and
freshwater fluxes in the upper ocean (Ivanov et al., 2016; Lind et al., 2018). Ocean heat coming from the North
Atlantic is released into the atmosphere at the Barents Sea, and these heat releases are projected to expand from
the northern Barents Sea to the Arctic Basin in a warming climate (Muilwijk et al., 2023; Shu et al., 2021).

These physical manifestations caused by enhanced Atlantification affect ecological changes such as increased
primary production and northward expansion of boreal species (Csapo et al., 2021; Ingvaldsen et al., 2021;
Polyakov et al., 2020). Interestingly, the largest increases in marine productivity are found in the northern Barents
and Kara seas (Arrigo & van Dijken, 2015; Lewis et al., 2020; Slagstad et al., 2015), which has continuously
increased in recent decades (Arrigo & van Dijken, 2015; Pabi et al., 2008). This increased productivity in the
Barents Sea is projected to persist in a warming climate (Bopp et al., 2013; Kwiatkowski et al., 2020; Vancoppenolle
et al., 2013) despite the declined phytoplankton biomass (Noh, Lim, et al., 2023). The Atlantification-induced
sea-ice reduction allows more sunlight to the upper ocean, resulting in an earlier start of the peak productiv-
ity (Kahru et al., 2016; Wassmann & Reigstad, 2011). Moreover, the weakened stratification and advection
of nutrient-rich Atlantic inflow enhance nutrient supply in the Barents Sea (Fransner et al., 2023; Randelhoff
et al., 2018; Vernet et al., 2020). These biogeochemical changes result in the poleward expansion of phytoplank-
ton with increased blooming in the Barents Sea, which is caused by increased temperature and stronger Atlantic
currents (Neukermans et al., 2018; Oziel et al., 2020). Although a previous study projected enhanced produc-
tivity in the Arctic Ocean in the Atlantic water pathway (Polyakov et al., 2020), detailed mechanisms of how
the Atlantification affects productivity are still uncertain. Additionally, the uncertainty in the changes in Arctic
productivity is still high (Cabré et al., 2014; Tagliabue et al., 2021), so we need to investigate the impacts of the
Atlantification on the productivity in the Arctic Sea to provide a reliable future projection of Arctic productivity.

In this study, we analyzed future Atlantification in response to the high-emission scenario and suggested the
mechanism for how the Atlantification affects productivity in the Barents Sea, especially on the boreal spring.
Based on the analyses of the inter-model spread, we estimated diverse responses of the Atlantification on the
changes in physical states and ocean productivity using both CMIP archives and observations. The relationship
between the projected Atlantification and productivity changes in the Barents Sea is firstly presented. Finally, a
summary and discussion about the biological impact of the Atlantification are provided.

2. Materials and Methods


2.1. CMIP5 and CMIP6 Data Sets

To analyze the effects of the Atlantification on the Arctic environments, we used 26 Earth system model (ESM)
outputs participating in the Coupled Model Intercomparison Projects 5 (Taylor et al., 2012) and 6 (Eyring
et al., 2016) (CMIP5 and CMIP6). The ESM selection is based on the availability of the outputs regarding marine
biogeochemical processes as well as 3D data of oceanic temperature and currents. Further information on indi-
vidual models used in this study is described in Table S1 in Supporting Information S1 (CMIP5) and S2 (CMIP6),
respectively. To project future changes in Arctic productivity (quantified by net primary production, NPP), we
applied historical and high-emission climate change scenarios up to 2,100. These high-emission scenarios are
known as representative concentration pathways (RCP85) in CMIP5 (Moss et al., 2010) and shared socioeco-
nomic pathways (SSP5-8.5) in CMIP6 (O'Neill et al., 2016). To compare the different ESMs in the same grid
system, the ESM outputs are re-gridded to the Gaussian grids in 1° × 1° horizontal resolutions and 3D oceanic

NOH ET AL. 2 of 13
Earth’s Future 10.1029/2023EF003709

variables are interpolated to 35 vertical layers to 5,000 m using distance-weighted average remapping (climate
data operators; remapdis and intlevel) (Schulzweida, 2019). The ocean mixed layer depth (MLD) is defined as
the depth where the density the 10 m depth value exceeds 0.125 kg/m 3 criterion according to CMIP6 protocol.

2.2. Reanalysis Data Set

We used a reanalysis data set from the Ocean ReAnalysis System 5 (ORAS5) to compare with the CMIP projec-
tions in Atlantification and associated physical variables such as temperature, sea-ice concentration, mixed layer
depth, and ocean currents, for the present climate covering the period from 1979 to 2020 (Zuo et al., 2019).
ORAS5 consists of the five ensembles that were re-gridded to 1° × 1° Gaussian grids before taking the ensemble
means in physical variables were used for the comparison. High-resolution satellite-based primary production
data were obtained from vertically generalized production model (VGPM). The VGPM model estimated the net
primary production (NPP) (ref by applying a chlorophyll-based algorithm to the Moderate Resolution Imag-
ing Spectroradiometer (MODIS) and Sea-Viewing Wide Field-of-View Sensor (SeaWiFS) satellite chlorophyll
observations (Behrenfeld & Falkowski, 1997) and covers approximately the period about 19 years (2002–2020)
based on periods in ocean color satellites.

2.3. Ocean Heat Transport and Atlantic Water

In the present investigation, we analyzed the ocean heat transport (HT) traversing the BSO employing a meth-
odology based on an extant research framework (Arthun et al., 2019; Pan et al., 2023; Shu et al., 2022). The HT
through BSO is calculated using monthly ocean temperature and currents as:


 HT(𝑡𝑡) = 𝜌𝜌0 𝑐𝑐𝑝𝑝 𝑈𝑈 (𝑡𝑡𝑡 𝑡𝑡𝑡 𝑡𝑡) ⋅ (𝑇𝑇 (𝑡𝑡𝑡 𝑡𝑡𝑡 𝑡𝑡) − 𝑇𝑇ref ) dS
𝑆𝑆

where ρ0 is the density of sea water, cp is the specific heat capacity of heat water, U is the ocean velocity perpen-
dicular to the section, T is the ocean temperature and Tref is reference temperature set to 0℃ (as done in e.g.,
Arthun et al., 2012, 2019; Pan et al., 2023). Here, t, y, z is the dimension variables corresponding to time, lati-
tude, and depth respectively. The geographical location of the BSO is defined by the longitude 20°E and latitude
spanning from 71°N to 74.5°N.

Furthermore, the Atlantic water is characterized by high temperature (>0℃) and typically occupies intermediate
depths, ranging from 150 to 900 m (Pan et al., 2023; Polyakov et al., 2020, 2023; Shu et al., 2019, 2022). The
Atlantic water exhibits a prominent warm core situated between 200 and 600 m depth, featuring temperatures in
the range of 2–5℃ (Figure S1 in Supporting Information S1). In order to investigate the prospective intrusion and
consequential impact of the Atlantic water, we defined the extent of Atlantic water based on the mean temperature
between depths of 150–900 m, demarcated by 2℃ isothermal contour.

