The Role of Local and Remote Processes for Wintertime Surface Energy Budget Extremes over Arctic Sea Ice

Lukas Papritz aInstitute for Atmospheric and Climate Science, ETH Zurich, Zurich, Switzerland

Search for other papers by Lukas Papritz in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0002-2047-9544
,
Sonja Murto bDepartment of Meteorology, Stockholm University, Stockholm, Sweden
cBolin Centre for Climate Research, Stockholm University, Stockholm, Sweden

Search for other papers by Sonja Murto in
Current site
Google Scholar
PubMed
Close
,
Matthias Röthlisberger aInstitute for Atmospheric and Climate Science, ETH Zurich, Zurich, Switzerland

Search for other papers by Matthias Röthlisberger in
Current site
Google Scholar
PubMed
Close
,
Rodrigo Caballero bDepartment of Meteorology, Stockholm University, Stockholm, Sweden
cBolin Centre for Climate Research, Stockholm University, Stockholm, Sweden

Search for other papers by Rodrigo Caballero in
Current site
Google Scholar
PubMed
Close
,
Gabriele Messori bDepartment of Meteorology, Stockholm University, Stockholm, Sweden
cBolin Centre for Climate Research, Stockholm University, Stockholm, Sweden
dDepartment of Earth Sciences and Centre of Natural Hazards and Disaster Science, Uppsala University, Uppsala, Sweden

Search for other papers by Gabriele Messori in
Current site
Google Scholar
PubMed
Close
,
Gunilla Svensson bDepartment of Meteorology, Stockholm University, Stockholm, Sweden
cBolin Centre for Climate Research, Stockholm University, Stockholm, Sweden

Search for other papers by Gunilla Svensson in
Current site
Google Scholar
PubMed
Close
, and
Heini Wernli aInstitute for Atmospheric and Climate Science, ETH Zurich, Zurich, Switzerland

Search for other papers by Heini Wernli in
Current site
Google Scholar
PubMed
Close
Free access

Abstract

Arctic warm extremes and anomalous sea ice melting have been linked to episodic injections of warm and moist air from midlatitudes, as well as airmass transformations inside the Arctic. However, the relative importance of remote and local processes for such events remains unclear. Here, we focus on events with extreme positive daily-mean surface energy budget (SEB) anomalies over Arctic sea ice in ERA5 data during extended winters (November–March during 1979–2020). Kinematic backward trajectories from the tropospheric column collocated with the SEB anomalies show that near-surface air is of Arctic origin, whereas air farther aloft is transported poleward from the midlatitudes and ascends. Despite the different origin of the air, the entire tropospheric column shows pronounced potential temperature anomalies (on the order of 10 K) building up along air-parcel trajectories over 2–4 days. Quantifying the contributions of horizontal and vertical transport as well as diabatic processes to the generation of these potential temperature anomalies emphasizes the relevance of horizontal advection across the climatological potential temperature gradient for the generation of the anomalies at all levels. Anomalies aloft are further enhanced by latent heating and those near the surface by subsidence, respectively. Surface heat fluxes over subpolar and polar oceans are key for warming and moistening the near-surface air of predominantly Arctic origin and maintaining a positive potential temperature anomaly, which due to its proximity to the surface unfolds the strongest impact on the SEB. This suggests that Arctic airmasses and their local transformations are crucial for generating the most extreme SEB anomalies.

© 2023 American Meteorological Society. This published article is licensed under the terms of the default AMS reuse license. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Lukas Papritz, lukas.papritz@env.ethz.ch

Abstract

Arctic warm extremes and anomalous sea ice melting have been linked to episodic injections of warm and moist air from midlatitudes, as well as airmass transformations inside the Arctic. However, the relative importance of remote and local processes for such events remains unclear. Here, we focus on events with extreme positive daily-mean surface energy budget (SEB) anomalies over Arctic sea ice in ERA5 data during extended winters (November–March during 1979–2020). Kinematic backward trajectories from the tropospheric column collocated with the SEB anomalies show that near-surface air is of Arctic origin, whereas air farther aloft is transported poleward from the midlatitudes and ascends. Despite the different origin of the air, the entire tropospheric column shows pronounced potential temperature anomalies (on the order of 10 K) building up along air-parcel trajectories over 2–4 days. Quantifying the contributions of horizontal and vertical transport as well as diabatic processes to the generation of these potential temperature anomalies emphasizes the relevance of horizontal advection across the climatological potential temperature gradient for the generation of the anomalies at all levels. Anomalies aloft are further enhanced by latent heating and those near the surface by subsidence, respectively. Surface heat fluxes over subpolar and polar oceans are key for warming and moistening the near-surface air of predominantly Arctic origin and maintaining a positive potential temperature anomaly, which due to its proximity to the surface unfolds the strongest impact on the SEB. This suggests that Arctic airmasses and their local transformations are crucial for generating the most extreme SEB anomalies.

© 2023 American Meteorological Society. This published article is licensed under the terms of the default AMS reuse license. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Lukas Papritz, lukas.papritz@env.ethz.ch

1. Introduction

The dwindling of the Arctic sea ice is one of the most prominent manifestations of global warming (e.g., Richter-Menge and Druckenmiller 2020). A major portion of this trend can be attributed to increasing atmospheric moisture content in the Arctic and its impact on downward radiative fluxes (Francis and Hunter 2006; Graversen and Burtu 2016; Vihma et al. 2016; Lee et al. 2017). Superposed on this negative trend of Arctic sea ice is a substantial regional and interannual variability induced by variations in wind stress and the surface energy budget components (e.g., Ogi and Wallace 2012; Kapsch et al. 2013; Boisvert et al. 2016; Wernli and Papritz 2018; Hofsteenge et al. 2022). Understanding the origin of variations in these forcings and how they impact sea ice has, therefore, received considerable attention by the scientific community.

Previous studies emphasized the coincidence of episodes of anomalously warm temperatures and of anomalies in the atmospheric components of the surface energy budget, hereafter referred to as SEB, with intrusions of warm and moist air from midlatitudes (Francis and Hunter 2006; Doyle et al. 2011; Baggett et al. 2016; Woods and Caballero 2016; Mortin et al. 2016; Persson et al. 2017; Rydsaa et al. 2021; Hofsteenge et al. 2022). During winter, such anomalies can strongly reduce sea ice growth or even promote its melt (e.g., Boisvert et al. 2016). However, the correspondence between extreme SEB and near-surface temperature anomalies is imperfect since SEB anomalies can arise due to anomalies in sensible heat fluxes, latent heat fluxes, and radiative fluxes or a combination of these. While sensible heat fluxes are proportional to the imbalance between atmospheric and surface temperature, and therefore directly related to atmospheric temperature anomalies, downward longwave radiative anomalies are influenced by temperature, water vapor, and clouds within 1 km of the surface (Curry et al. 1995; Ohmura 2001; Shupe and Intrieri 2004; Vargas Zeppetello et al. 2019; Clark et al. 2021). Intrusions of warm and humid air that replace the cold and dry Arctic air lead to a transition of the Arctic atmosphere from a cloud-free, clear state to a cloudy, opaque state with enhanced downwelling longwave radiation at the surface (Doyle et al. 2011; Persson et al. 2017; Pithan et al. 2018). If at the same time the near-surface air is anomalously warm, the combined effect of enhanced sensible heat fluxes and downwelling longwave radiation can lead to extreme SEB anomalies (e.g., Boisvert et al. 2016).

Intrusions of air from lower latitudes into the Arctic occur almost exclusively in narrow sectors, including the Nordic seas, the Bering Strait, and to a lesser extent also the Labrador Sea (Woods et al. 2013; Dufour et al. 2016; Naakka et al. 2019). They are related to planetary waves and large-scale circulation patterns that promote long-range poleward transport [see Henderson et al. (2021) and references therein, as well as Hofsteenge et al. (2022)]. For instance, positive geopotential height anomalies over Scandinavia and the Urals, often related to blocking, combined with negative anomalies over Greenland promote the northward transport of air in the Nordic seas (Luo et al. 2017; Messori et al. 2018; Nygård et al. 2019). While such large-scale patterns are conducive to meridional transport, recent studies have emphasized the relevance of synoptic-scale weather systems embedded in these patterns for effectively accomplishing the transport (Rydsaa et al. 2021; Murto et al. 2022). In particular, episodes of extreme warmth and sea ice melting, as well as intense moisture injections have been attributed—in part—to transport accomplished by midlatitude cyclones and blocks (Sorteberg and Walsh 2008; Woods et al. 2013; Messori et al. 2018; Papritz and Dunn-Sigouin 2020; Fearon et al. 2021). Throughout this study, we will consider warming linked to airmass transports from midlatitudes into the Arctic as a remote process. In addition, weather systems also play an important role in driving local warming within the Arctic. These include subsidence in anticyclones and diabatic heating, for example, in marine cold air outbreaks, which have been found to play an important role in certain cases (Ogi and Wallace 2012; Wernli and Papritz 2018; Papritz 2020). This suggests that both remote and local processes are relevant for SEB extremes in the Arctic but their relative importance remains an open question.

Murto et al. (2023) identified and tracked extreme positive SEB anomalies over Arctic sea ice in reanalysis data. One of the most prominent features of such SEB anomalies is the presence of strongly positive potential temperature anomalies that extend uniformly (of order 10 K) from the lower troposphere to the tropopause. Interestingly, this correspondence is not limited to extreme SEB anomalies but is a general feature of the Arctic atmosphere. This is illustrated in Fig. 1 showing a high vertical coherence between potential temperature anomalies, hereafter denoted θ*, and near-surface θ* as well as SEB anomalies averaged over the polar cap 85°N during winter. Hence, negative SEB anomalies tend to coincide with anomalously cold temperatures across the troposphere and vice versa for positive SEB anomalies.

Fig. 1.
Fig. 1.

Correlations between θ* at a given level with θ* at 940 hPa (red diamonds) and the SEB anomaly (bars) for all days during extended winter [November–March (NDJFM)] during 1979–2020 at 1200 UTC averaged over the polar cap poleward of 85°N. Data are from ERA5 (see section 2).

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

These findings might suggest that positive SEB anomalies occur when the entire tropospheric column of cold and dry Arctic air is replaced by a vertically coherent column of warm and moist air originating at lower latitudes. However, as warm air impinges on the dome of cold polar air, the warm air tends to ascend along slanted isentropes and therefore can become detached from the surface (Orbe et al. 2015; Bozem et al. 2019). Conversely, the air associated with warm and moist intrusions from lower latitudes can only have a strong impact on the SEB in cases where it displaces the cold lower tropospheric air instead of ascending over it (Messori et al. 2018). In cases where the air from lower latitudes ascends, other—more local—processes must account for the warming (and moistening) of the near-surface air. In fact, trajectory studies indicate that wintertime near-surface warm extremes in the high Arctic are often caused by air that is of Arctic origin and has been subject to substantial diabatic heating via turbulent fluxes or subsidence warming en route to the extreme (Binder et al. 2017; Papritz 2020).

In the present study, we focus on the events with the most extreme SEB anomalies during extended winters 1979–2020 as defined by Murto et al. (2023), so-called SEB life cycle events (LCEs). Major goals are to shed light on the origin of the air in the entire tropospheric column and to explore the relative importance of remote and local processes for generating positive θ* characteristic of these events. More specifically, we address the following questions:

  • What is the geographical origin of the air at different levels in the tropospheric column?

  • Where and when do potential temperature anomalies (θ*) form along the pathway of the air associated with the SEB events?

  • How much do horizontal and vertical transports, as well as diabatic heating, contribute to the formation of θ*?

  • What is the role of local airmass transformations for the generation of θ*?

