1. Introduction
In this study, we present observations of ocean turbulence and mixing in the Beaufort Sea’s Amundsen Gulf in the western Arctic Ocean (Fig. 1) from a series of conductivity–temperature–depth (CTD) and microstructure measurements collected in summer 2015. The Amundsen Gulf is the site of the Cape Bathurst polynya, part of the circum-Arctic system of flaw polynyas that are important in the dynamics of the Arctic ice sheet, the formation of Arctic deep water, and, in the case of the Cape Bathurst polynya, as a habitat for some of the highest densities of birds and mammals found anywhere in the Arctic (Arrigo and van Dijken 2004; Harwood and Stirling 1992; Dickson and Gilchrist 2002). It is also strategically located along the Northwest Passage and is expected to become a major commercial shipping line as summer sea ice continues to decrease (Prowse et al. 2009; Khon et al. 2010). Since the late 1990s, there has been a dramatic reduction in the extent and age of multiyear sea ice in the region, as well as large interannual variations in summer ice concentration and the duration of summertime ice clearance (Niemi et al. 2012). Over the whole Arctic, changes in sea ice alongside other rapid and dramatic Arctic system changes in the past few decades have perturbed regional ecosystems at all tropic levels (e.g., Grebmeier et al. 2006; Wassmann 2011, 2015) and have the potential to affect ecosystem services related to natural resources, food production, climate regulation, and cultural integrity (Post et al. 2009). In light of these ongoing changes to the broader physical environment, it is important to continue developing a detailed understanding of the physical oceanography and, in particular, the mixing characteristics of the region to facilitate studies that will model the environmental and ecological responses to future regional climate change (e.g., Carmack and MacDonald 2002; Rainville et al. 2011; Carmack et al. 2015).

(a) Map of the southeastern Beaufort Sea, showing the location of Amundsen Gulf. The glider path is shown by the thin black line inside the black-outlined rectangle. (b) Enlarged view of the region enclosed by the black-outlined rectangle in (a), showing the path of the glider. The start and end locations of the track are shown by the white rectangles; four intermediate waypoints are also shown and are numbered consecutively. The color on the glider’s track-line is Conservative Temperature along the 1026.15 kg m−3 isopycnal, indicating the location and spatial scale of the warm-core eddy discussed in the text (section 5c). The white star is the location of ArcticNet mooring CA08. Bathymetry data are from IBCAO 3.0 (Jakobsson et al. 2012).
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

(a) Map of the southeastern Beaufort Sea, showing the location of Amundsen Gulf. The glider path is shown by the thin black line inside the black-outlined rectangle. (b) Enlarged view of the region enclosed by the black-outlined rectangle in (a), showing the path of the glider. The start and end locations of the track are shown by the white rectangles; four intermediate waypoints are also shown and are numbered consecutively. The color on the glider’s track-line is Conservative Temperature along the 1026.15 kg m−3 isopycnal, indicating the location and spatial scale of the warm-core eddy discussed in the text (section 5c). The white star is the location of ArcticNet mooring CA08. Bathymetry data are from IBCAO 3.0 (Jakobsson et al. 2012).
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
(a) Map of the southeastern Beaufort Sea, showing the location of Amundsen Gulf. The glider path is shown by the thin black line inside the black-outlined rectangle. (b) Enlarged view of the region enclosed by the black-outlined rectangle in (a), showing the path of the glider. The start and end locations of the track are shown by the white rectangles; four intermediate waypoints are also shown and are numbered consecutively. The color on the glider’s track-line is Conservative Temperature along the 1026.15 kg m−3 isopycnal, indicating the location and spatial scale of the warm-core eddy discussed in the text (section 5c). The white star is the location of ArcticNet mooring CA08. Bathymetry data are from IBCAO 3.0 (Jakobsson et al. 2012).