3. Results
3.1. Enhanced Intrusion of the Atlantic Water

To examine the impact of the Atlantification on Arctic productivity, the ensemble–mean Atlantic water extents
are marked every 10 years from 2010 to 2100. The poleward intrusion of the Atlantic water is well captured in a
multi-model ensemble (MME) of the CMIP5 and CMIP6 ESMs (Figure 1a). This poleward intrusion is projected
to expand to the central Arctic Ocean in a high-emission scenario from Svalbard and the southern Barents Sea in
the present climate. The Atlantic water is transported to the central Arctic Ocean in two directions from the Fram
Strait and BSO (Tesi et al., 2021; Tsubouchi et al., 2021). In CMIP5 and CMIP6 ESMs, the extent of the Atlantic
water is also projected to expand toward the Fram Strait and the BSO under greenhouse warming (Figure 1a).
Therefore, we can suggest that the Atlantic water intrusion is consistent with the observational pattern (Muilwijk
et al., 2018, 2023; Tsubouchi et al., 2021) and that the intrusion might expand to the central Arctic Basin in the
future climate change scenario, which indicates the enhanced Atlantification.

The heat transport across the BSO has been enhanced in recent decades and is projected to increase 45.5 ± 34.9 TW
by the end of this century compared to the present level (Figure 1b), which is associated with the sea-ice decline

NOH ET AL. 3 of 13
Earth’s Future 10.1029/2023EF003709

Figure 1. Observed and CMIP5 and CMIP6 multi-model simulated time series compared with observation and reanalysis. (a) Multi-model ensemble (MME) of the
Atlantic water extent represented by the 2°C isothermal line every 10 years. The Atlantic water was defined as the subsurface layer at a depth of 150–900 m. (b) Annual
mean heat transport through the Barents Sea Opening (BSO) (c) Annual vertical mean temperature in the BSO (d) Annual mean primary production in the Barents Sea.
Black solid lines represent the time series of the MME in CMIP5 and 6 models for the period from 1901 to 2100. Gray shades show the range of inter-model diversities
(±1 s.d.). Reanalysis (heat transport and vertical mean temperature from the Ocean Reanalysis System 5 reanalysis, ORAS5) and satellite observations (primary
production from the vertically generalized production model, VGPM) are overlapped on the CMIP projections with solid blue (heat transport), red (mean temperature),
and green (primary production) lines, respectively. The Atlantic water intrusion of individual models is described in Figure S4 in Supporting Information S1.

in the Barents Sea (Arthun et al., 2012; Li et al., 2017). This positive trend of heat transport in MME continues
until 2100, with large uncertainty represented by inter-model diversity (Figure 1b). The increased heat transport
across the BSO is caused by the increased temperature of the Atlantic water across the BSO (Figure 1c), reaching
up to 8℃ at the end of the twenty-first century from 4℃ in the present climate. Sea surface temperature (SST) in
the Barents Sea usually exhibits increases more than in the other basins in the Arctic Ocean due to the increased
heat transport through the BSO (Figure S2 in Supporting Information S1), which implies the strong sensitivity of
the Barents Sea to anthropogenic carbon emissions.

As the CMIP5 and CMIP6 ESMs represent the largest increases of SST in the Barents Sea compared to the
other Arctic Oceans (Figure S2 in Supporting Information S1), the other physical components also experience
large changes in the Barents Sea. These physical changes are combined to alter biogeochemical states in the
Barents Sea such as the chlorophyll, nutrients, light availability, and finally the primary production (Kwiatkowski

NOH ET AL. 4 of 13
Earth’s Future 10.1029/2023EF003709

Figure 2. Changes in the primary production in the Barents Sea. (a) Seasonal primary productions in the Barents Sea are represented by the last 20 years of the
twentieth and twenty-first centuries in the historical scenario (1980–1999) and the future scenarios (2080–2099) in representative concentration pathway (RCP85) and
shared socioeconomic pathway (SSP5-85) respectively. The multi-model ensemble (MME) of the primary production in two periods is shown as a blue bar (historical)
and a red bar (RCP85 and SSP5-85). Future seasonal changes of MME in primary production are shown as gray bars. Error bars are represented by the standard
deviation of CMIP5 and CMIP6 model diversities. Future changes in the primary production in seasons are represented by boreal (b) winter (c) spring (d) summer and
(e) autumn. Future changes in primary production are calculated by the differences between the mean primary productions over the last 20 years of the twentieth and
twenty-centuries. Black solid lines represent 0°C surface isothermal lines of SST climatology in the last 20 years of the twentieth century, which can divide the regions
between the southwestern and northern Barents Sea (Schauer et al., 1997). Hatches indicate the region where more than two-thirds of the models agree with the sign of
the MME. The blue box in (b)–(e) represents the Barents Sea, which is also used in Figures 3 and 4.

et al., 2020). Marine productivity in the Arctic Ocean is projected to increase in the future climate change
scenario (Vancoppenolle et al., 2013) with declined phytoplankton biomass (Noh, Lim, et al., 2023). This
increased productivity in the Barents Sea is more distinctive than that in other basins in the Arctic Ocean (Figure
S3 in Supporting Information S1). The annual productivity in the Barents Sea has increased by approximately
16.4 ± 11.2 TgC/year in response to the high-emission scenario (Figure 1d). However, the increased productivity
in the Barents Sea shows a large inter-model spread ranging from 1.1% (0.2 TgC/year) to 310.8% (41.9 TgC/
year) increases compared with the current levels, in addition to one decreasing model: GISS-E2-R-CC. Primary
production is calculated as the vertical integration of the product of phytoplankton biomass and the growth
rate in phytoplankton species (Behrenfeld & Falkowski, 1997). These two factors are influenced by changes in
physical states such as temperature, sea–ice, and vertical mixing, and biogeochemical states such as light supply,
and nutrient availability. These physical and biogeochemical variables show strong seasonalities (Wassmann &
Reigstad, 2011), which suggests the need to investigate seasonal changes in detail.

3.2. Future Changes in Marine Productivity in the Barents Sea

To investigate how the productivity in the Barents Sea is increased under greenhouse warming, we initially
checked the seasonality of productivity in the Barents Sea (Figure 2). The sea-ice extent shows strong seasonal
dependency because the sea–ice covers a large part of the Barents Sea until boreal spring (March–May), and it
retreats to the northern Barents Sea in boreal summer (June to August) (Figure S5 in Supporting Information S1).
The sea-ice retreat provides additional light to the upper ocean so that the ocean becomes a favorable environ-
ment for phytoplankton growth. In the present climate, a large area of the Barents Sea is fully covered with the
sea–ice in boreal winter (December to February), and the productivity in winter is nearly zero (Figure 2a). As the
open-ocean area increases, the productivity increases in spring and peaks in the boreal summer.