For that purpose, we develop a trajectory-based method that allows quantifying the contributions of horizontal transport, that is, the advection of the air from climatologically warmer regions into the Arctic, vertical motion, and diabatic heating to the generation of θ*, similar to Röthlisberger and Papritz (2023a,b), who formulated such budgets using temperature instead of potential temperature. The study is structured as follows: section 2 introduces the methodology including the θ* decomposition, which is followed in section 3 by two case studies of LCEs—one occurring in the Atlantic sector and the other one in the Pacific—that illustrate key characteristics of the warming processes that cause θ* at different levels. We then present results for all LCEs in section 4 and end with a discussion of our results and conclusions in section 5.

2. Data and methods

a. Data

The study is based on ERA5, the latest reanalysis dataset produced by the European Centre for Medium-Range Weather Forecasts (ECMWF; Hersbach et al. 2020). We use hourly analysis fields on model levels for extended winters (November–March) between 1979 and 2020, with the first winter starting in November 1979. Fields have been interpolated in the horizontal to a 0.5° × 0.5° longitude–latitude grid.

b. SEB life cycle events

We use multiday life cycle events (LCEs) of extreme positive SEB anomalies over Arctic sea ice as introduced by Murto et al. (2023). In the following, we briefly outline the approach for the identification of LCEs; for details we refer to the aforementioned study. Each LCE consists of a sequence of coherent regions of extreme positive SEB anomalies, hereafter referred to as SEB anomaly patches. These SEB anomalies are defined relative to an SEB climatology that takes the seasonal cycle and the long-term trend into account, which ensures a uniform distribution of events across the study period. SEB anomaly patches are then defined as spatially coherent areas where the standardized SEB anomalies exceed a seasonally varying 95th-percentile threshold. That way, only the most extreme SEB anomalies are selected. These patches are then tracked in space and time to form LCEs. This is done iteratively starting from the patches with the largest mean SEB anomalies and area. Specifically, a candidate patch is connected to a track if it satisfies the following two criteria with respect to a parent patch that has already been connected to the track. First, the two patches need to be spatially close to each other; that is, the centers of mass must be within two equivalent radii of the parent patch, whereby the equivalent radius of a patch corresponds to the radius of a circle with the same area as the patch. Second, the centroids of air parcels arriving at 200 hPa above ground level (AGL) at the two patches must be located within 1500 km from each other two days prior to the occurrence of the patches. Following this procedure, we identify 142 LCEs, 72 of which could be attributed to an Atlantic airmass origin and 60 to a Pacific origin based on the location of the trajectory centroids two days prior to arrival (cf. Figs. 2a and 5e in Murto et al. 2023). For simplicity, LCEs with airmass origin over Siberia (2 events) or North America (8 events) are not considered here. The LCEs have a typical lifetime of 3–7 days and affect an area of several 106 km2. If not mentioned otherwise, the analysis presented in this study includes all SEB anomaly patches attributed to all or selected LCEs. Note that even though we do not explicitly consider the time evolution of LCEs (i.e., the life cycle of events from onset to decay), we still use the term LCE for consistency with Murto et al. (2023).

c. Air-parcel trajectories

For the analysis of the origin of the air in the tropospheric column collocated with LCEs and for quantifying the contributions of different processes to the generation of θ* associated with these events, we compute kinematic backward trajectories from the tropospheric column collocated with the SEB anomaly patches using the Lagrangian Analysis Tool (LAGRANTO; Wernli and Davies 1997; Sprenger and Wernli 2015). For each SEB anomaly patch, trajectories are initialized on an equidistant grid with 50 km grid spacing in the horizontal and at 50, 100, 200, 300, 400, 500, and 600 hPa AGL in the vertical.1 Initialization times of the trajectories are at 0300, 0900, 1500, and 2100 UTC on the day when the SEB anomaly patch occurs. Trajectories are calculated backward in time for 10 days and trajectory positions (longitude, latitude, pressure) as well as additional variables interpolated to these positions (i.e., potential temperature, potential temperature climatology) are stored in 3-hourly intervals. Note that we use hourly wind fields for the computation of the trajectories.

Complementary to the set of LCE trajectories, an additional set of reference trajectories is computed, which is meant to approximate the climatology of air parcel trajectories in the Arctic (pseudoclimatology). To account for the spatially nonuniform distribution of SEB anomaly patches across the Arctic, reference trajectory starting points are chosen to be identical to those associated with the SEB anomaly patches. For the reference trajectory calculation, we use data from the same calendar day but a randomly chosen year other than the actual year when the patch occurred. To increase the representativeness, this procedure is repeated twice such that the pseudoclimatology is obtained from averaging over twice as many trajectories as contained in the LCE trajectory dataset.

d. Potential temperature anomaly budget

Potential temperature anomalies θ* are given by
θ*=θθc,
where θc denotes the transient potential temperature climatology that takes the seasonal cycle as well as the long-term trend into account. The latter is essential when studying the dynamical processes that cause temperature anomalies in a transient climate, particularly in view of the rapid and nonlinear warming trends observed in the Arctic (Richter-Menge and Druckenmiller 2020). Similar to Messori et al. (2018), the climatology is computed in two steps: first, the time series is smoothed with a 21-day running-mean filter, removing short-term fluctuations but keeping the seasonality intact. Second, for a given date, all days within a centered 9-yr window and with the same calendar day are averaged, smoothing out interannual variations while at the same time retaining most of the nonlinear warming trend. For dates in the first (1979–83) or final (2016–20) 5 years, the averaging is done over the first and final 9 years of the study period, resulting in a climatology with no interannual variations during these years.
Taking the material derivative of Eq. (1) and making use of the thermodynamic energy equation Dθ/Dt=θ˙ yields an equation for the change of θ* along an air parcel trajectory:
Dθ*Dt=θ˙DθcDt.
Subsequent integration over the time interval tα to tβ and expanding the material derivative yields the potential temperature anomaly budget
θ*(tβ,xβ,pβ)=θ*(tα,xα,pα)+tαtβθ˙vhθcωθcpθctdτ.
Here, v and ω denote the horizontal wind and vertical pressure velocity, respectively, and ∇h the horizontal component of the gradient; x denotes the horizontal position, and p is pressure. Hence, θ* at time tβ is given by the initial anomaly at time tα and the time integrated diabatic heating as well as horizontal and vertical transports across the gradient of θc; that is, transport from a climatologically warm to a climatologically cold region leads to a positive θ*. Though defined as a transient climatology, θc changes only slowly in time such that the contribution from the last term is very small and is neglected in the following.
For the practical implementation, the individual terms are computed from 3-hourly incremental changes of θ and θc along the trajectories. Specifically, the contributions from diabatic processes D and vertical transport TV are given by
D=i=1n1θ(ti+1,xi+1,pi+1)θ(ti,xi,pi)=θ(tβ,xβ,pβ)θ(tα,xα,pα) and
TV=i=1n112(pi+1pi)[θcp(ti+1,xi+1,pi+1)+θcp(ti,xi,pi)],
where t1 = tα and tn = tβ and vertical derivatives of θc are approximated by finite differences as
θcp(ti,xi,pi)=θc(ti,xi,pi+Δp2)θc(ti,xi,piΔp2)Δp,
with Δp = 25 hPa. For that purpose, θc is interpolated to (ti,xi,pi±(Δp/2)) in addition to the actual trajectory positions (ti, xi, pi). To compute the contribution from horizontal transport TH we first note that
vhθc=DθcDt+θct+ωθcp.
Since θc varies slowly in time (i.e., it is two to three orders of magnitude smaller than the other terms), we neglect it and, therefore, compute TH as
TH=θc(tα,xα,pα)θc(tβ,xβ,pβ)TV.
This approach is preferable over a direct computation of the horizontal advection term since its spatial discretization using finite differences and approximating its time integral with instantaneous values would lead to unacceptably large numerical errors (see also Röthlisberger and Papritz 2023a,b).

Finally, the diabatic tendencies are further decomposed into contributions from latent heating in the free troposphere, heating in the planetary boundary layer, residual diabatic heating, and diabatic cooling. For that purpose, incremental changes Δθi = θ(ti+1, xi+1, pi+1) − θ(ti, xi, pi) are uniquely attributed to one of these categories according to the following criteria:

  1. Latent heating: Δθi > 0 K and Δqi < 0 g kg−1 h−1 and pipipbl0hPa and Δpi < 0 hPa h−1

  2. Boundary layer heating: Δθi > 0 K and pipipbl>0hPa

  3. Diabatic cooling: Δθi ≤ 0 K

where q is specific humidity and ppbl the pressure at the top of the boundary layer. The latter is interpolated from the diagnostic height of the planetary boundary layer available in the ERA5 archive. The rationale behind the first criterion is that condensation in the free troposphere is associated with ascending air and a decrease of specific humidity. The second criterion captures situations where an air parcel is located in the planetary boundary layer and experiences heating, for example by sensible heat fluxes such as in a marine cold air outbreak. The remaining positive θ changes that are not attributed to the latent heating or boundary layer heating categories are denoted as residual diabatic heating. A caveat of this approach is that each incremental change of θ is uniquely attributed to one category. In reality, however, several physical processes are at work at the same time; for example, longwave radiative cooling will occur concomitant with latent heating partially compensating each other. Note that our use of these categories primarily lies in studying diabatic heating contributions in specific flow situations rather than due to different physical processes. Finally, we define tgenesis as the last time step along the trajectory when θ* has the opposite sign as θ*(t=0h) (or is exactly 0 K). In the evaluation of the θ* budget we will mostly consider the time interval from t = tgenesis to t = 0 h.

e. Interpretation of the θ* budget

The processes generating positive θ* along air parcel trajectories directed toward the Arctic are schematically illustrated in Fig. 2. The transport terms represent the gain of θ* resulting from the transport of air from a region with climatologically high to a region with climatologically low potential temperature. Thereby, we distinguish the crossing of climatological isentropes in the horizontal (dark blue arrow in Fig. 2) and the vertical (light blue arrow in Fig. 2) to separate the effect of transport from lower latitudes into the Arctic from subsidence or ascent. Note that in this framework, flow along slanted climatological isentropes is associated with equal but opposite contributions from horizontal and vertical transport.

Fig. 2.
Fig. 2.

Schematic illustrating processes along air parcel trajectories causing positive potential temperature anomalies (θ*) in the Arctic. The transparent gray surface indicates a climatological isentrope and arrows illustrate various processes generating positive θ*, i.e., horizontal and vertical transports across climatological isentropes (dark and light blue, respectively), as well as diabatic processes including latent heat release (red) and turbulent fluxes into the boundary layer (orange).

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

For quantitatively understanding the relevance of horizontal transport, it provides insight to consider Fig. 3a, which quantifies the θ* acquired due to horizontal (isobaric) transport at 700 hPa (see Fig. 1 in the online supplemental material for additional levels) from a given geographical location into the polar cap poleward of 80°N assuming no diabatic processes (e.g., radiative cooling and/or mixing occurring along the transport). For instance, θ* of an air parcel moving from the British Isles to the high Arctic increases by around 20 K. The figure clearly reveals the climatological dome of polar cold air. Air parcels crossing this dome in the poleward direction experience a substantial gain in θ*. Furthermore, the pattern also reveals the well-known stationary wave patterns with the polar dome extending farther south over the eastern continents and farther north over the eastern ocean basins. The stationary wave pattern further leads to a zonal component of the gradient over the North Atlantic and to a lesser extent over the North Pacific such that air parcels moving zonally in these regions experience a decrease of θ* due to horizontal transport (cf. with black contours). In addition, the climatological potential temperature contrast between the subpolar oceans and the high Arctic is considerably stronger in the lower than the upper troposphere (cf. supplemental Fig. 1).

Fig. 3.
Fig. 3.