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
The Arctic Ocean is undersampled with respect to turbulence and mixing rates, and observations of ocean turbulence are notably scarce in the Beaufort Sea. A number of studies over the previous decade have deepened our understanding of mixing rates and mechanisms in the broader Canada Basin (e.g., Rainville and Winsor 2008; Timmermans et al. 2008a,b; Guthrie et al. 2013; Dosser et al. 2014; Shaw and Stanton 2014; Chanona et al. 2018), but, to our knowledge, only six previous studies (Padman and Dillon 1987; Bourgault et al. 2011; Shroyer 2012; Rippeth et al. 2015; Lincoln et al. 2016; Fine et al. 2018) have used microstructure measurements to characterize mixing rates in the Beaufort Sea directly. Many of these studies have focused on specific processes such as diffusive convection (Padman and Dillon 1987), mesoscale eddy heat loss (Fine et al. 2018), and canyon flows (Shroyer 2012). Outside of these, the studies collectively confirm that turbulence in the Beaufort Sea is generally very weak and mixing rates are generally very small away from the surface and significant topography. This is consistent with the indirect estimates of Guthrie et al. (2013) that report internal wave energies and consequent background mixing rates that are lower by about a factor of 5 in the Beaufort Sea than in the central and eastern Arctic Ocean. Lincoln et al. (2016) attribute the suppression of turbulent mixing at intermediate depths to the strong stratification that characterizes the central Canada Basin, which was observed to buffer intermediate depths from enhanced wind-driven internal wave energy. At the same time, a number of observations also suggest a key role of topography, and in particular the steep topography of the basin margins and of canyons, in driving a localized enhancement of turbulence and mixing (Shroyer 2012; Rippeth et al. 2015; Lincoln et al. 2016). In addition, an important role of mesoscale eddy features in modulating the space and time variability in mixing rates has been documented (Fine et al. 2018). There remains a pressing need to continue building a broad record of mixing estimates in the region to better understand the space and time geography of mixing rates and mechanisms, and their susceptibility to change.
With this study, we contribute to a more comprehensive understanding of the mixing environment within Amundsen Gulf of the southeastern Beaufort Sea by providing a novel description of the region’s turbulence and mixing characteristics. By using an autonomous ocean glider equipped with a CTD and turbulence sensors, we collected a large number of tightly resolved measurements of turbulent microstructure shear and temperature gradients, concurrent with observations of the hydrography, made possible by the long-duration, continuous, high-frequency sampling capabilities of the glider platform. This novel mode of sampling is significant because a large number of tightly resolved measurements are needed to accurately characterize turbulence, which tends to be described with intermittent, lognormally distributed variables that are easily undersampled (Baker and Gibson 1987; Gregg 1987). Further, it permits an examination of high-frequency time and space patterns that are difficult to resolve with traditional turbulence sampling techniques. We use these measurements to characterize the statistical distributions of estimates of the turbulent dissipation rate of kinetic energy, diapycnal mixing rate, and vertical heat flux during the period of observation. In addition, we use these observations to gain insight into the relative importance of different mechanisms, namely, tidal mixing, double diffusion, and near-surface mesoscale and smaller processes, that underpin and/or modulate the observed turbulence environment. To our knowledge, this is the first time such a highly resolved characterization of mixing from direct turbulence measurements has been presented for this region, and the first time an autonomous instrument has been used to characterize the statistics of turbulence and mixing in the western Arctic Ocean.
The remainder of the document is structured as follows. In section 2, we describe the CTD and microstructure measurements from the glider, and briefly outline the data processing methods. Section 3 uses the CTD measurements to describe the relevant hydrographic context. In section 4, we present the primary results of this study, the turbulence observations and the mixing rate and heat flux estimates derived from them. Section 5 presents a discussion of mixing mechanisms. We synthesize our results in section 6.
2. Measurements and data processing
a. Sampling strategy
We collected CTD and turbulence measurements in Amundsen Gulf using an autonomous 1000-m-rated Teledyne Webb Slocum G2 ocean glider, fitted with 1) a pumped Sea-Bird CTD measuring conductivity, temperature, and pressure and 2) an externally mounted turbulence-sensing package measuring shear and temperature microstructure (section 2b). The measurements used in this study are those first described by Scheifele et al. (2018), collected continuously over 11 days during the period from 25 August to 5 September 2015.