In the high-emission scenario, the SST also shows the most significant increase from Iceland to the Barents
Sea in all seasons (Figures S5a–S5d in Supporting Information S1) because the warm Atlantic water is tightly

NOH ET AL. 5 of 13
Earth’s Future 10.1029/2023EF003709

coupled with large air–ice–ocean interactions so that the Barents Sea is more
sensitive to greenhouse warming than other regions in the Arctic Ocean
(Smedsrud et al., 2013). The sea–ice in the Barents Sea in the CMIP5 and
CMIP6 also decreases the most in winter and spring (Figures S5e–S5H in
Supporting Information S1) when the sea–ice covers an area larger than in
other seasons. Comparing the mean states in primary production between
present (1981–1999) and future (2081–2099) climate, although produc-
tivity peaks in boreal summer, the increased productivity is larger in
boreal spring (ΔPPMAM = 46.7 ± 30.4 TgC/year) than in boreal summer
(ΔPPJJA = 19.2 ± 26.2 TgC/year). The productivity in spring has increased
about four times compared with the present climate level (Figure 2a). The larg-
est increase in spring is clearly shown in the spatial pattern of the productivity
changes in the Barents Sea (Figures 2b–2e). In response to the high-emission
scenario, the increased productivity in spring is usually concentrated in the
northern Barents Sea between the Svalbard Islands and Novaya Zemlya. The
spring sea-ice edge, which represents 15% of the iso-concentration line of
sea–ice, moves poleward to the northern Barents Sea in response to green-
house warming, resulting in the largest changes in productivity.

Although the productivity in summer (JJA) is the largest (PPJJA = 66.4 ± 37.7
and 85.6 ± 48.5 TgC/year in the present and future climate) among all seasons
(Figure 2a), the increase in the productivity in summer is relatively small
when compared with that in spring. This discrepancy is caused by seasonal-
ity changes in primary productivity (Figure 3a). The sea–ice in the Barents
Sea is fully covered in spring, and a large part of the sea–ice (approximately
50%) still exists in future climate (Figure 3b). In the present climate, the
Figure 3. Seasonality changes in primary productivity and sea-ice in the
Barents Seas. Seasonality of primary production (a) and sea-ice concentration sea–ice starts to decrease from June to July, which promotes phytoplank-
(b) in the Barents Sea are represented by the last 20 years of the 20th and 21st ton bloom by providing additional light. However, in the future climate,
century in the historical scenario (1980–1999) and the RCP85 and SSP5-85 the climatological sea–ice concentration largely decreases in the warming
(2080–2099) respectively. The MME of seasonality in two periods is shown climate, and the timing of the sea–ice decline moves from April to May. This
as the blue line (historical) and the red line (RCP85 and SSP5-85). Blue and
early decline in sea–ice induces a shift in the peak-timing of the productivity
red shades show the range of inter-model diversities (±1 s.d.). Mean seasonal
changes in PP are integrated above the northern Barents Sea as visually from July to May. Therefore, the change in productivity becomes the largest
delineated by the blue boxes in Figure 2. in spring, and a relatively small increase occurs in summer because the nutri-
ents are consumed by the bloomed phytoplankton in spring (Wassmann &
Reigstad, 2011).

3.3. Physical Environments and Marine Productivity in the Present and Future

To further understand the increased productivity in the Barents Sea, we investigated the physical changes associ-
ated with the expansion of the Atlantic water. Initially, we compared observational changes from satellite data and
reanalysis data sets in the last decades with future changes simulated in CMIP5 and CMIP6. The primary produc-
tion has increased in the Barents Sea and this positive trend will continue until the end of the twenty-first century
(Figures 4a and 4b). This increasing trend, however, shows a large inter-model spread, ranging from 0.3 TgC/year
to 124.3 TgC/year. In response to greenhouse warming, the largest SST increase occurs near the sea–ice edge
(Figures 4c and 4d) caused by enhanced Atlantic water intrusion (Shu et al., 2021, 2022). SST changes also show
a large inter-model diversity in the CMIP5 and CMIP6 ESMs ranging from even a slight decrease by −0.5℃ to
an extreme increase by approximately 7℃. SST changes are closely related to the primary production changes in
the Barents Sea (r = 0.55, P < 0.01). The increased temperature can weaken the temperature limitation, which is
usually described as the exponential function in the model parameterization (Eppley, 1972). This implies that the
model with a larger increase in SST tends to simulate more increased productivity (Figure 4e).

The sea–ice concentration (SIC) decrease in the Barents Sea is usually positioned outside the sea-ice edge both
in the reanalysis and CMIP outputs (Figures 4f and 4g). The mean SIC in the Barents Sea decreases by approx-
imately 53.2% ± 24.0% from 90.7% ± 7.5% in the current climate to 37.5% ± 27.0% in the future climate. The
SIC area in the Barents Sea is strongly correlated with the trend of the Atlantic heat transport across the BSO

NOH ET AL. 6 of 13
Earth’s Future 10.1029/2023EF003709

Figure 4. Projections in physical states and primary production in the Barents Sea and relationships with primary production.
Primary production (PP) changes on satellite-based observation (a) and CMIP5 and CMIP6 multi-model ensemble (MME)
(b). SST changes on the reanalysis (c) and CMIP5 and CMIP6 MME (d). SIC changes on the reanalysis (f) and CMIP5 and
CMIP6 MME (g). MLD changes on the reanalysis (i) and CMIP6 MME (j). Owing to short observation periods of primary
production (2002–2018), the changes in the reanalysis data sets are calculated by the differences between two periods P1
(2002–2010) and P2 (2011–2018). CMIP5 and CMIP6 MME changes are calculated as the difference between the last
20 years of the twentieth and twenty-first centuries. Black solid lines represent 0°C isothermal lines of SST climatology in the
last 20 years of the twentieth century. Inter-model relationship between PP changes and other physical variables changes in
the Barents Sea: SST (e), SIC (h), and MLD (k). Scatters represent the mean changes in the variables in the Barents Sea and
are shown with different colors and markers depending on individual models. CMIP5 and CMIP6 models are represented as
empty scatters and filled scatters, respectively. Mean differences in PP, SST, SIC, and MLD between present and projected
future climate are integrated above the northern Barents Sea as visually delineated by the blue boxes in Figure 2.

rather than the global mean surface air temperature (Li et al., 2017). As the reduction of the Arctic SIC has
accelerated productivity with the additional light supply (Arrigo et al., 2008), the Barents Sea becomes a more
favorable environment for phytoplankton growth. As expected, the SIC declines show a strong relationship with
the increased productivity in the Barents Sea (r = 0.65, P < 0.01) (Figure 4h) with large inter-model diversity
of sea-ice declines ranging from −2.3% to −91.7%. This relationship implies that the weakened light limitation
due to the retreat of the sea-ice edge can contribute to the larger intensity of spring bloom and result in increased
productivity in the Barents Sea.