Generation of θ* due to horizontal and vertical transport. (a) θ* acquired by an air parcel transported isobarically at 700 hPa from a given geographical location into the polar cap poleward of 80°N with the 6- and 18-K and the 12-K contours highlighted by thin and thick black lines, respectively. The figure is obtained by subtracting the polar cap (≥80°N) average of the potential temperature climatology θc from θc at a given geographical location. Note that variations of θc within the 80°N polar cap are small as compared to θ* acquired due to transport from lower latitudes into the polar cap. (b) θ* acquired by an air parcel subsiding from 700 to 800 hPa at a fixed geographical location. Both panels show averages over extended winters 1979–2020 and the polar cap 80°N is indicated by a black circle.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

Similarly, Fig. 3b shows the gain in θ* due to subsidence from 700 to 800 hPa at a given geographical location. An air parcel descending from 700 to 800 hPa gains θ* of about 5–7 K due to vertical transport in most of the regions. Variations in the efficiency of subsidence in creating θ* are determined by variations of the climatological stratification, i.e., θc/p. Consequently, subsidence is particularly efficient in creating large θ* along the slopes of Greenland’s ice sheet and the coldest regions over the eastern continents where the climatological stratification is high, whereas it is considerably less efficient over the Nordic seas and near Kamchatka.

Finally, θ* can increase due to diabatic heating such as the release of latent heat, for example as warm and humid midlatitude air impinges on the dome of polar cold air, ascends, and cools, followed by cloud formation (red arrow in Fig. 2), or sensible heat fluxes from the ocean to the atmosphere when Arctic air sweeps over a comparatively warm ocean surface (orange arrow in Fig. 2).

3. Case studies

To illustrate the methodology and obtain a sense of the diversity of airstreams associated with LCEs, we first consider two case studies. For both cases, animations illustrating the evolution of air parcel trajectories reaching the LCE at 50, 200, and 500 hPa AGL and the formation of θ* are provided in the supplemental material.

a. Life cycle event 44

The first case, LCE 44 [numbering according to Murto et al. (2023)] in December 1986, was accompanied by one of the most severe near-surface warm extremes in the Arctic in recent decades (Messori et al. 2018). The affected area extended in a narrow plume from the ice-covered portions of the Greenland Sea across the North Pole toward Siberia [see also Murto et al. (2023) for a discussion of its evolution and characteristics]. The LCE formed in conjunction with a moist-air intrusion from the Atlantic sector that reached the area affected by the LCE (purple hatched in Fig. 4) at 200 hPa AGL and above (see supplemental Fig. 2 and animations in the online supplemental material).

Fig. 4.
Fig. 4.

Case study of LCE 44 showing the positions of trajectories initialized at 50 hPa AGL from all patches associated with the LCE. Maps are shown for lags of (a) −8, (b) −6, (c) −4, and (d) −2 days prior to 1200 UTC 12 Dec 1986, corresponding to the peak of the LCE. Trajectory positions are colored according to pressure. Additionally shown are sea level pressure (gray contours at intervals of 10 hPa; blue and gray shading for values below 990 hPa and above 1030 hPa, respectively) and the 5000 m (blue thin), 5300 m (blue thick), and 5600 m (blue thin) isohypses of 500 hPa geopotential height. The purple hatched area indicates the region affected by the LCE and the solid gray line shows the 60°N latitude circle. Note that only positions from every second trajectory are drawn for better visibility.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

In contrast to the air farther aloft, the near-surface air originated in the Arctic. Figure 4 shows the location of air parcels during the days prior to arriving at 50 hPa AGL at the LCE. Eight days prior to the LCE’s peak, air parcels were mainly located poleward of 60°N at an elevation above 850 hPa (Fig. 4a). Subsequently, they descended and moved westward over open ocean in the Barents Sea (Figs. 4b,c), turned poleward and re-entered the Arctic upon which they moved across the pole in a narrow plume (Fig. 4d). In addition, a few air parcels originated over the Canadian Archipelago and went around the southern tip of Greenland, thus leaving and returning to the Arctic. Furthermore, others descended from above the Greenland ice sheet (Figs. 4b,c). Two days prior to the peak of the LCE, near-surface air parcels have merged into two coherent blobs, one near Fram Strait and the other north of Siberia, already deep within the high Arctic (Fig. 4d). Thus, the near-surface air associated with the exceptional warmth had a very different history than the air farther aloft, which has the characteristic evolution of midlatitude moist-air intrusions (cf. Woods and Caballero 2016). As a result, the mechanisms that caused θ* likely differ between the near-surface air and the air farther aloft.

Figures 5a and 5c show the fractions of trajectories with ttgenesis for trajectories arriving at 50 and 500 hPa AGL and the evolution of θ* (black line). Recall that tgenesis refers to the last time when θ* has the opposite sign than the final anomaly. For most of the trajectories θ* started to form less than 4 days prior to the LCE with a longer tail (i.e., earlier onset of θ*) for lower tropospheric air parcels (gray shading in Figs. 5a,c). For about 50% of the air parcels, the anomalies formed only within 48 h prior to the LCE. Furthermore, the final θ* (at t = 0 h) associated with near-surface air parcels was substantially larger than for air parcels aloft (more than 10 K vs roughly 7 K).

Fig. 5.
Fig. 5.

θ* budgets for LCEs (left) 44 and (right) 137. Shown is the temporal evolution of θ* (black) and of accumulated contributions to θ* from horizontal transport (dark blue), vertical transport (light blue), latent heating (red), boundary layer heating (orange), residual diabatic heating (yellow), and diabatic cooling (purple) for trajectories initialized at (a),(b) 500 and (c),(d) 50 hPa AGL (axis on the right) from all SEB anomaly patches associated with the given LCE. For each trajectory, terms are set to zero for t = tgenesis and accumulated for the period 0 ≥ ttgenesis. The resulting time series are then averaged over all trajectories. Gray bars show the fraction of trajectories with ttgenesis (axis on the left). t = 0 h corresponds to the time of arrival in the LCE and t = tgenesis to the time when θ* was last ≤0 K.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

The temporal evolution of the contributions of the θ* budget terms accumulated over the period from tgenesis to t = 0 h shows that θ* of near-surface air parcels formed mainly through horizontal and vertical transport, as well as diabatic heating in the planetary boundary layer (Fig. 5c), which occurred when the air parcels transited open ocean in the Barents Sea (see the animation in the supplemental material). Note that in this region, horizontal transport is particularly efficient in creating θ* due to the large near-surface climatological temperature contrast between the Barents Sea and the high Arctic (supplemental Figs. 1a,b). Substantial amounts of these positive contributions were balanced by diabatic cooling. Similar to near-surface air parcels, air parcels aloft gained θ* primarily via horizontal transport, followed by latent heating (Fig. 5a). Since latent heating is largely associated with ascent and adiabatic cooling, the latent heating contribution was entirely consumed by a negative contribution from vertical transport.

b. Life cycle event 137

The second case, LCE 137, was associated with meridional transport from the Pacific sector aloft and a persistent anticyclone within the Arctic (Fig. 6). Note that the anticyclone was confined to the lower troposphere with no upper-level anticyclone present (cf. SLP contours and 500 hPa isohypses in Fig. 6). With a very persistent anticyclone located within the Arctic, this event differs from the typical LCE as described in Murto et al. (2023), yet it illustrates how diverse the processes can be that lead to extreme SEB anomalies over Arctic sea ice.

Fig. 6.
Fig. 6.

As in Fig. 4, but for LCE 137 with lags relative to the peak of that LCE at 1200 UTC 3 Jan 1994.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

Air parcels aloft originated in the Pacific sector and were episodically injected into the high Arctic associated with transient extratropical cyclones and tropospheric cutoffs (see supplemental Fig. 3 and supplemental animations). In contrast, the bulk of near-surface air parcels was located within the Arctic 8 days prior to the peak of the LCE (Fig. 6a). Three bundles of air parcels can be discerned: one collocated with the anticyclone at roughly 700 hPa, one over Siberia also at midtropospheric elevations, and one over and north of Alaska located at low altitudes. The air parcels associated with the latter two bundles were gradually swept up by the anticyclone over the Beaufort Sea, in which they remained trapped and subsided (Figs. 6b–d).

Considering the θ* budget, Figs. 5b and 5d show that θ* formed over a longer time period compared to LCE 44, as evident from the longer tail of the fraction of trajectories with ttgenesis. Furthermore, final θ* was almost twice as large aloft than near the surface. While aloft the θ* budget terms show a qualitatively similar evolution as in the case of LCE 44 (Figs. 5a,b), positive contributions for near-surface air parcels were dominated by vertical transport, while horizontal transport was less important, and diabatic warming was negligible (Fig. 5d). Thus, in this case, the near-surface air acquired its θ* primarily via subsidence in the anticyclone. A substantial portion of the θ* increase due to vertical transport was, however, compensated by diabatic cooling.

c. Summary of case studies

Comparing the evolution of the θ* budget terms for air parcels near the surface and aloft for the two case studies reveals several interesting characteristics. First, either horizontal or vertical transport dominates the positive contributions at all levels. In particular, this also applies to the near-surface air with Arctic origin, which is required to reach an area with substantially higher θc than in the high Arctic before θ* starts to form. Second, θ* arises as a residual between opposing processes, whose individual contributions typically exceed the magnitude of θ*. Third, heating in the planetary boundary layer can contribute to the warming of the near-surface air, but its contribution to the final θ* remains smaller than that from transport. That said, heating in the planetary boundary layer, especially due to surface turbulent heat fluxes over open ocean, might be more important for transforming the originally Arctic near-surface air while it moves into a region with warmer θc before θ* starts to form, as well as for eroding an initially negative θ*. This possibility will be explored to a greater extent in the climatological part of the study.

4. Results for all LCEs

In this section, we apply the previous methodology to all Atlantic and Pacific LCEs.

a. Air parcel origin and evolution of θ*

Before turning attention to the climatological θ* budgets, we first consider when and where positive values of θ* form. The so obtained temporal and spatial scales of θ* formation will be helpful in interpreting the θ* budgets. Figure 7 shows the temporal evolution of θ* and histograms of tgenesis for trajectories initialized at 500 and 50 hPa AGL.

Fig. 7.
Fig. 7.

Histograms of tgenesis for all Atlantic and Pacific LCEs (gray bars) and temporal evolution of mean (solid lines) and interquartile range (shading) of θ* for Atlantic (blue) and Pacific (red) LCEs for trajectories initialized at (a) 500 and (b) 50 hPa AGL. Dashed black lines show the 75th percentile of θ* associated with climatological trajectories (see section 2c). Bars are shown every 6 h and the leftmost bar includes trajectories with tgenesis ≤ −240 h.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

Near-surface air parcels tend to be anomalously cold 10 days prior to arrival at the LCE. The initial cold anomaly is slightly more pronounced for Atlantic than Pacific events and the anomalies form over a shorter time period in the former group. The anomalies emerge between t = −72 and −48 h in the mean (Fig. 7b) and they exceed ordinary variations of θ* as described by the 75th percentile of θ* along reference trajectories (dashed line) at about t = −48 h. Hence, air parcels become unusually warm only 2 days prior to arrival. The fraction of air parcels for which tgenesis ≤ −240 h (i.e., for which the generation of θ* cannot be fully captured by 10-day backward trajectories) is acceptably small (3%–4% of all trajectories). Considering the distribution of tgenesis further reveals a long tail indicating that for some trajectories the formation of the anomalies occurs over considerably longer time periods than in the mean, such as in the case of LCE 137 discussed in section 3b. Unlike lower tropospheric trajectories, mid- and upper tropospheric trajectories associated with Atlantic LCEs feature only weak initial cold anomalies and none for Pacific LCEs (Fig. 7a). Furthermore, anomalies tend to form slightly earlier and the distribution of tgenesis is wider with an earlier peak.

Complementary to the temporal scales over which θ* forms, we consider the geographical origin of LCE air parcels and their location at t = tgenesis. For that purpose, Fig. 8 shows trajectory densities at t = −240 h and at t = tgenesis for trajectories associated with Atlantic and Pacific LCEs and initialized at 50, 200, and 500 hPa AGL. At t = −240 h, near-surface air parcels are mainly located in the Arctic, including the high Arctic, Eurasia, and the Canadian Archipelago (Figs. 8a,d) consistent with the two case studies. Only a small portion of air parcels originate over the North Atlantic and North Pacific Oceans.