The 186-km horizontal path of the glider, immediately northwest of the gulf’s sill, is shown in Fig. 1. The glider spent the first 5 days in the central gulf, where the water depth exceeds 400 m, and the remaining time on three traverses of the continental shelf near Banks Island. Along this path, the glider collected 348 discrete quasi-vertical measurement profiles, at a nominal glide angle of 26° from the horizontal. The first 112 profiles, in water ~410 m deep, extend from the near surface to a fixed depth of 300 m; later profiles typically extend to within 15 m of the local bottom, which ranged between 205 and 430 m in depth. The location of each profile is approximated with its mean coordinates, neglecting horizontal translation that occurs over the course of one profile. The mean and standard deviation distances between consecutive profiles are 536 and 357 m, respectively.
b. Turbulence measurements and data processing
The glider carried an externally mounted, self-contained microstructure sensing package known as a Microrider, also used in recent studies by Fer et al. (2014), Peterson and Fer (2014), Palmer et al. (2015), and Schultze et al. (2017). The Microrider is manufactured by Rockland Scientific and is factory-installed on the glider. Our configuration of the Microrider had two velocity shear probes and two fast-response thermistors, each sampling at 512 Hz, measuring microstructure velocity and temperature gradients, respectively. All sensors sampled continuously during the deployment, but one of the two shear probes failed after the first three days of measurement.
We derive independent estimates of the turbulent kinetic energy (TKE) dissipation rate ε from each of the four microstructure channels. This rate is a measure of the intensity of turbulence in the flow and is proportional to the rate of diapycnal mixing in the Osborn (1980) model (see section 4c). We briefly outline our methods to derive ε from the shear and temperature microstructure measurements below; a more detailed description of the methods and their limitations is given in Scheifele et al. (2018).
We calculate the TKE dissipation rate from the temperature microstructure measurements using power spectra of temperature gradients, calculated over the same 40-s segments and 4-s subsegments that we used to calculate the shear spectra. We fit a theoretical form for the temperature gradient spectrum (the Batchelor spectrum) to the observed temperature gradient power spectrum using the maximum likelihood estimator method outlined by Ruddick et al. (2000). In this procedure, ε is a variable fitting parameter that is optimized by minimizing the difference between the observed and theoretical spectra. We refer to this optimized value as εT, with the subscript T indicating a temperature gradient-derived dissipation rate estimate.
Both εU and εT estimates are then subjected to a series of quality control criteria that remove suspect estimates. These routines are designed to flag and remove values where, for example, the glider’s flight was not steady, shear probes contacted small marine organisms or debris, Taylor’s frozen turbulence hypothesis is violated in the calculation of power spectra, etc. They are detailed in Scheifele et al. (2018). Quality control removes 22% of εU estimates and 34% of εT estimates.
Dual estimates of ε from each set of probes are arithmetically averaged to obtain single εU and εT values for each 40-s segment. As described in detail in Scheifele et al. (2018), we find agreement within a factor of 2 from the two types of estimates when ε exceeds 3 × 10−11 W kg−1, a threshold we identify as the noise floor of the shear-derived estimates. However, our analysis indicates that the noise floor of the shear measurements biases the statistical distribution of εU at values of ε as large as 1 × 10−10 W kg−1, while temperature-derived estimates of ε are reliable to values as small as 2 × 10−12 W kg−1. The averaged εU and εT estimates are combined into a single best ε estimate using the following method. When εU ≥ 1 × 10−10 W kg−1, we keep only εU because the shear-derived estimate relies more directly on the definition of the dissipation rate and is more reliable in energetic conditions (Gregg 1999). However, if the εU estimate is unavailable because it failed quality control, we keep the coincident εT estimate instead, if this is available. If both are available, but they differ by more than a factor of 10, both are discarded. Below 1 × 10−10 W kg−1, we keep only εT. However, when εT < 2 × 10−12 W kg−1, we set ε to zero, following the approach used by Gregg et al. (2012). We are left with 22 153 unique ε estimates for the remaining analysis; of these, 4699 (21%) are set to zero.