Although the mixed layer depth (MLD) is globally projected to be shallower with enhanced stratification
(Kwiatkowski et al., 2020; Li et al., 2020), MLD has been deepened in subregions of the Arctic Ocean because
expanded ice-free regions make the mixed layer more sensitive to wind (Peralta-Ferriz & Woodgate, 2015).
Specifically, the Atlantic water enhances the vertical mixing in the Barents–Kara seas in response to the
increased oceanic heat loss (Shu et al., 2021). In the CMIP6 ESMs, the MLD in the Barents Sea also increases

NOH ET AL. 7 of 13
Earth’s Future 10.1029/2023EF003709

to approximately 25.7 ± 28.1 m in a warming climate, suggesting that additional nutrients can be supplied to
the upper ocean. In other words, the enhanced vertical mixing weakens nutrient limitation, which results in the
increased productivity (Randelhoff et al., 2020). The strong relationship between the deeper MLD and increased
productivity (r = 0.71, P < 0.01) suggests that the increased productivity in the Barents Sea is influenced by
the weakened nutrient limitation. This relationship implies that ESMs with stronger vertical mixing can lead to
increased productivity in the Arctic Ocean.
These physical factors that affect productivity changes are strongly correlated among them; thus, their direct
effect from the linear correlation should be interpreted carefully. To estimate the relative contributions of
increased productivity in the Barents Sea for the SST, SIC, and MLD, we conducted multiple-linear regres-
sions in inter-model diversity (ΔPP = α ΔSST + β ΔSIC + γ ΔMLD). Because only CMIP6 provides the MLD
data, 12 CMIP6 models are used to calculate the contributions of individual factors. So, statistical significance
can be limited, and the results should be carefully interpreted. The multiple regression results show that SST
changes explain the least approximately 13% of the inter-model diversity in the productivity changes in response
to anthropogenic emissions and that SIC changes explain approximately 27%. In addition, productivity changes
can be explained by MLD changes of approximately 38%. Although the MLD changes explain the largest portion
of the inter-model diversity in the productivity changes, the contribution of each physical factor to future MME
changes in productivity can be different. Because the ratio of the MME changes relative to the inter-model spread
is the highest in the SIC (2–4 std), the contribution of the sea-ice changes to the MME productivity change can
be higher.

3.4. Role of Atlantification in Marine Productivity

Using the inter-model relationship between changes in productivity and other physical variables, we suggest that
productivity changes are caused by the combined result of the weakened limitations in temperature, light, and
nutrients. Atlantification triggers these physical changes in MME (Shu et al., 2021), so we attempt to quantify
how the intensity of Atlantification influences physical and biogeochemical changes in inter-model diversity.
We first quantified the strength of Atlantification as the heat transport across the BSO. Based on this definition,
we find a significant relationship between the intensified Atlantification and the increased primary production
(r = 0.51, P < 0.01) (Figure 5a). Both spatial patterns in the inter-model regression and MME changes commonly
show increased productivity in the Barents Sea, and their similar spatial patterns demonstrate the effect of the
Atlantification on productivity (Figure 5b). These results suggest that enhanced heat transport across the BSO
plays an important role in the productivity changes in the Barents Sea. From the inter-model spread, we can
estimate that the Atlantification contributes to the increased productivity under greenhouse warming by approx-
imately 26% of the inter-model diversity.
Atlantification modulates productivity by affecting the physical environment. The Atlantification is highly corre-
lated with the favorable conditions of the physical variables for larger productivity. The correlation coefficients
with the heat transport from the BSO are 0.68, −0.56, and 0.69 for SST (Figure 5d), SIC (Figure 5g), and vertical
mixing (Figure 5j), respectively, which are all statistically significant at the 99% confidence level. These corre-
lations imply that the enhanced Atlantification contributes to the increased SST, reduced SIC by the warm water
advection, and enhanced vertical mixing due to surface heat loss induced by the retreating of the sea-ice edge.
These effects can be intensified via the SIC-albedo feedback, yielding a higher correlation than present values.
To further show the effect of the Atlantification on physical changes, inter-model regression patterns are shown
for SST (Figure 5e), SIC (Figure 5h), and MLD (Figure 5k). Strikingly, the regression patterns in the inter-model
space are similar to the future projections in physical states (Figures 4d–4g, and 4j). In addition to the inter-model
spread, the effects of the Atlantification consistently appear in the MME regression as well, supporting the
robustness of the Atlantification effects. Interestingly, the spatial patterns of SST and SIC associated with the
Atlantification are quite similar to those of the productivity changes (Figures 5b–5c), although the pattern of
the vertical mixing is a bit noisy. These results suggest that future changes in the Atlantification can be critical for
future changes in the primary productivity in the Barents Sea by modulating physical variables.

4. Conclusion and Discussion


In recent decades, the influence of the Atlantic water has been enlarged in the Arctic Ocean because of the
increased heat transport from the BSO and has manifested the changes in physical and ecological states,

NOH ET AL. 8 of 13
Earth’s Future 10.1029/2023EF003709

Figure 5. Effect of the Atlantification on the physical states and primary production in the Barents Sea. Inter-model
relationship between heat transport (HT) from the BSO and (a) primary production (PP), (d) sea surface temperature (SST),
(g) sea–ice concentration (SIC), (j) mixed layer depth (MLD). Scatters represent the mean changes of the variables in the
Barents Sea and are shown with different colors and markers depending on ESMs. CMIP5 and CMIP6 ESMs are represented
as open scatter and filled scatter, respectively. The legend of the scatters describing individual models is consistent with
the legend in Figure 3 except for the GFDL-ESM4 model which did not provide the u-current. Inter-model regression and
MME regression map between the BSO HT changes and the spatial pattern of (b), c) PP, (e), f) SST, (h), i) SIC, (K,L) MLD
changes under RCP85 and SSP5-85 to historical simulation, respectively. In individual models, detrended anomalies in HT
are regressed on anomalies in other variables during the twenty-first century to represent the impact of HT variability on
variables. The regression coefficients are averaged to represent MME regressions. Hatches indicate the region where the
regression coefficients exceed the 95% confidence level and where more than two-thirds of models agree. Mean differences
in PP, SST, SIC, and MLD between present and projected future climate are integrated above the northern Barents Sea as
visually delineated by the blue boxes in Figure 2.

particularly in the Barents Sea (Ingvaldsen et al., 2021). Atlantification has influenced physical changes such as
increased surface temperature, reduced sea–ice extent, and increased vertical mixing (Shu et al., 2021) and these
physical changes have affected the productivity in the Barents Sea (Fransner et al., 2023; Polyakov et al., 2020).
We estimated diverse projected changes in Atlantification and firstly suggested the detailed mechanism of how
the Atlantification influences the spring productivity changes in the Barents Sea based on satellite observa-
tions, reanalysis, and model outputs participated in the CMIP5 and CMIP6. The present result exhibits that the
Atlantification has been enhanced (45.5 ± 34.9 TW) and Atlantic water intrusion is projected to expand to the
central Arctic Ocean. These Atlantification-induced physical changes contribute to the increased levels of spring

NOH ET AL. 9 of 13
Earth’s Future 10.1029/2023EF003709

productivity in the Barents Sea (ΔPPMAM = 46.7 ± 30.4 TgC/year) by weakening temperature, light, and nutrient
limitations for phytoplankton growth. Based on the estimation using the inter-model regression, the intensified
Atlantification can explain the increased productivity by 26% of total changes in response to anthropogenic emis-
sions. In summary, the future intensified Atlantification may increase the spring productivity in the Barents Sea
interacting with the Arctic systems from atmosphere, ocean, to cryosphere.