Fig. 8.
Fig. 8.

Spatial distribution of trajectories initialized at (a),(d) 50, (b),(e) 200, and (c),(f) 500 hPa AGL associated with (a)–(c) Atlantic and (d)–(f) Pacific LCEs. Trajectory positions are shown for t = −240 h [red hatched; contours at 1% and 2% (106 km2)−1], t = tgenesis (shading), and t = 0 h [black; contour at 1% (106 km2)−1]. Further shown in (b) and (e) are the 6-, 12-, and 18-K contours of the gain of θ* due to horizontal transport at 700 hPa from a given geographical location into the 80°N polar cap (thin and thick black lines; as in Fig. 3a).

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

For Atlantic LCEs and near-surface air parcels, the anomaly starts to form mainly over open ocean in the Nordic seas and the Barents Sea, while for Pacific LCEs anomaly genesis occurs in the northern North Pacific, east of Kamchatka and near the Bering Strait. All these regions are climatologically warmer than the predominant origin of the air parcels (i.e., their location at t = −240 h) as well as the regions with Arctic sea ice (see black contours in Figs. 8b,e and 3a). It is worth noting that trajectory densities at t = tgenesis are sharply delimited by the sea ice edge and landmasses. This hints at the relevance of diabatic heating via turbulent surface heat fluxes for the erosion of the originally negative θ* and counteracting the effect of horizontal transport from the cold Arctic origin to the climatologically warmer ocean regions at t = tgenesis.

At t = −240 h, air parcels initialized at 200 and 500 hPa AGL are located only 5°–15° latitude equatorward of the latitude at which they arrive at the LCE (supplemental Fig. 4b). However, they undergo a stronger equatorward excursion (supplemental Fig. 4b) and the great circle distance between origin and arrival is considerably larger as compared to near-surface air parcels (supplemental Fig. 4a) since the former originate predominantly upstream of the Atlantic and Pacific gateways into the Arctic. More specifically, air parcels originate primarily over Canada, Labrador and Greenland for Atlantic LCEs (Figs. 8b,c) and eastern Eurasia for Pacific LCEs (Figs. 8e,f). During the first 5–7 days, they tend to move equatorward, followed by a steep poleward trajectory during the final days (supplemental Fig. 4b). The locations of air parcels at t = tgenesis arriving at 200 and 500 hPa AGL lie downstream of the southern inflection point of the trajectory in the respective sector of the Arctic (Figs. 8b,c) and the North Atlantic and the North Pacific ocean basins (Figs. 8e,f), respectively.

In line with the case studies, the different origins of near-surface air parcels and air parcels further aloft suggest that LCEs do not result from deep vertically coherent injections of midlatitude air into the Arctic. Instead, they are systematically caused by the vertical superposition of originally Arctic air in the lowermost troposphere and midlatitude air aloft.

This finding is further corroborated by the evolution of the great circle and latitudinal distances between the centroids of trajectories initialized near the surface and further aloft (supplemental Figs. 4c,d). The great circle distance decreases from originally more than 2500 km between near-surface air parcels and air parcels at 500 hPa AGL to below 500 km only between t = −36 and 0 h (supplemental Fig. 4a). Thus, the coherent transport of the airstreams across the tropospheric column is only a short phase right before the occurrence of an LCE. Likewise, the evolution of the latitudinal distance shows a similar rapid decrease during the final 2 days (supplemental Fig. 4b). Furthermore, the latitudinal distance between near-surface air parcels and air parcels aloft has pronounced peaks at about t = −120 and −72 h for Atlantic and Pacific LCEs, respectively, reflecting that air parcels aloft first move from their origin equatorward before turning poleward (supplemental Figs. 4b,d; also compare Figs. 8c,f).

b. θ* budgets

Given the different origins of air parcels and the location where θ* starts to form, we now aim to shed light on the processes that cause θ*. For that purpose, we first consider the θ* budget for the reference trajectories (see section 2c). There, a balance exists between positive and negative contributions to θ* (supplemental Fig. 5a). This balance is essentially set by the mean circulation in the Arctic: midlatitude air is transported poleward, whereby it ascends into the mid- and upper troposphere (negative contribution from vertical transport), indicating that the flow is at least partially along slanted climatological isentropes. Adiabatic cooling during ascent facilitates condensation, which in turn compensates part of the negative contribution from vertical transport. From the mid- and upper troposphere, the air then subsides within the Arctic into the lower troposphere while it cools radiatively such that the first-order balance is between diabatic cooling and vertical transport. Any θ* forming along LCE trajectories must, thus, result from departures from this balance between positive and negative contributions to θ* along reference trajectories (supplemental Figs. 5b,c).

Figures 9a and 9b show the magnitude of θ* at the time of the LCEs (t = 0 h; diamonds) and the θ* budget accumulated over the time period t = tgenesis to t = 0 h (bars) for Atlantic and Pacific LCEs, comprising on average about 3 to 4 days (see section 4a). This reveals that θ* peaks in the lower troposphere with more than 10 K. In the case of Atlantic LCEs, θ* declines with altitude to 6–7 K at 600 hPa AGL, whereas the decrease is more modest for Pacific LCEs (see also Fig. 7a in Murto et al. 2023).

Fig. 9.
Fig. 9.

(a),(b) Accumulated contributions of horizontal transport (dark blue), vertical transport (light blue), latent heating (red), boundary layer heating (orange), residual diabatic heating (yellow), and diabatic cooling (purple) to the final θ* (diamonds) for all (a) Atlantic and (b) Pacific LCEs. For each trajectory, the terms are accumulated from t = tgenesis to t = 0 h. (c),(d) As in (a) and (b), but for the difference between trajectories with the 20% largest (upper quintile) and smallest (lower quintile) θ* at t = 0 h. Quintiles are defined by level.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

For LCEs, horizontal transport dominates the positive contributions at all levels. This is consistent with air parcel locations at t = tgenesis (Fig. 8), which implies that air parcels must cross the climatological dome of cold air as they track poleward (see also the 12-K contour shown in Figs. 8b,e). This crossing of the climatological dome of polar cold air occurs during the final ∼48 h prior to an LCE when lower and upper tropospheric air parcels approach each other and move fairly coherently poleward (cf. supplemental Fig. 4a). Similar to the reference trajectories, vertical transport contributes positively in the lower and negatively in the upper troposphere reflecting descent and ascent, respectively, and with stronger positive contributions in the case of Pacific LCEs compared to Atlantic LCEs. Further positive contributions stem from latent heating in the mid- and upper troposphere, as well as from heating in the planetary boundary layer, the latter more for Atlantic than Pacific LCEs. Finally, diabatic cooling balances a substantial part—but not all—of the positive contributions across all levels, with larger magnitudes in the lower troposphere. Comparing with the reference trajectories, this suggests a decisive role for horizontal transport for the formation of θ* associated with LCEs.

Insight into the spatial variability of the θ* budgets for LCEs is obtained from Lagrangian forward projections of the individual θ* budget terms (Fig. 10 and supplemental Fig. 6). Lagrangian forward projections are computed by first assigning a given characteristic quantity of a trajectory, such as the θ* budget contributions, to the trajectory location at t = 0 h (Liniger and Davies 2003), and then averaging over all values assigned to the same location. Thus, the maps in Fig. 10 show, for a given location and altitude, the average magnitude of θ* and how much a particular process contributes to its formation. This shows that lower tropospheric θ* is largest near the sea ice edge, reaching up to 16 K, and declines with fetch over sea ice (Figs. 10a,i). The reasons for the existence of this gradient are twofold. First, the horizontal transport contribution declines as the air traverses the high Arctic and eventually moves across climatological isentropes from cold to warm (Figs. 10b,f,j,n). Second, diabatic contributions become more negative with fetch (Figs. 10d,l) due to prolonged diabatic cooling and the absence of any notable diabatic warming (supplemental Figs. 6d,h,l,p). This gradient of θ* amplitude is much less pronounced in the upper troposphere (Figs. 10e,m), primarily due to weaker diabatic cooling contributions (Figs. 10h,p) and a reduced or even reversed gradient in the contributions of horizontal transport across the Arctic (Figs. 10f,n). Finally, positive contributions to lower tropospheric θ* due to vertical transport tend to peak along the coastlines and orography, particularly north of Alaska and Greenland (Figs. 10c,k), while in the upper troposphere, vertical transport contributions are almost uniformly negative (Figs. 10g,o).

Fig. 10.
Fig. 10.

Lagrangian forward projections of θ* budget terms for (a)–(h) Atlantic and (i)–(p) Pacific LCEs and trajectories initialized at (a)–(d) and (i)–(l) 50 and (e)–(h) and (m)–(p) 500 hPa AGL. Shown are (first column) θ* at t = 0 h, and the contributions of (second column) horizontal transport, (third column) vertical transport, and (fourth column) diabatic processes to θ* accumulated over ttgenesis. The black contour shows the spatial distribution of trajectories at t = 0 h [contour at 1% (106 km2)−1].

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

c. Peculiarities of SEB trajectories

As for the case studies, the θ* budgets again reveal a clear compensation between the contributions of different processes such that the final θ* emerges as a residual, which is notably smaller than the contributions from the individual processes. Compensation results from the anticorrelation between the contributions from certain physical processes: Latent heat release is closely linked to ascent, such that in the upper troposphere positive contributions from latent heat release are often counteracted by negative contributions due to vertical transport. Furthermore, diabatic cooling often goes along with subsidence and as a result vertical transport compensates some of the diabatic cooling in the lower troposphere. Nevertheless, this decomposition of θ* by processes does not provide quantitative information about which processes have unusual contributions as compared to the average Arctic air parcel, which does not obtain a strongly positive θ*.

Some insight into the unusualness can be gained, however, by stratifying trajectories according to the magnitude of the final anomaly and comparing the contributions. We do so by considering the differences of the contributions of LCE trajectories with the 20% largest (upper quintile) and smallest (lower quintile) θ* at t = 0 h among all LCE trajectories at a given level (Figs. 9c,d). Note the pronounced difference of θ* at t = 0 h of 10–17 K between the two sets (diamonds). Even though some cancellation between terms remains, the cancellation is clearly less than the magnitude of the anomalies, with the exception of the most elevated air parcels.

Across levels, horizontal transport is the dominant term contributing additional θ* (Figs. 9c,d). For upper tropospheric trajectories, the figures further reveal more negative contributions from vertical transport for trajectories with large θ*. Hence, upper tropospheric trajectories with a large θ* (upper quintile) ascend more since tgenesis than trajectories with a weaker θ* (lower quintile). Related to the enhanced ascent, the former also experience more condensation, as reflected by additional positive contributions from latent heating. In the lower troposphere, there is much less cancellation and the additional θ* is mainly caused by stronger horizontal transport contributions, and in the case of Pacific LCEs also vertical transport. The latter is likely linked to anomalously high frequency of blocking over Alaska preceding Pacific LCEs and associated subsidence (supplemental Fig. 7d). Interestingly, also diabatic cooling is generally more intense for warmer trajectories, which on one hand could be due to stronger longwave emission resulting from higher temperature and water vapor content, and on the other hand also a longer timespan over which the anomaly forms.

In summary, this analysis emphasizes the relevance of horizontal transport across the climatological horizontal temperature gradient associated with the polar dome of cold air for the generation of the θ* between t = tgenesis and t = 0 h. Other processes contribute less to the formation of θ* and they can enhance or oppose the formation of θ* depending on the vertical level and geographical region.

d. Role of airmass transformations en route

Even though the θ* changes prior to tgenesis do not contribute to the final θ*, their analysis is nevertheless insightful for understanding the chain of events that leads to large positive near-surface θ* during LCEs for several reasons. First of all, near-surface air parcels tend to have a slightly negative θ* 10 days prior to an LCE (Fig. 7). More importantly, though, the Arctic origin of near-surface air parcels (location at t = −240 h; red hatching in Figs. 8a,d) implies that air parcels must first move from a region that features comparable climatological temperatures as the sea ice–covered portion of the Arctic to the climatologically warmer subpolar oceans, where the air parcels preferentially reach a zero θ* for the last time before the LCE (location at t = tgenesis; shading in Figs. 8a,d). Only then can horizontal transport contribute toward a positive θ* at t = 0 h. In addition, radiative cooling, which in the Arctic typically amounts to 1–2 K day−1 (e.g., Papritz 2020), would inevitably contribute to the formation of negative θ*. Consequently, warming processes must compensate for the initial negative θ* and the negative contributions due to transport and radiative cooling in order for the air parcels to reach a zero θ* at t = tgenesis. In the following, we shall shed light on these warming processes prior to t = tgenesis.