c. Arithmetic versus geometric averaging
Turbulence in the ocean is patchy in space and intermittent in time, and the distributions of dissipation rates and mixing coefficients are typically lognormal-like, spanning many orders of magnitude (Baker and Gibson 1987). Arithmetic mean values may differ from median and geometric mean values by orders of magnitude in such data, so it is important to distinguish between these various metrics and recognize their distinct physical interpretations (Kirkwood 1979). The geometric mean (GM) is defined as
3. Hydrography
In Fig. 2, we present the fields of Conservative Temperature Θ and squared buoyancy frequency N2 = −(g/ρθ)∂ρθ/∂z. Here g is the gravitational acceleration and ρθ is the potential density referenced to the sea surface, derived from the CTD measurements and smoothed by a 15-s running mean filter when used in the calculation of N2. For each field, a mean vertical profile and a spatial cross section are shown. The horizontal coordinate in the cross sections is the glider’s geographic along-track distance coordinate s measured along the two-dimensional track shown in Fig. 1b. To guide the eye, each cross section is broken into multiple panels at waypoints where the glider changed its direction of travel.

(a) Arithmetic mean profile and spatial cross section of Conservative Temperature Θ. (b) Geometric mean profile and spatial cross section of stratification N2. For the mean profiles, alternating colored background shading indicates the approximate depth ranges of the hydrographic layers defined in the text (PW, WH, and AW are labeled). For the spatial sections, the horizontal axis is broken and consecutively labeled 1–4 at the waypoints marked in Fig. 1, indicating where the glider changed direction. The white-outlined rectangle in (a) indicates the mesoscale eddy discussed in the text.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

(a) Arithmetic mean profile and spatial cross section of Conservative Temperature Θ. (b) Geometric mean profile and spatial cross section of stratification N2. For the mean profiles, alternating colored background shading indicates the approximate depth ranges of the hydrographic layers defined in the text (PW, WH, and AW are labeled). For the spatial sections, the horizontal axis is broken and consecutively labeled 1–4 at the waypoints marked in Fig. 1, indicating where the glider changed direction. The white-outlined rectangle in (a) indicates the mesoscale eddy discussed in the text.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
(a) Arithmetic mean profile and spatial cross section of Conservative Temperature Θ. (b) Geometric mean profile and spatial cross section of stratification N2. For the mean profiles, alternating colored background shading indicates the approximate depth ranges of the hydrographic layers defined in the text (PW, WH, and AW are labeled). For the spatial sections, the horizontal axis is broken and consecutively labeled 1–4 at the waypoints marked in Fig. 1, indicating where the glider changed direction. The white-outlined rectangle in (a) indicates the mesoscale eddy discussed in the text.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
We observe five distinct hydrographic layers, similar to those used to describe layering in the Canada Basin (e.g., Carmack et al. 1989). From shallowest to deepest, these are a warm surface mixed layer (SML), a strongly stratified near-surface cold halocline (CH), a cold Pacific Water (PW) layer, an intermediate warm halocline (WH) where temperature increases with depth, and a warm Atlantic Water (AW) layer. We define the boundaries of the layers using their local Conservative Temperature and Absolute Salinity SA structure using criteria similar to those described in Jones (2001), Jackson et al. (2010), and Timmermans et al. (2014); the boundaries and hydrographic characteristics of the layers we observe are summarized in Table 1. The layering can be seen most easily in the Conservative Temperature profile (Fig. 2a).
Properties of the hydrographic layers. Layers are defined by their Conservative Temperature Θ and Absolute Salinity SA. The ranges given for depth, Conservative Temperature Θ, potential density anomaly σ0, and stratification N2 are for the central 90% of the data. The layer labels are SML: surface mixed layer; CH: cold halocline; PW: Pacific Water layer; WH: warm halocline; AW: Atlantic Water layer.


Second, the stratification is strong everywhere in the subsurface relative to that in lower-latitude oceans. Typical values for N2 in the North Atlantic and North Pacific pycnoclines are O(10−6) s−2 (Emery et al. 1984). This benchmark is comparable to the smallest N2 values we observe in the AW but is nearly two orders of magnitude smaller than N2 in the PW layer and is three orders of magnitude smaller than N2 in the CH. Numerous previous studies in the Beaufort Sea (e.g., Guthrie et al. 2013; Lincoln et al. 2016; Chanona et al. 2018) have noted that stratification is a key controlling feature of the mixing characteristics in this region. We will build on these results in section 4 by combining our measurements of the turbulence and density fields to demonstrate that density stratification frequently inhibits turbulent mixing in our dataset.