Phytoplankton communities have different thermal tolerance for growth and plankton functional groups (PFTs),
and they react differently in response to global warming depending on their composition (Laufkotter et al., 2015).
The Atlantification drives poleward expansion of temperature phytoplankton species such as coccolithophore,
which is a marine-calcifying phytoplankton species (Neukermans et al., 2018; Oziel et al., 2020). As the physical
and biogeochemical states become the “Atlantified” (Ingvaldsen et al., 2021), the Atlantic phytoplankton species
expands poleward and replaces the Arctic phytoplankton species (Orkney et al., 2020), which is represented
as the “borealization” (Polyakov et al., 2020). Future Atlantification is projected to be intensified under global
warming (Figure 1), so these ecosystem shifts will also persist or even accelerate. Therefore, additional detailed
investigations are necessary to predict the marine ecosystem changes in the Barents Sea considering the different
effects of PFTs.

Productivity changes are strongly correlated with the Atlantic water only in the spring. In the warm (summer and
autumn) seasons, the MLD is too shallow to provide additional heat fluxes from the Atlantic water. However, in
the cold (winter and spring) seasons, the MLD is already deep enough to allow the interaction between the surface
to subsurface layer and becomes deeper in response to greenhouse warming. In a warming climate, the deep-
ening of MLD facilitates positive feedbacks between sea–ice dynamics and heat transport through BSO (Lind
et al., 2018; Shu et al., 2021). Notably, our current analysis is confined to linear processes and does not encom-
pass these intricate dynamics. The projected responses in MLD indicate amplified MLD seasonality in the warm-
ing climate, and the increased seasonality implies that the increased heat transport due to Atlantification impacts
the surface process only during cold seasons when the vertical mixing in the upper ocean is sufficiently vigorous
to reach the Atlantic water. Consequently, further investigations employing modeling approaches to elucidate the
seasonality and feedbacks associated with Atlantification are needed for predictive capacity concerning future
productivity within the Barents Sea ecosystem.

Conflict of Interest
The authors declare no conflicts of interest relevant to this study.

Data Availability Statement


ORAS5 reanalysis is provided at https://www.ecmwf.int/en/forecasts/dataset/ocean-reanalysis-system-5 (Zuo
et al., 2019) and VGPM primary production estimation is provided at http://orca.science.oregonstate.edu/1080.
by.2160.monthly.hdf.vgpm.m.chl.m.sst.php (Behrenfeld & Falkowski, 1997), respectively. The CMIP5 and
CMIP6 archives are freely available from https://esgf-node.llnl.gov/ (Eyring et al., 2016; Taylor et al., 2012).
The data used in this study will be available from https://doi.org/10.6084/m9.figshare.22637329.v1 (Noh, Oh,
et al., 2023). All figures were generated by using software package Python with the matplotlib and basemap
modules (https://matplotlib.org/, https://matplotlib.org/basemap/). The map coastlines are derived by the Global
Self-consistent, Hierarchical, High-resolution Geography (GSHHG) Database (www.soest.hawaii.edu/pwessel/
gshhg/), which has been distributed under the GNU Lesser General Public License and is provided with the
Acknowledgments
This research was supported by the basemap Python module.
National Research Foundation of
Korea (NRF-2022R1A3B1077622,
NRF-2021M3I6A1086808) and the References
National Supercomputing Center with
supercomputing resources, associated Aagaard, K., Coachman, L. K., & Carmack, E. (1981). On the halocline of the Arctic Ocean. Deep Sea Research Part I: Oceanographic Research
technical support KSC-2023-CHA-0001. Papers, 28(6), 529–545. https://doi.org/10.1016/0198-0149(81)90115-1
H.-G. Lim was supported by a project Ardyna, M., & Arrigo, K. R. (2020). Phytoplankton dynamics in a changing Arctic Ocean. Nature Climate Change, 10(10), 1–12. https://doi.
titled “Ocean Circulation and ecosystem org/10.1038/s41558-020-0905-y
variability and predictability research Arrigo, K. R., van Dijken, G., & Pabi, S. (2008). Impact of a shrinking Arctic ice cover on marine primary production. Geophysical Research
in the earth system model (PEA0175)” Letters, 35(19), L19603. https://doi.org/10.1029/2008GL035028
funded by the Korea Institute of Ocean Arrigo, K. R., & van Dijken, G. L. (2015). Continued increases in Arctic Ocean primary production. Progress in Oceanography, 136(C), 60–70.
Science and Technology. https://doi.org/10.1016/j.pocean.2015.05.002