For the purpose, Fig. 11 shows the θ* budget accumulated from t = −240 h to t = tgenesis for the entire troposphere, that is, the θ* budget for the time before the final positive anomaly starts to form. Note that the initial θ* is negative but relatively small in comparison to the other budget terms, such that positive and negative contributions nearly balance each other. In line with equatorward transport into climatologically warmer regions, horizontal transport contributions are negative for lower tropospheric air parcels associated with Atlantic LCEs (Fig. 11a). For Pacific LCEs, negative contributions from horizontal transport occur across the entire tropospheric column (Fig. 11b), which is in line with the more poleward origin of air parcels aloft as compared to Atlantic LCEs (see supplemental Fig. 4b). For LCEs in both regions, vertical transport and diabatic heating contribute positively and, hence, compensate negative contributions due to diabatic cooling and horizontal transport (Figs. 11a,b).

Fig. 11.
Fig. 11.

θ* budget as in Figs. 9a and 9b, but for terms accumulated from t = 240 h to t = tgenesis. In this case the trajectories obtain a nonzero initial anomaly (gray).

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

While at upper levels additional diabatic heating is mainly due to latent heating, heating in the planetary boundary layer becomes increasingly important toward the lower troposphere, more so for Atlantic than for Pacific LCEs. Given that at t = tgenesis the near-surface air is preferentially located over open ocean (Figs. 8a,d), the heating likely occurs in a marine cold air outbreak type of flow, where cold and dry Arctic air is advected over a comparatively warm ocean, resulting in upward surface sensible and latent heat fluxes. The relevance of cold air outbreaks is further corroborated by the geographical location of the maximum 24-hourly boundary layer heating (Fig. 12).

Fig. 12.
Fig. 12.

Spatial distribution of trajectories at the time of maximum 24-hourly diabatic heating in the planetary boundary layer during their 10-day evolution for trajectories initialized between 50 and 200 hPa AGL associated with (a) Atlantic and (b) Pacific LCEs. The 50% sea ice concentration contour is indicated by a solid black line.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

For Atlantic LCEs, several maxima are apparent: the Labrador Sea, the Greenland Sea south of Fram Strait, and most notably the Barents Sea (Fig. 12a), while for Pacific LCEs the maximum heating occurs in the Sea of Okhotsk and the Bering Sea (Fig. 12b). These regions are well known for cold air outbreaks (Fletcher et al. 2016; Papritz and Spengler 2017). Furthermore, they are fairly close to the location of air parcels at t = tgenesis (Figs. 8a,d), namely the last time when they transition from negative to positive θ*.

Provided that near-surface air parcels associated with LCEs are of Arctic origin and anomalously cold initially, we conclude that in the absence of such diabatic heating, diabatic cooling and equatorward transport would inevitably lead to more negative θ* along the trajectories, preventing them from attaining a zero θ* at t = tgenesis, let alone a strong positive θ* during the LCEs. Such airmass transformations can, thus, be seen as a precursor and prerequisite to the formation of a positive near-surface θ* by horizontal transport associated with LCEs. Despite this, the budget shown in Fig. 11 is not a unique feature of near-surface trajectories associated with LCEs but instead appears as a frequent characteristic of trajectories over subpolar oceans (not shown).

5. Conclusions

One of the most prominent features of extreme wintertime positive SEB anomalies over Arctic sea ice is the presence of potential temperature anomalies (θ*) on the order of +10 K across the tropospheric column. In this study, we used kinematic backward trajectories to explore the origin of the air and the airmass transformations leading to the positive θ* during 132 life cycle events of extreme Arctic SEB anomalies (LCEs; see Murto et al. 2023). We focused on LCEs for which air parcels at 200 hPa AGL originate in the Atlantic and Pacific sectors. Based on the kinematic trajectories, we introduced a method to quantify the contributions of horizontal and vertical transport across the climatological gradient of potential temperature, as well as diabatic processes such as latent heat release and heating in the planetary boundary layer to θ*.

a. Synthesis

Synthesizing the results from the trajectory analyses, a three-dimensional conceptual picture of the airstreams constituting the anomalously warm tropospheric column during LCEs and their characteristic thermodynamic evolution emerges. A robust feature of Atlantic and Pacific LCEs is that air arriving near the surface (i.e., below 200 hPa AGL) has Arctic origin, whereas the air farther aloft originates at midlatitudes and is transported over larger distances. To simplify the discussion in the following, the former will be referred to as near-surface air and the latter as air aloft. Typical representations of these airstreams are schematically depicted in Fig. 13. Regarding the questions posed at the end of the introduction, our conclusions are as follows:

  1. What is the geographical origin of the air at different levels in the tropospheric column? Ten days prior to an LCE, air aloft is located at midlatitudes, predominantly in the lower troposphere, a feature that applies equally to Atlantic and Pacific LCEs. Typically, the air associated with Atlantic LCEs is transported from the North American continent across the North Atlantic and ascends as it moves poleward east of Greenland (upper airstream in Fig. 13 left). Similarly, the air associated with Pacific LCEs originates over Eurasia and ascends over the central North Pacific (upper airstream in Fig. 13 right). In contrast to the air aloft, the near-surface air originates in the Arctic and undergoes modest equatorward excursions en route, though remaining at subpolar latitudes. Thereby, the air traverses open ocean areas in the Barents Sea and the Nordic seas (Atlantic LCEs; lower airstream in Fig. 13 left), as well as east of Kamchatka and south of Bering Strait (Pacific LCEs; lower airstream in Fig. 13 right).

Fig. 13.
Fig. 13.

Schematic illustrating typical pathways of near-surface air and air aloft (see text) associated with LCEs in the (left) Atlantic and (right) Pacific sectors. Airstreams are colored according to θ* from blue (negative θ*) to red (positive θ*) and arrows indicate where horizontal transport (dark blue), vertical transport (light blue), latent heating (red), and boundary layer heating (orange) contribute to the generation of θ*. Furthermore, black dots indicate the typical location of trajectories at t = tgenesis (i.e., the time when θ* was last 0 K), corresponding approximately to 2–4 days before arrival at the LCE. The brown dashed contour shows the sea ice edge.

Citation: Journal of Climate 36, 21; 10.1175/JCLI-D-22-0883.1

The latitudinal distance between near-surface air and air aloft is maximized between 3 and 5 days prior to the occurrence of an LCE when the air aloft is at its southernmost latitude and the near-surface air is still deep within the Arctic (cf. black dots in Fig. 13). The airstreams join over the subpolar and polar oceans and vertically coherent transport is limited to the final 1–2 days as the air is transiting through the principal Arctic gateways (i.e., Fram Strait and Bering Strait).

  1. 2) Where and when do potential temperature anomalies (θ*) form along the pathway of the air associated with the SEB events? Across levels, positive θ* forms over a time period of 2–4 days prior to an LCE with earlier onset of θ* formation for air aloft. At the time when the air last obtained a zero θ*, near-surface air is located over the subpolar and polar oceans, whereas air aloft is located closer to or at midlatitudes over the North Atlantic and North Pacific Oceans (black dots in Fig. 13).
  2. 3) How much do horizontal and vertical transports, as well as diabatic heating contribute to the formation of θ*? The θ* commonly forms as a residual between opposing processes, whose individual contributions exceed the final θ*. Horizontal transport and diabatic cooling—mostly due to radiation—have the largest magnitude of all terms in the θ* budget. Horizontal transport contributes positively across levels, whereas vertical transport has positive contributions for near-surface air due to descent and negative contributions for air aloft due to ascent. Latent heating and boundary layer heating have notable positive contributions above and below 200 hPa AGL, respectively.

Considering the differences in the contributions between air parcels with the 20% largest and smallest θ* emphasizes rapid horizontal transport from climatologically warmer to colder regions as the key process causing larger θ* at all levels. Latent heat release provides additional positive contributions to θ* of air aloft, which are opposed by more negative contributions of comparable magnitude due additional ascent. Subsidence can be an important cause for large θ* of near-surface air, especially for Pacific LCEs (see LCE 137). The more important role of vertical transport for Pacific as compared to Atlantic LCEs is consistent with the more frequent occurrence of blocking in the former case.

  1. 4) What is the role of local airmass transformations for the generation of θ*? Given that near-surface air has a modest cold anomaly 10 days before an LCE occurs and that it originates in a region that is climatologically as cold or colder than the high Arctic where the LCE occurs, the air has to adjust to warmer climatological temperatures as found over the subpolar and polar oceans before it can attain a positive θ* via horizontal transport. Such an adjustment typically occurs in a marine cold air outbreak, where the air is warmed by upward turbulent heat fluxes from the ocean, or via vertical transport across the climatological potential temperature gradient. These airmass transformations can be seen as a necessary but not unique precursor for LCEs.

b. Concluding remarks

Based on our results, the view that the most extreme SEB anomalies over Arctic sea ice are associated with vertically coherent injections of warm and humid midlatitude air appears oversimplified. Instead, our findings show that warm and humid air from midlatitudes ascends and joins forces with originally Arctic near-surface air. Thereby, the air at different levels moves coherently only during the final 1–2 days. This is consistent with the idea that poleward flowing air tends to ascend along slanted moist isentropes over the polar dome (e.g., Orbe et al. 2015; Bozem et al. 2019), constituting the air aloft, while the air in the polar dome is trapped and recirculates at low altitudes between the high Arctic and the subpolar oceans, providing the near-surface air after diabatic transformation. As a consequence, the θ* spanning the tropospheric column collocated with the most extreme SEB anomalies over Arctic sea ice results from a combination of poleward transport from lower latitudes and local diabatic airmass transformations over the subpolar and polar oceans as well as subsidence within the Arctic.

The importance of local processes for the generation of near-surface temperature extremes over Arctic sea ice has previously been noted by Binder et al. (2017) and Papritz (2020), whereas the systematic Arctic origin of near-surface air during extreme SEB anomalies has not previously been recognized. This finding may be of particular relevance for understanding the aerosol composition during moist-air intrusions into the Arctic, which in turn plays an essential role in cloud formation processes and associated radiative effects (e.g., Dada et al. 2022). Despite the relevance of local warming processes, the driving dynamical processes in many cases may have an origin at lower latitudes. This includes on one hand synoptic weather systems as identified in many previous studies (e.g., Sorteberg and Walsh 2008; Messori et al. 2018; Nygård et al. 2019; Murto et al. 2022) and confirmed here (supplemental Fig. 7), and on the other hand also established teleconnections with the tropics and midlatitudes (e.g., Baggett et al. 2016; Hofsteenge et al. 2022).

Here, we focused on the origin of large positive θ* in the entire tropospheric column, whose presence is one of the most striking characteristics of the air collocated with extreme SEB anomalies. Temperature directly influences surface sensible heat fluxes and downwelling longwave radiation. However, water vapor and liquid clouds also play an essential role for longwave radiative anomalies (Curry et al. 1995; Shupe and Intrieri 2004; Persson et al. 2017; Vargas Zeppetello et al. 2019; Clark et al. 2021), which strongly contribute to extreme SEB anomalies (Murto et al. 2023). In the presence of moisture sources such as over open ocean, we argue that water vapor is in the first place slaved to temperature via the Clausius–Clapeyron relationship such that anomalously warm air will inevitably have higher moisture content. This is confirmed for the LCEs by vertical profiles (Murto et al. 2023; see also supplemental Fig. 8) and justifies our focus on θ*. Furthermore, this mechanism appears to be especially important for originally Arctic near-surface air transiting over open ocean, since in associated marine cold air outbreaks ocean evaporation immediately follows warming by sensible heat fluxes (Papritz and Spengler 2017). Finally, the fact that a significant amount of the downward longwave radiation at the surface is emitted in the lowermost 1 km of the atmosphere (Ohmura 2001; Persson et al. 2017; Vargas Zeppetello et al. 2019) underpins the relevance of the near-surface air, and thus, the role of local processes for extreme SEB anomalies over Arctic sea ice (see also supplemental Fig. 8).