While most of the subsurface appears to be generally uniform in the horizontal plane, we observe substantial horizontal mesoscale and smaller [O(1) km] Conservative Temperature variability in the PW layer (Fig. 2a). Most distinctive is the presence of a mesoscale eddy between waypoints 2 and 3. This variability and its potential implications for the layer’s heat budget are discussed further in section 5c.
4. Turbulence and mixing
a. Turbulent dissipation rates
As is typical for ocean turbulence observations (e.g., Gregg 1987; Lueck et al. 2002), we observe an ε distribution that spans many orders of magnitude. Further, the positive skew in the distribution (Fig. 3a) suggests a relatively small number of strongly turbulent events occurring in a less turbulent background flow field. We note that the tail on the left side of the distribution is artificially cut off by the estimated noise floor of our microstructure temperature measurements: values of ε below the detection limit (21% of the data; section 2b), representing turbulence too weak for us to observe, are set to zero for the calculation of arithmetic mean and median values, and are not depicted in Fig. 3a. Resolvable ε realizations vary over five orders of magnitude, from O(10−12) to O(10−8) W kg−1. The interquartile range (IQR) of these estimates is (3.0–160) × 10−12 W kg−1.

Histograms of (a) the turbulent dissipation rate ε and (b) the buoyancy Reynolds number ReB. For each, the number in the top right indicates the percentage of data that fall within the axis limits; the remaining data are below the detection limit and are not displayed. The interquartile range for each set, including below-detection-limit data, is the span between the two dash–dotted vertical lines. For ε, the geometric and arithmetic mean values are also indicated (GM and AM, respectively). For ReB, the approximate critical value
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

Histograms of (a) the turbulent dissipation rate ε and (b) the buoyancy Reynolds number ReB. For each, the number in the top right indicates the percentage of data that fall within the axis limits; the remaining data are below the detection limit and are not displayed. The interquartile range for each set, including below-detection-limit data, is the span between the two dash–dotted vertical lines. For ε, the geometric and arithmetic mean values are also indicated (GM and AM, respectively). For ReB, the approximate critical value
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Histograms of (a) the turbulent dissipation rate ε and (b) the buoyancy Reynolds number ReB. For each, the number in the top right indicates the percentage of data that fall within the axis limits; the remaining data are below the detection limit and are not displayed. The interquartile range for each set, including below-detection-limit data, is the span between the two dash–dotted vertical lines. For ε, the geometric and arithmetic mean values are also indicated (GM and AM, respectively). For ReB, the approximate critical value
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Consistent with previous studies in the region (e.g., Bourgault et al. 2011; Guthrie et al. 2013; Rippeth et al. 2015; Lincoln et al. 2016), we find that turbulence is typically very weak. The geometric mean dissipation rate, representing the central tendency of a lognormally distributed variable, is 2.8 [2.7, 2.8] × 10−11 W kg−1.1 The median value is 2.3 [2.2, 2.4] × 10−11 W kg−1. In 68% of these observations, ε is smaller than 1.0 × 10−10 W kg−1, a common benchmark for “low turbulence” open ocean dissipation rates (Gregg 1999; Lueck et al. 2002). Only 0.4% of the distribution lies above 1 × 10−8 W kg−1. However, the small number of large ε estimates do play an important role in setting the integrated dissipation rate over the period of observation: the arithmetic mean value of ε is 4.9 [4.0, 6.8] × 10−10 W kg−1, more than an order of magnitude larger than the geometric mean and median values.