NOH ET AL. 10 of 13
Earth’s Future 10.1029/2023EF003709

Arthun, M., Eldevik, T., & Smedsrud, L. H. (2019). The role of Atlantic heat transport in future Arctic winter sea ice loss. Journal of Climate,
32(11), 3327–3341. https://doi.org/10.1175/JCLI-D-18-0750.1
Arthun, M., Eldevik, T., Smedsrud, L. H., Skagseth, O., & Ingvaldsen, R. B. (2012). Quantifying the influence of Atlantic heat on Barents Sea
ice variability and retreat. Journal of Climate, 25(13), 4736–4743. https://doi.org/10.1175/JCLI-D-11-00466.1
Asbjørnsen, H., Årthun, M., Skagseth, Ø., & Eldevik, T. (2020). Mechanisms underlying recent Arctic Atlantification. Geophysical Research
Letters, 47(15), e2020GL088036. https://doi.org/10.1029/2020GL088036
Behrenfeld, M. J., & Falkowski, P. G. (1997). Photosynthetic rates derived from satellite-based chlorophyll concentration. Limnology & Ocean-
ography, 42(1), 1–20. https://doi.org/10.4319/lo.1997.42.1.0001
Bopp, L., Resplandy, L., Orr, J. C., Doney, S. C., Dunne, J. P., Gehlen, M., et al. (2013). Multiple stressors of ocean ecosystems in the 21st
century: Projections with CMIP5 models. Biogeosciences, 10(10), 6225–6245. https://doi.org/10.5194/bg-10-6225-2013
Cabré, A., Marinov, I., & Leung, S. (2014). Consistent global responses of marine ecosystems to future climate change across the IPCC AR5
Earth system models. Climate Dynamics, 45(5–6), 1253–1280. https://doi.org/10.1007/s00382-014-2374-3
Cohen, J., Screen, J. A., Furtado, J. C., Barlow, M., Whittleston, D., Coumou, D., et al. (2014). Recent Arctic amplification and extreme
mid-latitude weather. Nature Geoscience, 7(9), 627–637. https://doi.org/10.1038/NGEO2234
Csapo, H. K., Grabowski, M., & Westawski, J. M. (2021). Coming home—Boreal ecosystem claims Atlantic sector of the Arctic. The Science of
the Total Environment, 771, 144817. https://doi.org/10.1016/j.scitotenv.2020.144817
Docquier, D., & Koenigk, T. (2021). A review of interactions between ocean heat transport and Arctic sea ice. Environmental Research Letters,
16(12), 123002. https://doi.org/10.1088/1748-9326/ac30be
Eppley, R. W. (1972). Temperature and phytoplankton growth in the sea. Fishery Bulletin, 70(4), 1063–1085.
Eyring, V., Bony, S., Meehl, G. A., Senior, C. A., Stevens, B., Stouffer, R. J., & Taylor, K. E. (2016). Overview of the coupled model inter-
comparison project phase 6 (CMIP6) experimental design and organization. Geoscientific Model Development, 9(5), 1937–1958. https://doi.
org/10.5194/gmd-9-1937-2016
Fransner, F., Olsen, A., Arthun, M., Counillon, F., Tjiputra, J., Samuelsen, A., & Keenlyside, N. (2023). Phytoplankton abundance in the Barents
Sea is predictable up to five years in advance. Communications Earth & Environment, 4(141), 141. https://doi.org/10.1038/s43247-023-00791-9
Ingvaldsen, R. B., Assmann, K. M., Primicerio, R., Fossheim, M., Polyakov, I. V., & Dolgov, A. V. (2021). Physical manifestations and ecological
implications of Arctic Atlantification. Nature Reviews Earth and Environment, 2(12), 874–889. https://doi.org/10.1038/s43017-021-00228-x
Ivanov, V., Alexeev, V., Koldunov, N. V., Repina, I., Sando, A. B., Smedsrud, L. H., & Smirnov, A. (2016). Arctic Ocean heat impact on regional
ice decay: A suggested positive feedback. Journal of Physical Oceanography, 46(5), 1437–1456. https://doi.org/10.1175/JPO-D-15-0144.1
Jang, Y.-S., Jun, S.-Y., Son, S.-W., Min, S.-K., & Kug, J.-S. (2021). Delayed impacts of Arctic Sea-ice loss on Eurasian Severe cold winters.
Journal of Geophysical Research, 126(23), e2021JD035286. https://doi.org/10.1029/2021JD035286
Jansen, E., Christensen, J. H., Dokken, T., Nisancioglu, K. H., Vinther, B. M., Capron, E., et al. (2020). Past perspectives on the present era of
abrupt Arctic climate change. Nature Climate Change, 10(8), 714–721. https://doi.org/10.1038/s41558-020-0860-7
Kahru, M., Lee, Z., Mitchell, B. G., & Nevison, C. D. (2016). Effects of sea ice cover on satellite-detected primary production in the Arctic Ocean.
Biology Letters, 12(11), 20160223. https://doi.org/10.1098/rsbl.2016.0223
Kim, D.-S., Jun, S.-Y., Lee, M.-I., & Kug, J.-S. (2022). Significant relationship between Arctic warming and East Asia hot summers. International
Journal of Climatology, 42(16), 1–9. https://doi.org/10.1002/joc.7844
Kug, J.-S., Jeong, J.-H., Jang, Y.-S., Kim, B.-M., Folland, C. K., Min, S.-K., & Son, S.-W. (2015). Two distinct influences of Arctic warming on
cold winters over North America and East Asia. Nature Geoscience, 8(10), 759–762. https://doi.org/10.1038/ngeo2517
Kwiatkowski, L., Torres, O., Bopp, L., Aumont, O., Chamberlain, M., Christian, J. R., et al. (2020). Twenty-first century ocean warming, acid-
ification, deoxygenation, and upper-ocean nutrient and primary production decline from CMIP6 model projections. Biogeosciences, 17(13),
3439–3470. https://doi.org/10.5194/bg-17-3439-2020
Laufkotter, C., Vogt, M., Gruber, N., Aita-Noguchi, M., Aumont, O., Bopp, L., et al. (2015). Drivers and uncertainties of future global marine
primary production in marine ecosystem models. Biogeosciences, 12(23), 6955–6984. https://doi.org/10.5194/bg-12-6955-2015
Lewis, K. M., van Dijken, G. L., & Arrigo, K. R. (2020). Changes in phytoplankton concentration now drive increased Arctic Ocean primary
production. Science, 369(6500), 198–202. https://doi.org/10.1126/science.aay8380
Li, D., Zhang, R., & Knutson, T. R. (2017). On the discrepancy between observed and CMIP5 multi-model simulated Barents Sea winter sea ice
decline. Nature Communications, 8(1), 7. https://doi.org/10.1038/ncomms14991
Li, G., Cheng, L., Zhu, J., Trenberth, K. E., Mann, M. E., & Abraham, J. P. (2020). Increasing ocean stratification over the past half-century.
Nature Climate Change, 10(12), 1116–1123. https://doi.org/10.1038/s41558-020-00918-2
Lind, S., Ingvaldsen, R. B., & Furevik, T. (2018). Arctic warming hotspot in the northern Barents Sea linked to declining sea-ice import. Nature
Climate Change, 8(7), 634–639. https://doi.org/10.1038/s41558-018-0205-y
Meredith, M., Sommerkorn, M., Cassotta, S., Derksen, C., Ekaykin, A., Hollowed, A., et al. (2019). Polar regions. In IPCC special report on the
ocean and cryosphere in a changing climate.
Moss, R. H., Edmonds, J. A., Hibbard, K. A., Manning, M. R., Rose, S. K., van Vuuren, D. P., et al. (2010). The next generation of scenarios for
climate change research and assessment. Nature, 463(7282), 1–10. https://doi.org/10.1038/nature08823
Muilwijk, M., Nummelin, A., Heuze, C., Polyakov, I. V., Zanowski, H., & Smdsrud, L. H. (2023). Divergence in climate model projections of
future Arctic Atlantification. Journal of Climate, 36(6), 1727–1749. https://doi.org/10.1175/JCLI-D-22-0349.1
Muilwijk, M., Smedsrud, L. H., Ilicak, M., & Drange, H. (2018). Atlantic Water heat transport variability in the 20th century Arctic Ocean from a
global ocean model and observations. Journal of Geophysical Research: Oceans, 123(11), 8159–8179. https://doi.org/10.1029/2018JC014327
Neukermans, G., Oziel, L., & Babin, M. (2018). Increased intrusion of warming Atlantic water leads to rapid expansion of temperate phytoplank-
ton in the Arctic. Global Change Biology, 24(6), 2545–2553. https://doi.org/10.1111/gcb.14075
Noh, K. M., Lim, H.-G., Yang, E. J., & Kug, J.-S. (2023). Emergent constraint for future decline in Arctic phytoplankton concentration. Earth's
Future, 11(4), e2022EF003427. https://doi.org/10.1029/2022EF003427
Noh, K. M., Oh, J.-H., Lim, H.-G., Song, H., & Kug, J.-S. (2023). Data for “role of Atlantification in enhanced primary productivity in the Barents
Sea” [Dataset]. Figshare. https://doi.org/10.6084/m9.figshare.22637329.v1
Onarheim, I. H., & Arthun, M. (2017). Toward an ice-free Barents Sea. Geophysical Research Letters, 44(16), 8387–8395. https://doi.
org/10.1002/2017GL074304
O'Neill, B. C., Tebaldi, C., van Vuuren, D. P., Eyring, V., Friedlingstein, P., Hurtt, G., et al. (2016). The scenario model intercomparison project
(ScenarioMIP) for CMIP6. Geoscientific Model Development, 9(9), 3461–3482. https://doi.org/10.5194/gmd-9-3461-2016
Orkney, A., Platt, T., Narayanaswamy, B. E., Kostakis, I., & Bouman, H. A. (2020). Bio-optical evidence for increasing phaeocystis dominance in
the Barents Sea. Philosophical Transactions of the Royal Society A: Mathematical, Physical & Engineering Sciences, 378(2181), 20190357.
https://doi.org/10.1098/rsta.2019.0357