In summary, these findings suggest that intrusions of midlatitude air into the Arctic alone are not the single cause of collocated SEB anomalies over Arctic sea ice. Rather, the large-scale flow patterns are such that midlatitude air aloft coincides with anomalously warm, humid, and cloudy near-surface air of Arctic origin, which has been transformed by local processes. This results in a vertically deep intrusion of warm and moist air with the near-surface air likely exhibiting the strongest forcing on the SEB.

1

This corresponds to approximately 950, 900, 800, 700, 600, 500, and 400 hPa.

Acknowledgments.

We thank three anonymous reviewers for their insightful and attentive comments that helped to improve the manuscript. Furthermore, we acknowledge the ECMWF for providing access to the ERA5 data set and Michael Sprenger (ETH Zürich) for his continuous support with LAGRANTO. SM was supported by Knut och Alice Wallenbergs Stiftelse, Grant 2016-0024, and GM and MR by the European Union’s H2020 research and innovation programme under ERC Grants 948309 (CENÆ project) and 787652 (INTEXseas project), respectively. Python, the open-source software package R (http://www.r-project.org/), and the NCAR Command Language (http://www.ncl.ucar.edu/) have been used for data analysis and visualization.

Data availability statement.

The ERA5 data can be downloaded from the Copernicus Climate Service (https://climate.copernicus.eu/climate-reanalysis).

REFERENCES

  • Baggett, C., S. Lee, and S. Feldstein, 2016: An investigation of the presence of atmospheric rivers over the North Pacific during planetary-scale wave life cycles and their role in Arctic warming. J. Atmos. Sci., 73, 43294347, https://doi.org/10.1175/JAS-D-16-0033.1.

    • Search Google Scholar
    • Export Citation
  • Binder, H., M. Boettcher, C. M. Grams, H. Joos, S. Pfahl, and H. Wernli, 2017: Exceptional air mass transport and dynamical drivers of an extreme wintertime Arctic warm event. Geophys. Res. Lett., 44, 12 02812 036, https://doi.org/10.1002/2017GL075841.

    • Search Google Scholar
    • Export Citation
  • Boisvert, L. N., A. A. Petty, and J. C. Stroeve, 2016: The impact of the extreme winter 2015/16 Arctic cyclone on the Barents–Kara Seas. Mon. Wea. Rev., 144, 42794287, https://doi.org/10.1175/MWR-D-16-0234.1.

    • Search Google Scholar
    • Export Citation
  • Bozem, H., and Coauthors, 2019: Characterization of transport regimes and the polar dome during Arctic spring and summer using in-situ aircraft measurements. Atmos. Chem. Phys., 19, 15 04915 071, https://doi.org/10.5194/acp-19-15049-2019.

    • Search Google Scholar
    • Export Citation
  • Clark, J. P., E. E. Clothiaux, S. B. Feldstein, and S. Lee, 2021: Drivers of global clear sky surface downwelling longwave irradiance trends from 1984 to 2017. Geophys. Res. Lett., 48, e2021GL093961, https://doi.org/10.1029/2021GL093961.

    • Search Google Scholar
    • Export Citation
  • Curry, J. A., J. L. Schramm, M. C. Serreze, and E. E. Ebert, 1995: Water vapor feedback over the Arctic Ocean. J. Geophys. Res., 100, 14 22314 229, https://doi.org/10.1029/95JD00824.

    • Search Google Scholar
    • Export Citation
  • Dada, L., and Coauthors, 2022: A central Arctic extreme aerosol event triggered by a warm air-mass intrusion. Nat. Commun., 13, 5290, https://doi.org/10.1038/s41467-022-32872-2.

    • Search Google Scholar
    • Export Citation
  • Doyle, J. G., G. Lesins, C. P. Thackray, C. Perro, G. J. Nott, T. J. Duck, R. Damoah, and J. R. Drummond, 2011: Water vapor intrusions into the high Arctic during winter. Geophys. Res. Lett., 38, L12806, https://doi.org/10.1029/2011GL047493.

    • Search Google Scholar
    • Export Citation
  • Dufour, A., O. Zolina, and S. K. Gulev, 2016: Atmospheric moisture transport to the Arctic: Assessment of reanalyses and analysis of transport components. J. Climate, 29, 50615081, https://doi.org/10.1175/JCLI-D-15-0559.1.

    • Search Google Scholar
    • Export Citation
  • Fearon, M. G., J. D. Doyle, D. R. Ryglicki, P. M. Finocchio, and M. Sprenger, 2021: The role of cyclones in moisture transport into the Arctic. Geophys. Res. Lett., 48, e2020GL090353, https://doi.org/10.1029/2020GL090353.

    • Search Google Scholar
    • Export Citation
  • Fletcher, J., S. Mason, and C. Jakob, 2016: The climatology, meteorology, and boundary layer structure of marine cold air outbreaks in both hemispheres. J. Climate, 29, 19992014, https://doi.org/10.1175/JCLI-D-15-0268.1.

    • Search Google Scholar
    • Export Citation
  • Francis, J. A., and E. Hunter, 2006: New insight into the disappearing Arctic sea ice. Eos, Trans. Amer. Geophys. Union, 87, 509511, https://doi.org/10.1029/2006EO460001.

    • Search Google Scholar
    • Export Citation
  • Graversen, R. G., and M. Burtu, 2016: Arctic amplification enhanced by latent energy transport of atmospheric planetary waves. Quart. J. Roy. Meteor. Soc., 142, 20462054, https://doi.org/10.1002/qj.2802.

    • Search Google Scholar
    • Export Citation
  • Henderson, G. R., B. S. Barrett, L. J. Wachowicz, K. S. Mattingly, J. R. Preece, and T. L. Mote, 2021: Local and remote atmospheric circulation drivers of Arctic change: A review. Front. Earth Sci., 9, 709896, https://doi.org/10.3389/feart.2021.709896.

    • Search Google Scholar
    • Export Citation
  • Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 19992049, https://doi.org/10.1002/qj.3803.

    • Search Google Scholar
    • Export Citation
  • Hofsteenge, M. G., R. G. Graversen, J. H. Rydsaa, and Z. Rey, 2022: The impact of atmospheric Rossby waves and cyclones on the Arctic sea ice variability. Climate Dyn., 59, 579594, https://doi.org/10.1007/s00382-022-06145-z.

    • Search Google Scholar
    • Export Citation
  • Kapsch, M.-L., R. G. Graversen, and M. Tjernström, 2013: Springtime atmospheric energy transport and the control of Arctic summer sea-ice extent. Nat. Climate Change, 3, 744748, https://doi.org/10.1038/nclimate1884.

    • Search Google Scholar
    • Export Citation
  • Lee, S., T. Gong, S. B. Feldstein, J. A. Screen, and I. Simmonds, 2017: Revisiting the cause of the 1989–2009 Arctic surface warming using the surface energy budget: Downward infrared radiation dominates the surface fluxes. Geophys. Res. Lett., 44, 10 65410 661, https://doi.org/10.1002/2017GL075375.

    • Search Google Scholar
    • Export Citation
  • Liniger, M. A., and H. C. Davies, 2003: Substructure of a map streamer. Quart. J. Roy. Meteor. Soc., 129, 633651, https://doi.org/10.1256/qj.02.28.

    • Search Google Scholar
    • Export Citation
  • Luo, B., D. Luo, L. Wu, L. Zhong, and I. Simmonds, 2017: Atmospheric circulation patterns which promote winter Arctic sea ice decline. Environ. Res. Lett., 12, 054017, https://doi.org/10.1088/1748-9326/aa69d0.

    • Search Google Scholar
    • Export Citation
  • Messori, G., C. Woods, and R. Caballero, 2018: On the drivers of wintertime temperature extremes in the high Arctic. J. Climate, 31, 15971618, https://doi.org/10.1175/JCLI-D-17-0386.1.

    • Search Google Scholar
    • Export Citation
  • Mortin, J., G. Svensson, R. G. Graversen, M.-L. Kapsch, J. C. Stroeve, and L. N. Boisvert, 2016: Melt onset over Arctic sea ice controlled by atmospheric moisture transport. Geophys. Res. Lett., 43, 66366642, https://doi.org/10.1002/2016GL069330.

    • Search Google Scholar
    • Export Citation
  • Murto, S., R. Caballero, G. Svensson, and L. Papritz, 2022: Interaction between Atlantic cyclones and Eurasian atmospheric blocking drives wintertime warm extremes in the high Arctic. Wea. Climate Dyn., 3, 2144, https://doi.org/10.5194/wcd-3-21-2022.

    • Search Google Scholar
    • Export Citation
  • Murto, S., L. Papritz, G. Messori, R. Caballero, G. Svensson, and H. Wernli, 2023: Extreme surface energy budget anomalies in the high Arctic in winter. J. Climate, 36, 35913609, https://doi.org/10.1175/JCLI-D-22-0209.1.

    • Search Google Scholar
    • Export Citation
  • Naakka, T., T. Nygård, T. Vihma, J. Sedlar, and R. Graversen, 2019: Atmospheric moisture transport between mid-latitudes and the Arctic: Regional, seasonal and vertical distributions. Int. J. Climatol., 39, 28622879, https://doi.org/10.1002/joc.5988.

    • Search Google Scholar
    • Export Citation
  • Nygård, T., R. G. Graversen, P. Uotila, T. Naakka, and T. Vihma, 2019: Strong dependence of wintertime Arctic moisture and cloud distributions on atmospheric large-scale circulation. J. Climate., 32, 87718790, https://doi.org/10.1175/JCLI-D-19-0242.1.

    • Search Google Scholar
    • Export Citation
  • Ogi, M., and J. M. Wallace, 2012: The role of summer surface wind anomalies in the summer Arctic sea ice extent in 2010 and 2011. Geophys. Res. Lett., 39, L09704, https://doi.org/10.1029/2012GL051330.

    • Search Google Scholar
    • Export Citation
  • Ohmura, A., 2001: Physical basis for the temperature-based melt-index method. J. Appl. Meteor. Climatol., 40, 753761, https://doi.org/10.1175/1520-0450(2001)040<0753:PBFTTB>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Orbe, C., P. A. Newman, D. W. Waugh, M. Holzer, L. D. Oman, F. Li, and L. M. Polvani, 2015: Airmass origin in the Arctic. Part I: Seasonality. J. Climate, 28, 49975014, https://doi.org/10.1175/JCLI-D-14-00720.1.

    • Search Google Scholar
    • Export Citation
  • Papritz, L., 2020: Arctic lower-tropospheric warm and cold extremes: Horizontal and vertical transport, diabatic processes, and linkage to synoptic circulation features. J. Climate, 33, 9931016, https://doi.org/10.1175/JCLI-D-19-0638.1.

    • Search Google Scholar
    • Export Citation
  • Papritz, L., and T. Spengler, 2017: A Lagrangian climatology of wintertime cold air outbreaks in the Irminger and Nordic seas and their role in shaping air-sea heat fluxes. J. Climate, 30, 27172737, https://doi.org/10.1175/JCLI-D-16-0605.1.

    • Search Google Scholar
    • Export Citation
  • Papritz, L., and E. Dunn-Sigouin, 2020: What configuration of the atmospheric circulation drives extreme net and total moisture transport into the Arctic. Geophys. Res. Lett., 47, e2020GL089769, https://doi.org/10.1029/2020GL089769.