The variability of the ε field has a notable spatial structure that can be identified in the mean vertical profiles and horizontal cross section of the field (Fig. 4a). In the vertical, there is an ε minimum in the core of the cold PW layer at ~100-m depth: here ε is typically O(10−11) W kg−1. The geometric average dissipation rate near the sea surface and the seafloor indicates that ε is typically an order of magnitude larger in proximity to these boundaries. Consistent with the dataset as a whole, the arithmetic mean in each of these depth bins is an order of magnitude larger than the geometric mean (the arithmetic mean of ε is 4.4 [1.0, 20] × 10−9 W kg−1 at ~20-m depth, 1.1 [0.98, 1.2] × 10−10 W kg−1 at ~110-m depth, and 1.1 [0.79 1.4] × 10−9 W kg−1 at ~350-m depth). Larger uncertainty bounds in the mean of the near-surface bin reflects a stronger influence of high-end outliers of the distribution there. Laterally, the most obvious source of variability is a prominent near-bottom patch of elevated dissipation at the base of the continental slope, where ε can be as high as O(10−8) W kg−1. This turbulent patch is found between s = 52 and 81 km and is identified in Fig. 4a by a white-outlined rectangle.

Mean vertical profiles and horizontal cross sections of (a) ε and (b) ReB. Waypoints are indicated as in Fig. 2. For each, the geometric mean profile is given in 25-m bins (blue), with error bars indicating the standard error in the mean using the geometric standard deviation; for ε, the arithmetic mean profile is also given (black), with error bars indicating the 95% confidence interval based on bootstrap resampling. As for all geometric mean values presented, below-detection limit values set to zero for all other calculations are set to the detection limit (2.0 × 10−12 W kg−1) for the calculation of the geometric mean. In both cross sections, the white-outlined rectangle between waypoints 2 and 3 identifies the patch of enhanced turbulence discussed in the text. In the ReB cross section, red pixels indicate where a turbulent diapycnal flux is expected and gray pixels indicate an expected absence of turbulent diapycnal mixing. The approximate critical value
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

Mean vertical profiles and horizontal cross sections of (a) ε and (b) ReB. Waypoints are indicated as in Fig. 2. For each, the geometric mean profile is given in 25-m bins (blue), with error bars indicating the standard error in the mean using the geometric standard deviation; for ε, the arithmetic mean profile is also given (black), with error bars indicating the 95% confidence interval based on bootstrap resampling. As for all geometric mean values presented, below-detection limit values set to zero for all other calculations are set to the detection limit (2.0 × 10−12 W kg−1) for the calculation of the geometric mean. In both cross sections, the white-outlined rectangle between waypoints 2 and 3 identifies the patch of enhanced turbulence discussed in the text. In the ReB cross section, red pixels indicate where a turbulent diapycnal flux is expected and gray pixels indicate an expected absence of turbulent diapycnal mixing. The approximate critical value
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Mean vertical profiles and horizontal cross sections of (a) ε and (b) ReB. Waypoints are indicated as in Fig. 2. For each, the geometric mean profile is given in 25-m bins (blue), with error bars indicating the standard error in the mean using the geometric standard deviation; for ε, the arithmetic mean profile is also given (black), with error bars indicating the 95% confidence interval based on bootstrap resampling. As for all geometric mean values presented, below-detection limit values set to zero for all other calculations are set to the detection limit (2.0 × 10−12 W kg−1) for the calculation of the geometric mean. In both cross sections, the white-outlined rectangle between waypoints 2 and 3 identifies the patch of enhanced turbulence discussed in the text. In the ReB cross section, red pixels indicate where a turbulent diapycnal flux is expected and gray pixels indicate an expected absence of turbulent diapycnal mixing. The approximate critical value
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Dissipation rates in the turbulent patch are anomalously high relative to the rest of the field but modify the statistics of the full dataset only marginally (Table 2). For example, the arithmetic mean of ε excluding estimates from within the patch is 4.4 [3.4, 6.3] × 10−10 W kg−1, only 10% smaller than the estimate from the dataset as a whole. However, the arithmetic mean of data only from within the patch is 11 [9.7, 13] × 10−10 W kg−1, an increase by a factor of 2.2 over the mean calculated from the full set of data. For a similar comparison of the geometric mean value, median value, and IQR see Table 2.