NOH ET AL. 11 of 13
Earth’s Future 10.1029/2023EF003709

Oziel, L., Baudena, A., Ardyna, M., Massicotte, P., Randelhoff, A., Sallée, J. B., et al. (2020). Faster Atlantic currents drive poleward expansion
of temperate phytoplankton in the Arctic Ocean. Nature Communications, 11(1705), 1–8. https://doi.org/10.1038/s41467-020-15485-5
Pabi, S., van Dijken, G. L., & Arrigo, K. R. (2008). Primary production in the Arctic Ocean, 1998-2006. Journal of Geophysical Research,
113(C8), C08005. https://doi.org/10.1029/2007JC004578
Pan, R., Shu, Q., Wang, Q., Wang, S., Song, Z., He, Y., & Qiao, F. (2023). Future Arctic climate change in CMIP6 strikingly intensified by
NEMO-family climate models. Geophysical Research Letters, 50(4), e2022GL102077. https://doi.org/10.1029/2022GL102077
Perlalta-Ferriz, C., & Woodgate, R. A. (2015). Seasonal and interannual variability of pan-Arctic surface mixed layer properties from 1979 to
2012 from hydrographic data, and the dominance of stratification for multiyear mixed layer depth shoaling. Progress in Oceanography, 134,
19–53. https://doi.org/10.1016/j.pocean.2014.12.005
Pithan, F., & Mauritsen, T. (2014). Arctic amplification dominated by temperature feedbacks in contemporary climate models. Nature Geosci-
ence, 7(3), 181–184. https://doi.org/10.1038/ngeo2071
Polyakov, I. V., Alkire, M. B., Bluhm, B. A., Brown, K. A., Carmack, E. C., Chierici, M., et al. (2020). Borealization of the Arctic Ocean in
response to anomalous advection from sub-Arctic seas. Frontiers in Marine Science, 7. https://doi.org/10.3389/fmars.2020.00491
Polyakov, I. V., Ingvaldsen, R. B., Pnyushkov, A. V., Bhatt, U. S., Francis, J. A., Janout, M., et al. (2023). Fluctuating Atlantic inflows modulate
Arctic Atlantification. Science, 381(6661), 972–979. https://doi.org/10.1126/science.adh5158
Polyakov, I. V., Pnyushkov, A. V., Alkire, M. B., Ashik, I. M., Baumann, T. M., Carmack, E. C., et al. (2017). Greater role for Atlantic inflows on
sea-ice loss in the Eurasian Basin of the Arctic Ocean. Science, 356(6335), 285–291. https://doi.org/10.1126/science.aai8204
Randelhoff, A., Holding, J., Janout, M., Sejr, M. K., Babin, M., Tremblay, J.-E., & Alkire, M. B. (2020). Pan-Arctic ocean primary production
constrained by turbulent nitrate fluxes. Frontiers in Marine Science, 7, 1–15. https://doi.org/10.3389/fmars.2020.00150
Randelhoff, A., Reigstad, M., Chierici, M., Sundfjord, A., Ivanov, V., Cape, M., et al. (2018). Seasonality of the physical and biogeochemical
hydrography in the inflow to the Arctic Ocean through Fram Strait. Frontiers in Marine Science, 5, 3778–3816. https://doi.org/10.3389/
fmars.2018.00224
Schulzweida, U. (2019). CDO user guide. Zenodo, 1–230. https://doi.org/10.5281/zenodo.3539275
Screen, J. A., & Simmonds, I. (2010). The central role of diminishing sea ice in recent Arctic temperature amplification. Nature, 464(7293),
1334–1337. https://doi.org/10.1038/nature09051
Serreze, M. C., Barrett, A. P., Stroeve, J. C., Kindig, D. N., & Holland, M. M. (2009). The emergence of surface-based Arctic amplification. The
Cryosphere, 3(1), 11–19. https://doi.org/10.5194/tc-3-11-2009
Shu, Q., Wang, Q., Arthun, M., Wang, S., Song, Z., Zhang, M., & Qiao, F. (2022). Arctic Ocean Amplification in a warming climate in CMIP6
models. Science Advances, 8(30). https://doi.org/10.1126/sciadv.abn9755
Shu, Q., Wang, Q., Song, Z., & Qiao, F. (2021). The poleward enhanced Arctic Ocean cooling machine in a warming climate. Nature Communi-
cations, 12(1), 2966. https://doi.org/10.1038/s41467-021-23321-7
Shu, Q., Wang, Q., Su, J., Li, X., & Qiao, F. (2019). Assessment of the Atlantic water layer in the Arctic Ocean in CMIP5 climate models. Climate
Dynamics, 53(9), 5279–5291. https://doi.org/10.1007/s00382-019-04870-6
Slagstad, D., Wassmann, P. F. J., & Ellingson, I. (2015). Physical constrains and productivity in the future Arctic Ocean. Frontiers in Marine
Science, 2. https://doi.org/10.3389/fmars.2015.00085
Smedsrud, L. H., Esau, I., Ingvaldsen, R. B., Eldevik, T., Haugan, P. M., Li, C., et al. (2013). The role of the Barents Sea in the Arctic climate
system. Reviews of Geophysics, 51(3), 415–449. https://doi.org/10.1002/rog.20017
Tagliabue, A., Kwiatkowski, L., Bopp, L., Butenschön, M., Cheung, W., Lengaigne, M., & Vialard, J. (2021). Persistent uncertainties in ocean
net primary production climate change projections at regional scales raise challenges for assessing impacts on ecosystem services. Frontiers in
Climate, 3. https://doi.org/10.3389/fclim.2021.738224
Taylor, K. E., Stouffer, R. J., & Meehl, G. A. (2012). An overview of CMIP5 and the experiment design. Bulletin of the American Meteorological
Society, 93(4), 485–498. https://doi.org/10.1175/BAMS-D-11-00094.1
Tesi, T., Muschitiello, F., Mollenhauer, G., Miserocchi, S., Langone, L., Ceccarelli, C., et al. (2021). Rapid Atlantification along the Fram Strait
at the beginning of the 20th century. Science Advances, 7(48). https://doi.org/10.1126/sciadv.abj2946
Tsubouchi, T., Vage, K., Hansen, B., Larsen, K. M. H., Osterhus, S., Johnson, C., et al. (2021). Increased ocean heat transport into the Nordic Seas
and Arctic Ocean over the period 1993–2016. Nature Climate Change, 11(1), 21–26. https://doi.org/10.1038/s41558-020-00941-3
Vancoppenolle, M., Bopp, L., Madec, G., Dunne, J., Ilyina, T., Halloran, P. R., & Steiner, N. (2013). Future Arctic Ocean primary productiv-
ity from CMIP5 simulations: Uncertain outcome, but consistent mechanisms. Global Biogeochemical Cycles, 27(3), 605–619. https://doi.
org/10.1002/gbc.20055
Vernet, M., Carstensen, J., Reigstad, M., & Svensen, C. (2020). Editorial: Carbon Bridge to the Arctic. Frontiers in Marine Science, 7, 1–4.
https://doi.org/10.3389/fmars.2020.00204
Wassmann, P., & Reigstad, M. (2011). Future Arctic Ocean seasonal ice zones and implications for Pelagic-Benthic coupling. Oceanography,
24(3), 220–231. https://doi.org/10.5670/oceanog.2011.74
Woods, C., & Caballero, R. (2016). The role of moist intrusions in winter Arctic warming and Sea ice decline. Journal of Climate, 29(12),
4473–4485. https://doi.org/10.1175/JCLI-D-15-0773.1
Zuo, H., Balmaseda, M. A., Tietsche, S., Mogensen, K., & Mayer, M. (2019). The ECMWF operational ensemble reanalysis-analysis system
for ocean and sea ice: A description of the system and assessment. Ocean Science, 15(3), 779–808. https://doi.org/10.5194/os-15-779-2019