    • Search Google Scholar
    • Export Citation
  • Persson, P. O. G., M. D. Shupe, D. Perovich, and A. Solomon, 2017: Linking atmospheric synoptic transport, cloud phase, surface energy fluxes, and sea-ice growth: Observations of midwinter SHEBA conditions. Climate Dyn., 49, 13411364, https://doi.org/10.1007/s00382-016-3383-1.

    • Search Google Scholar
    • Export Citation
  • Pithan, F., and Coauthors, 2018: Role of air-mass transformations in exchange between the Arctic and mid-latitudes. Nat. Geosci., 11, 805812, https://doi.org/10.1038/s41561-018-0234-1.

    • Search Google Scholar
    • Export Citation
  • Richter-Menge, J., and M. L. Druckenmiller, 2020: The Arctic [in “State of the Climate in 2019”]. Bull. Amer. Meteor. Soc., 101 (Suppl.), S239S285, https://doi.org/10.1175/2020BAMSStateoftheClimate.1.

    • Search Google Scholar
    • Export Citation
  • Röthlisberger, M., and L. Papritz, 2023a: A global quantification of the physical processes leading to near-surface cold extremes. Geophys. Res. Lett., 50, e2022GL101670, https://doi.org/10.1029/2022GL101670.

    • Search Google Scholar
    • Export Citation
  • Röthlisberger, M., and L. Papritz, 2023b: Quantifying the physical processes leading to atmospheric hot extremes at a global scale. Nat. Geosci., 16, 210216, https://doi.org/10.1038/s41561-023-01126-1.

    • Search Google Scholar
    • Export Citation
  • Rydsaa, J. H., R. G. Graversen, T. I. H. Heiskanen, and P. J. Stoll, 2021: Changes in atmospheric latent energy transport into the Arctic: Planetary versus synoptic scales. Quart. J. Roy. Meteor. Soc., 147, 22812292, https://doi.org/10.1002/qj.4022.

    • Search Google Scholar
    • Export Citation
  • Shupe, M. D., and J. M. Intrieri, 2004: Cloud radiative forcing of the Arctic surface: The influence of cloud properties, surface albedo, and solar zenith angle. J. Climate, 17, 616628, https://doi.org/10.1175/1520-0442(2004)017<0616:CRFOTA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Sorteberg, A., and J. E. Walsh, 2008: Seasonal cyclone variability at 70°N and its impact on moisture transport into the Arctic. Tellus, 60A, 570586, https://doi.org/10.1111/j.1600-0870.2007.00314.x.

    • Search Google Scholar
    • Export Citation
  • Sprenger, M., and H. Wernli, 2015: The LAGRANTO Lagrangian analysis tool – version 2.0. Geosci. Model Dev., 8, 25692586, https://doi.org/10.5194/gmd-8-2569-2015.

    • Search Google Scholar
    • Export Citation
  • Vargas Zeppetello, L. R., A. Donohoe, and D. S. Battisti, 2019: Does surface temperature respond to or determine downwelling longwave radiation? Geophys. Res. Lett., 46, 27812789, https://doi.org/10.1029/2019GL082220.

    • Search Google Scholar
    • Export Citation
  • Vihma, T., and Coauthors, 2016: The atmospheric role in the Arctic water cycle: A review on processes, past and future changes, and their impacts. J. Geophys. Res. Biogeosci., 121, 586620, https://doi.org/10.1002/2015JG003132.

    • Search Google Scholar
    • Export Citation
  • Wernli, H., and H. C. Davies, 1997: A Lagrangian-based analysis of extratropical cyclones. I: The method and some applications. Quart. J. Roy. Meteor. Soc., 123, 467489, https://doi.org/10.1002/qj.49712353811.

    • Search Google Scholar
    • Export Citation
  • Wernli, H., and L. Papritz, 2018: Role of polar anticyclones and mid-latitude cyclones for Arctic summertime sea-ice melting. Nat. Geosci., 11, 108113, https://doi.org/10.1038/s41561-017-0041-0.

    • Search Google Scholar
    • Export Citation
  • Woods, C., and R. Caballero, 2016: The role of moist intrusions in winter Arctic warming and sea ice decline. J. Climate, 29, 44734485, https://doi.org/10.1175/JCLI-D-15-0773.1.

    • Search Google Scholar
    • Export Citation
  • Woods, C., R. Caballero, and G. Svensson, 2013: Large-scale circulation associated with moisture intrusions into the Arctic during winter. Geophys. Res. Lett., 40, 47174721, https://doi.org/10.1002/grl.50912.

    • Search Google Scholar
    • Export Citation

Supplementary Materials

Save
  • Baggett, C., S. Lee, and S. Feldstein, 2016: An investigation of the presence of atmospheric rivers over the North Pacific during planetary-scale wave life cycles and their role in Arctic warming. J. Atmos. Sci., 73, 43294347, https://doi.org/10.1175/JAS-D-16-0033.1.

    • Search Google Scholar
    • Export Citation
  • Binder, H., M. Boettcher, C. M. Grams, H. Joos, S. Pfahl, and H. Wernli, 2017: Exceptional air mass transport and dynamical drivers of an extreme wintertime Arctic warm event. Geophys. Res. Lett., 44, 12 02812 036, https://doi.org/10.1002/2017GL075841.

    • Search Google Scholar
    • Export Citation
  • Boisvert, L. N., A. A. Petty, and J. C. Stroeve, 2016: The impact of the extreme winter 2015/16 Arctic cyclone on the Barents–Kara Seas. Mon. Wea. Rev., 144, 42794287, https://doi.org/10.1175/MWR-D-16-0234.1.

    • Search Google Scholar
    • Export Citation
  • Bozem, H., and Coauthors, 2019: Characterization of transport regimes and the polar dome during Arctic spring and summer using in-situ aircraft measurements. Atmos. Chem. Phys., 19, 15 04915 071, https://doi.org/10.5194/acp-19-15049-2019.

    • Search Google Scholar
    • Export Citation
  • Clark, J. P., E. E. Clothiaux, S. B. Feldstein, and S. Lee, 2021: Drivers of global clear sky surface downwelling longwave irradiance trends from 1984 to 2017. Geophys. Res. Lett., 48, e2021GL093961, https://doi.org/10.1029/2021GL093961.

    • Search Google Scholar
    • Export Citation
  • Curry, J. A., J. L. Schramm, M. C. Serreze, and E. E. Ebert, 1995: Water vapor feedback over the Arctic Ocean. J. Geophys. Res., 100, 14 22314 229, https://doi.org/10.1029/95JD00824.

    • Search Google Scholar
    • Export Citation
  • Dada, L., and Coauthors, 2022: A central Arctic extreme aerosol event triggered by a warm air-mass intrusion. Nat. Commun., 13, 5290, https://doi.org/10.1038/s41467-022-32872-2.

    • Search Google Scholar
    • Export Citation
  • Doyle, J. G., G. Lesins, C. P. Thackray, C. Perro, G. J. Nott, T. J. Duck, R. Damoah, and J. R. Drummond, 2011: Water vapor intrusions into the high Arctic during winter. Geophys. Res. Lett., 38, L12806, https://doi.org/10.1029/2011GL047493.

    • Search Google Scholar
    • Export Citation
  • Dufour, A., O. Zolina, and S. K. Gulev, 2016: Atmospheric moisture transport to the Arctic: Assessment of reanalyses and analysis of transport components. J. Climate, 29, 50615081, https://doi.org/10.1175/JCLI-D-15-0559.1.

    • Search Google Scholar
    • Export Citation
  • Fearon, M. G., J. D. Doyle, D. R. Ryglicki, P. M. Finocchio, and M. Sprenger, 2021: The role of cyclones in moisture transport into the Arctic. Geophys. Res. Lett., 48, e2020GL090353, https://doi.org/10.1029/2020GL090353.

    • Search Google Scholar
    • Export Citation
  • Fletcher, J., S. Mason, and C. Jakob, 2016: The climatology, meteorology, and boundary layer structure of marine cold air outbreaks in both hemispheres. J. Climate, 29, 19992014, https://doi.org/10.1175/JCLI-D-15-0268.1.

    • Search Google Scholar
    • Export Citation
  • Francis, J. A., and E. Hunter, 2006: New insight into the disappearing Arctic sea ice. Eos, Trans. Amer. Geophys. Union, 87, 509511, https://doi.org/10.1029/2006EO460001.

    • Search Google Scholar
    • Export Citation
  • Graversen, R. G., and M. Burtu, 2016: Arctic amplification enhanced by latent energy transport of atmospheric planetary waves. Quart. J. Roy. Meteor. Soc., 142, 20462054, https://doi.org/10.1002/qj.2802.

    • Search Google Scholar
    • Export Citation
  • Henderson, G. R., B. S. Barrett, L. J. Wachowicz, K. S. Mattingly, J. R. Preece, and T. L. Mote, 2021: Local and remote atmospheric circulation drivers of Arctic change: A review. Front. Earth Sci., 9, 709896, https://doi.org/10.3389/feart.2021.709896.

    • Search Google Scholar
    • Export Citation
  • Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 19992049, https://doi.org/10.1002/qj.3803.

    • Search Google Scholar
    • Export Citation
  • Hofsteenge, M. G., R. G. Graversen, J. H. Rydsaa, and Z. Rey, 2022: The impact of atmospheric Rossby waves and cyclones on the Arctic sea ice variability. Climate Dyn., 59, 579594, https://doi.org/10.1007/s00382-022-06145-z.

    • Search Google Scholar
    • Export Citation
  • Kapsch, M.-L., R. G. Graversen, and M. Tjernström, 2013: Springtime atmospheric energy transport and the control of Arctic summer sea-ice extent. Nat. Climate Change, 3, 744748, https://doi.org/10.1038/nclimate1884.

    • Search Google Scholar
    • Export Citation
  • Lee, S., T. Gong, S. B. Feldstein, J. A. Screen, and I. Simmonds, 2017: Revisiting the cause of the 1989–2009 Arctic surface warming using the surface energy budget: Downward infrared radiation dominates the surface fluxes. Geophys. Res. Lett., 44, 10 65410 661, https://doi.org/10.1002/2017GL075375.

    • Search Google Scholar
    • Export Citation
  • Liniger, M. A., and H. C. Davies, 2003: Substructure of a map streamer. Quart. J. Roy. Meteor. Soc., 129, 633651, https://doi.org/10.1256/qj.02.28.

    • Search Google Scholar
    • Export Citation
  • Luo, B., D. Luo, L. Wu, L. Zhong, and I. Simmonds, 2017: Atmospheric circulation patterns which promote winter Arctic sea ice decline. Environ. Res. Lett., 12, 054017, https://doi.org/10.1088/1748-9326/aa69d0.

    • Search Google Scholar
    • Export Citation
  • Messori, G., C. Woods, and R. Caballero, 2018: On the drivers of wintertime temperature extremes in the high Arctic. J. Climate, 31, 15971618, https://doi.org/10.1175/JCLI-D-17-0386.1.

    • Search Google Scholar
    • Export Citation
  • Mortin, J., G. Svensson, R. G. Graversen, M.-L. Kapsch, J. C. Stroeve, and L. N. Boisvert, 2016: Melt onset over Arctic sea ice controlled by atmospheric moisture transport. Geophys. Res. Lett., 43, 66366642, https://doi.org/10.1002/2016GL069330.

    • Search Google Scholar
    • Export Citation
  • Murto, S., R. Caballero, G. Svensson, and L. Papritz, 2022: Interaction between Atlantic cyclones and Eurasian atmospheric blocking drives wintertime warm extremes in the high Arctic. Wea. Climate Dyn., 3, 2144, https://doi.org/10.5194/wcd-3-21-2022.

    • Search Google Scholar
    • Export Citation
  • Murto, S., L. Papritz, G. Messori, R. Caballero, G. Svensson, and H. Wernli, 2023: Extreme surface energy budget anomalies in the high Arctic in winter. J. Climate, 36, 35913609, https://doi.org/10.1175/JCLI-D-22-0209.1.