Select statistics of ε and |Kρ| observed in all of the data, all data except that within the turbulent patch, and data only from within the turbulent patch. The turbulent patch is defined as the region inside the white-outlined rectangle in Figs. 4 and 7, between s = 52 and 81 km on the horizontal axis. Bracketed numbers indicate lower and upper bounds based on the 95% confidence interval from bootstrap resampling in the case of arithmetic mean and median values and the standard error in the mean from the geometric standard deviation in the case of geometric mean values.


Further information about the variability in the ε field is available from the glider’s three repeat transects over the continental shelf slope. A comparison of the depth-averaged dissipation rate estimates along the three transects is shown in Fig. 5, for each of which ε is plotted as a function of distance from the glider’s easternmost waypoint, geometrically averaged in 2.5-km bins. This bin-averaged dissipation rate remained of the same order of magnitude over the 7 days needed to complete the transects (note that the ε axis in Fig. 5 is linear, not logarithmic) and varied between (1–5) × 10−11 W kg−1. From the first and last transects, it appears that ε is systematically larger in the central gulf than it is on the shelf slope, but the second transect does not exhibit this pattern; nonetheless, when all transects are averaged together (not shown), the pattern of enhanced ε in the central gulf relative to on the shelf slope remains. Note that the turbulent patch discussed previously and identified in Fig. 4 by the white-outlined rectangle is situated to the immediate left of the leftmost axis limit in Fig. 5.

The three repeat ε transects (left vertical axis) over the continental shelf slope. The horizontal axis is the distance from waypoint 3 shown in Fig. 1b. Thick lines are 2.5-km, geometric mean bin-averages of cast-averaged ε; colored markers in the background are the individual geometric mean cast averages. The bathymetry is shown with gray shading in the background (right vertical axis) for reference.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

The three repeat ε transects (left vertical axis) over the continental shelf slope. The horizontal axis is the distance from waypoint 3 shown in Fig. 1b. Thick lines are 2.5-km, geometric mean bin-averages of cast-averaged ε; colored markers in the background are the individual geometric mean cast averages. The bathymetry is shown with gray shading in the background (right vertical axis) for reference.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
The three repeat ε transects (left vertical axis) over the continental shelf slope. The horizontal axis is the distance from waypoint 3 shown in Fig. 1b. Thick lines are 2.5-km, geometric mean bin-averages of cast-averaged ε; colored markers in the background are the individual geometric mean cast averages. The bathymetry is shown with gray shading in the background (right vertical axis) for reference.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
A notable attribute of the ε transects in Fig. 5 is a number of localized peaks in the depth-averaged ε value that appear in all three occupations of the transect at approximately the same horizontal position. Examples are the local maxima at ~5–7-, ~18–19-, and ~23-km distance. Note that it is not possible to decouple time and space variability in measurements taken from a glider transect; here, we have treated the ε observations as a spatial series in order to highlight what appear to be geographic features, but we will discuss temporal variability and its implications in section 5a.
b. The influence of stratification
Combining observations of ε with those of N2 and the kinematic viscosity ν, we construct estimates of the buoyancy Reynolds number, ReB = ε/(νN2), which quantifies the energetic capacity of the flow to develop vertical overturns that lead to diapycnal mixing (Figs. 3b and 4b). Note that in our computation of ReB we bin N2 similar to ε, described in section 2b, after applying a 15-s running mean smoothing. The ReB parameter is a measure of the relative separation between the Ozmidov and Kolmogorov scales, that is, the length scale over which turbulence first becomes confined by stratification and the length scale over which turbulence becomes dissipated by viscosity, respectively. On length scales between these scales (the inertial subrange) we expect turbulent kinetic energy transfer to follow a scale-free cascade and be responsible for overturning density structure. A small ReB value thus indicates that viscosity is acting to directly dissipate the turbulent motions responsible for overturning, whereas a large ReB value indicates that the large overturning scales are undamped by viscous dissipation. Evidence from laboratory, numerical and field studies suggests that the ReB parameter has a critical value near
We impose an
Further, it is apparent from the mean vertical profile and spatial cross section of the ReB field (Fig. 4b) that where turbulent mixing is expected to occur is not homogeneously distributed in space. Rather, most of the turbulent mixing expected in our dataset occurs within 100 m of the seafloor in the isolated patch of enhanced ε that we observed at the edge of the shelf slope region (i.e., inside the white-outlined rectangles in Fig. 4). Only here is ReB commonly of O(10) or larger, with individual values occasionally reaching as large as O(103). The white-outlined rectangle representing the region of enhanced ε in Fig. 4a encloses only 8% of ε estimates, but it encloses 41% of the occurrences where ReB ≥ 10 and 64% of those where ReB ≥ 100. Inside the rectangle, 37% of the ε estimates indicate that ReB ≥ 10; in contrast, only 5% of the estimates indicate that ReB ≥ 10 outside this region.