References From the Supporting Information


Arora, V. K., Scinocca, J. F., Boer, G. J., Christian, J. R., Denman, K. L., Flato, G. M., et al. (2011). Carbon emission limits required to
satisfy future representative concentration pathways of greenhouse gases. Geophysical Research Letters, 38(5), L05805. https://doi.
org/10.1029/2010gl046270
Boucher, O., Servonnat, J., Albright, A. L., Aumont, O., Balkanski, Y., Bastrikov, V., et al. (2020). Presentation and evaluation of the
IPSL-CM6A-LR climate model. Journal of Advances in Modeling Earth Systems, 12(7), 1029–1052. https://doi.org/10.1029/2019ms002010
Danabasoglu, G., Lamarque, J. F., Bacmeister, J., Bailey, D. A., DuVivier, A. K., Edwards, J., et al. (2020). The community Earth system model
version 2 (CESM2). Journal of Advances in Modeling Earth Systems, 12, 106–135. https://doi.org/10.1029/2019ms001916
Dufresne, J. L., Foujols, M. A., Denvil, S., Caubel, A., Marti, O., Aumont, O., et al. (2013). Climate change projections using the IPSL-CM5
Earth system model: From CMIP3 to CMIP5. Climate Dynamics, 40(9–10), 2123–2165. https://doi.org/10.1007/s00382-012-1636-1

NOH ET AL. 12 of 13
Earth’s Future 10.1029/2023EF003709

Dunne, J. P., Horowitz, L. W., Adcroft, A. J., Ginoux, P., Held, I. M., John, J. G., et al. (2020). The GFDL Earth system model version 4.1
(GFDL-ESM 4.1): Overall coupled model description and simulation characteristics. Journal of Advances in Modeling Earth Systems, 12(11),
e2019MS002015. https://doi.org/10.1029/2019ms002015
Dunne, J. P., John, J. G., Adcroft, A. J., Griffies, S. M., Hallberg, R. W., Shevliakova, E., et al. (2012). GFDL’s ESM2 global coupled climate–
carbon Earth system models. Part I: Physical formulation and baseline simulation characteristics. Journal of Climate, 25(19), 6646–6665.
https://doi.org/10.1175/jcli-d-11-00560.1
Giorgetta, M. A., Jungclaus, J., Reick, C. H., Legutke, S., Bader, J., Böttinger, M., et al. (2013). Climate and carbon cycle changes from 1850 to
2100 in MPI-ESM simulations for the coupled model intercomparison project phase 5. Journal of Advances in Modeling Earth Systems, 5(3),
572–597. https://doi.org/10.1002/jame.20038
Hajima, T., Watanabe, M., Yamamoto, A., Tatebe, H., Noguchi, M. A., Abe, M., et al. (2020). Development of the MIROC-ES2L Earth system
model and the evaluation of biogeochemical processes and feedbacks. Geoscientific Model Development, 13(5), 2197–2244. https://doi.
org/10.5194/gmd-13-2197-2020
Jones, C. D., Hughes, J. K., Bellouin, N., Hardiman, S. C., Jones, G. S., Knight, J., et al. (2011). The HadGEM2-ES implementation of CMIP5
centennial simulations. Geoscientific Model Development, 4(3), 543–570. https://doi.org/10.5194/gmd-4-543-2011
Lindsay, K., Bonan, G. B., Doney, S. C., Hoffman, F. M., Lawrence, D. M., Long, M. C., et al. (2014). Preindustrial-control and twentieth-century
carbon cycle experiments with the Earth system model CESM1(BGC). Journal of Climate, 27(24), 8981–9005. https://doi.org/10.1175/
jcli-d-12-00565.1
Mauritsen, T., Bader, J., Becker, T., Behrens, J., Bittner, M., Brokopf, R., et al. (2019). Developments in the MPI-M Earth system model
version 1.2 (MPI-ESM1.2) and its response to increasing CO2. Journal of Advances in Modeling Earth Systems, 11(4), 998–1038. https://doi.
org/10.1029/2018ms001400
Romanou, A., Gregg, W. W., Romanski, J., Kelley, M., Bleck, R., Healy, R., et al. (2013). Natural air-sea flux of CO2 in simulations of the
NASA-GISS climate model: Sensitivity to the physical ocean model formulation. Ocean Modelling, 66, 26–44. https://doi.org/10.1016/j.
ocemod.2013.01.008
Seland, Ø., Bentsen, M., Olivié, D., Toniazzo, T., Gjermundsen, A., Graff, L. S., et al. (2020). Overview of the Norwegian Earth System Model
(NorESM2) and key climate response of CMIP6 DECK, historical, and scenario simulations. Geoscientific Model Development, 13(12),
6165–6200. https://doi.org/10.5194/gmd-13-6165-2020
Sellar, A. A., Jones, C. G., Mulcahy, J. P., Tang, Y., Yool, A., Wiltshire, A., et al. (2019). UKESM1: Description and evaluation of the UK Earth
system model. Journal of Advances in Modeling Earth Systems, 11(12), 4513–4558. https://doi.org/10.1029/2019ms001739
Séférian, R., Delire, C., Decharme, B., Voldoire, A., Salas y Mélia, D., Chevallier, M., et al. (2016). Development and evaluation of CNRM Earth
system model—CNRM-ESM1. Geoscientific Model Development, 9(4), 1423–1453. https://doi.org/10.5194/gmd-9-1423-2016
Séférian, R., Nabat, P., Michou, M., Saint-Martin, D., Voldoire, A., Colin, J., et al. (2019). Evaluation of CNRM Earth system model,
CNRM-ESM2-1: Role of Earth system processes in present-day and future climate. Journal of Advances in Modeling Earth Systems, 11(12),
4182–4227. https://doi.org/10.1029/2019ms001791
Swart, N. C., Cole, J. N. S., Kharin, V. V., Lazare, M., Scinocca, J. F., Gillett, N. P., et al. (2019). The Canadian Earth system model version 5
(CanESM5.0.3). Geoscientific Model Development, 12(11), 4823–4873. https://doi.org/10.5194/gmd-12-4823-2019
Ziehn, T., Chamberlain, M. A., Law, R. M., Lenton, A., Bodman, R. W., Dix, M., et al. (2020). The Australian Earth system model:
ACCESS-ESM1.5. Journal of Southern Hemisphere Earth Systems Science, 70(1), 193–214. https://doi.org/10.1071/es19035

NOH ET AL. 13 of 13

You might also like