    • Search Google Scholar
    • Export Citation
  • Naakka, T., T. Nygård, T. Vihma, J. Sedlar, and R. Graversen, 2019: Atmospheric moisture transport between mid-latitudes and the Arctic: Regional, seasonal and vertical distributions. Int. J. Climatol., 39, 28622879, https://doi.org/10.1002/joc.5988.

    • Search Google Scholar
    • Export Citation
  • Nygård, T., R. G. Graversen, P. Uotila, T. Naakka, and T. Vihma, 2019: Strong dependence of wintertime Arctic moisture and cloud distributions on atmospheric large-scale circulation. J. Climate., 32, 87718790, https://doi.org/10.1175/JCLI-D-19-0242.1.

    • Search Google Scholar
    • Export Citation
  • Ogi, M., and J. M. Wallace, 2012: The role of summer surface wind anomalies in the summer Arctic sea ice extent in 2010 and 2011. Geophys. Res. Lett., 39, L09704, https://doi.org/10.1029/2012GL051330.

    • Search Google Scholar
    • Export Citation
  • Ohmura, A., 2001: Physical basis for the temperature-based melt-index method. J. Appl. Meteor. Climatol., 40, 753761, https://doi.org/10.1175/1520-0450(2001)040<0753:PBFTTB>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Orbe, C., P. A. Newman, D. W. Waugh, M. Holzer, L. D. Oman, F. Li, and L. M. Polvani, 2015: Airmass origin in the Arctic. Part I: Seasonality. J. Climate, 28, 49975014, https://doi.org/10.1175/JCLI-D-14-00720.1.

    • Search Google Scholar
    • Export Citation
  • Papritz, L., 2020: Arctic lower-tropospheric warm and cold extremes: Horizontal and vertical transport, diabatic processes, and linkage to synoptic circulation features. J. Climate, 33, 9931016, https://doi.org/10.1175/JCLI-D-19-0638.1.

    • Search Google Scholar
    • Export Citation
  • Papritz, L., and T. Spengler, 2017: A Lagrangian climatology of wintertime cold air outbreaks in the Irminger and Nordic seas and their role in shaping air-sea heat fluxes. J. Climate, 30, 27172737, https://doi.org/10.1175/JCLI-D-16-0605.1.

    • Search Google Scholar
    • Export Citation
  • Papritz, L., and E. Dunn-Sigouin, 2020: What configuration of the atmospheric circulation drives extreme net and total moisture transport into the Arctic. Geophys. Res. Lett., 47, e2020GL089769, https://doi.org/10.1029/2020GL089769.

    • Search Google Scholar
    • Export Citation
  • Persson, P. O. G., M. D. Shupe, D. Perovich, and A. Solomon, 2017: Linking atmospheric synoptic transport, cloud phase, surface energy fluxes, and sea-ice growth: Observations of midwinter SHEBA conditions. Climate Dyn., 49, 13411364, https://doi.org/10.1007/s00382-016-3383-1.

    • Search Google Scholar
    • Export Citation
  • Pithan, F., and Coauthors, 2018: Role of air-mass transformations in exchange between the Arctic and mid-latitudes. Nat. Geosci., 11, 805812, https://doi.org/10.1038/s41561-018-0234-1.

    • Search Google Scholar
    • Export Citation
  • Richter-Menge, J., and M. L. Druckenmiller, 2020: The Arctic [in “State of the Climate in 2019”]. Bull. Amer. Meteor. Soc., 101 (Suppl.), S239S285, https://doi.org/10.1175/2020BAMSStateoftheClimate.1.

    • Search Google Scholar
    • Export Citation
  • Röthlisberger, M., and L. Papritz, 2023a: A global quantification of the physical processes leading to near-surface cold extremes. Geophys. Res. Lett., 50, e2022GL101670, https://doi.org/10.1029/2022GL101670.

    • Search Google Scholar
    • Export Citation
  • Röthlisberger, M., and L. Papritz, 2023b: Quantifying the physical processes leading to atmospheric hot extremes at a global scale. Nat. Geosci., 16, 210216, https://doi.org/10.1038/s41561-023-01126-1.

    • Search Google Scholar
    • Export Citation
  • Rydsaa, J. H., R. G. Graversen, T. I. H. Heiskanen, and P. J. Stoll, 2021: Changes in atmospheric latent energy transport into the Arctic: Planetary versus synoptic scales. Quart. J. Roy. Meteor. Soc., 147, 22812292, https://doi.org/10.1002/qj.4022.

    • Search Google Scholar
    • Export Citation
  • Shupe, M. D., and J. M. Intrieri, 2004: Cloud radiative forcing of the Arctic surface: The influence of cloud properties, surface albedo, and solar zenith angle. J. Climate, 17, 616628, https://doi.org/10.1175/1520-0442(2004)017<0616:CRFOTA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Sorteberg, A., and J. E. Walsh, 2008: Seasonal cyclone variability at 70°N and its impact on moisture transport into the Arctic. Tellus, 60A, 570586, https://doi.org/10.1111/j.1600-0870.2007.00314.x.

    • Search Google Scholar
    • Export Citation
  • Sprenger, M., and H. Wernli, 2015: The LAGRANTO Lagrangian analysis tool – version 2.0. Geosci. Model Dev., 8, 25692586, https://doi.org/10.5194/gmd-8-2569-2015.

    • Search Google Scholar
    • Export Citation
  • Vargas Zeppetello, L. R., A. Donohoe, and D. S. Battisti, 2019: Does surface temperature respond to or determine downwelling longwave radiation? Geophys. Res. Lett., 46, 27812789, https://doi.org/10.1029/2019GL082220.

    • Search Google Scholar
    • Export Citation
  • Vihma, T., and Coauthors, 2016: The atmospheric role in the Arctic water cycle: A review on processes, past and future changes, and their impacts. J. Geophys. Res. Biogeosci., 121, 586620, https://doi.org/10.1002/2015JG003132.

    • Search Google Scholar
    • Export Citation
  • Wernli, H., and H. C. Davies, 1997: A Lagrangian-based analysis of extratropical cyclones. I: The method and some applications. Quart. J. Roy. Meteor. Soc., 123, 467489, https://doi.org/10.1002/qj.49712353811.

    • Search Google Scholar
    • Export Citation
  • Wernli, H., and L. Papritz, 2018: Role of polar anticyclones and mid-latitude cyclones for Arctic summertime sea-ice melting. Nat. Geosci., 11, 108113, https://doi.org/10.1038/s41561-017-0041-0.

    • Search Google Scholar
    • Export Citation
  • Woods, C., and R. Caballero, 2016: The role of moist intrusions in winter Arctic warming and sea ice decline. J. Climate, 29, 44734485, https://doi.org/10.1175/JCLI-D-15-0773.1.

    • Search Google Scholar
    • Export Citation
  • Woods, C., R. Caballero, and G. Svensson, 2013: Large-scale circulation associated with moisture intrusions into the Arctic during winter. Geophys. Res. Lett., 40, 47174721, https://doi.org/10.1002/grl.50912.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    Correlations between θ* at a given level with θ* at 940 hPa (red diamonds) and the SEB anomaly (bars) for all days during extended winter [November–March (NDJFM)] during 1979–2020 at 1200 UTC averaged over the polar cap poleward of 85°N. Data are from ERA5 (see section 2).

  • Fig. 2.

    Schematic illustrating processes along air parcel trajectories causing positive potential temperature anomalies (θ*) in the Arctic. The transparent gray surface indicates a climatological isentrope and arrows illustrate various processes generating positive θ*, i.e., horizontal and vertical transports across climatological isentropes (dark and light blue, respectively), as well as diabatic processes including latent heat release (red) and turbulent fluxes into the boundary layer (orange).

  • Fig. 3.

    Generation of θ* due to horizontal and vertical transport. (a) θ* acquired by an air parcel transported isobarically at 700 hPa from a given geographical location into the polar cap poleward of 80°N with the 6- and 18-K and the 12-K contours highlighted by thin and thick black lines, respectively. The figure is obtained by subtracting the polar cap (≥80°N) average of the potential temperature climatology θc from θc at a given geographical location. Note that variations of θc within the 80°N polar cap are small as compared to θ* acquired due to transport from lower latitudes into the polar cap. (b) θ* acquired by an air parcel subsiding from 700 to 800 hPa at a fixed geographical location. Both panels show averages over extended winters 1979–2020 and the polar cap 80°N is indicated by a black circle.

  • Fig. 4.

    Case study of LCE 44 showing the positions of trajectories initialized at 50 hPa AGL from all patches associated with the LCE. Maps are shown for lags of (a) −8, (b) −6, (c) −4, and (d) −2 days prior to 1200 UTC 12 Dec 1986, corresponding to the peak of the LCE. Trajectory positions are colored according to pressure. Additionally shown are sea level pressure (gray contours at intervals of 10 hPa; blue and gray shading for values below 990 hPa and above 1030 hPa, respectively) and the 5000 m (blue thin), 5300 m (blue thick), and 5600 m (blue thin) isohypses of 500 hPa geopotential height. The purple hatched area indicates the region affected by the LCE and the solid gray line shows the 60°N latitude circle. Note that only positions from every second trajectory are drawn for better visibility.

  • Fig. 5.

    θ* budgets for LCEs (left) 44 and (right) 137. Shown is the temporal evolution of θ* (black) and of accumulated contributions to θ* from horizontal transport (dark blue), vertical transport (light blue), latent heating (red), boundary layer heating (orange), residual diabatic heating (yellow), and diabatic cooling (purple) for trajectories initialized at (a),(b) 500 and (c),(d) 50 hPa AGL (axis on the right) from all SEB anomaly patches associated with the given LCE. For each trajectory, terms are set to zero for t = tgenesis and accumulated for the period 0 ≥ ttgenesis. The resulting time series are then averaged over all trajectories. Gray bars show the fraction of trajectories with ttgenesis (axis on the left). t = 0 h corresponds to the time of arrival in the LCE and t = tgenesis to the time when θ* was last ≤0 K.

  • Fig. 6.

    As in Fig. 4, but for LCE 137 with lags relative to the peak of that LCE at 1200 UTC 3 Jan 1994.

  • Fig. 7.

    Histograms of tgenesis for all Atlantic and Pacific LCEs (gray bars) and temporal evolution of mean (solid lines) and interquartile range (shading) of θ* for Atlantic (blue) and Pacific (red) LCEs for trajectories initialized at (a) 500 and (b) 50 hPa AGL. Dashed black lines show the 75th percentile of θ* associated with climatological trajectories (see section 2c). Bars are shown every 6 h and the leftmost bar includes trajectories with tgenesis ≤ −240 h.

  • Fig. 8.

    Spatial distribution of trajectories initialized at (a),(d) 50, (b),(e) 200, and (c),(f) 500 hPa AGL associated with (a)–(c) Atlantic and (d)–(f) Pacific LCEs. Trajectory positions are shown for t = −240 h [red hatched; contours at 1% and 2% (106 km2)−1], t = tgenesis (shading), and t = 0 h [black; contour at 1% (106 km2)−1]. Further shown in (b) and (e) are the 6-, 12-, and 18-K contours of the gain of θ* due to horizontal transport at 700 hPa from a given geographical location into the 80°N polar cap (thin and thick black lines; as in Fig. 3a).

  • Fig. 9.

    (a),(b) Accumulated contributions of horizontal transport (dark blue), vertical transport (light blue), latent heating (red), boundary layer heating (orange), residual diabatic heating (yellow), and diabatic cooling (purple) to the final θ* (diamonds) for all (a) Atlantic and (b) Pacific LCEs. For each trajectory, the terms are accumulated from t = tgenesis to t = 0 h. (c),(d) As in (a) and (b), but for the difference between trajectories with the 20% largest (upper quintile) and smallest (lower quintile) θ* at t = 0 h. Quintiles are defined by level.