c. Density diffusivity estimates
A histogram of the density diffusivity estimates, separated into upgradient and downgradient subsets, is given in Fig. 6a. The discontinuity between 8 × 10−8 and 3 × 10−6 m2 s−1 reflects the distinction between our estimates of Kρ that assume turbulent mixing versus molecular diffusion, with all estimates to the right of the discontinuity computed using Eq. (3) and all estimates to the left of the discontinuity computed using Eq. (5). The discontinuity is an artifact of the Osborn model’s inability to describe the transition between turbulent and nonturbulent density fluxes. Although alternative models that describe a smooth transition between the turbulent and molecular regimes have been proposed (e.g., Bouffard and Boegman 2013), the treatment of these transitional regime estimates do not have a significant impact on the characterization of the integrated diffusivity: arithmetic Kρ averages are largely unaffected by variability (or inaccuracies) in the smaller orders of magnitude Kρ estimates. To illustrate, we compare the arithmetic mean of all Kρ estimates using the Osborn model [Eq. (3)] indiscriminately versus imposing the

Histogram of (a) the diapycnal mixing coefficient for density Kρ for all nonzero measurements and (b) the vertical heat flux FH for turbulent regime (ReB ≥ 10) estimates only. Positive Kρ indicates downgradient density diffusion; negative Kρ indicates upgradient density diffusion. The arithmetic mean of all Kρ estimates is indicated. For FH, the dash–dotted lines indicate the 5th and 95th percentiles.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

Histogram of (a) the diapycnal mixing coefficient for density Kρ for all nonzero measurements and (b) the vertical heat flux FH for turbulent regime (ReB ≥ 10) estimates only. Positive Kρ indicates downgradient density diffusion; negative Kρ indicates upgradient density diffusion. The arithmetic mean of all Kρ estimates is indicated. For FH, the dash–dotted lines indicate the 5th and 95th percentiles.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Histogram of (a) the diapycnal mixing coefficient for density Kρ for all nonzero measurements and (b) the vertical heat flux FH for turbulent regime (ReB ≥ 10) estimates only. Positive Kρ indicates downgradient density diffusion; negative Kρ indicates upgradient density diffusion. The arithmetic mean of all Kρ estimates is indicated. For FH, the dash–dotted lines indicate the 5th and 95th percentiles.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
The arithmetic mean of all Kρ estimates is 4.5 [2.3, 15] × 10−6 m2 s−1, three orders of magnitude larger than the arithmetic mean of the molecular regime estimates (−3.8 [−3.9, −3.7] × 10−9 m2 s−1). This implies that the upper 7% of Kρ estimates are responsible for drawing up the average mixing rate by a factor of approximately 2000 from that set by 93% of the observations in the molecular regime, a result that highlights the importance of the relatively small number of turbulent mixing events in setting the mean mixing rate. The arithmetic mean represents the 94th percentile of data: another indication of the disproportionate importance of the turbulent fluxes in setting the bulk transformation of buoyancy.
Like the dissipation rate, Kρ shows systematic vertical structure (Fig. 7a; left panel): it is typically smallest at the depths of the cold PW layer around 100-m depth; it is typically one to three orders of magnitude larger near the surface and near the bottom. In these spatial views (Fig. 7a), Kρ is computed locally as either a turbulent or molecular regime contribution as appropriate given the local ReB value; the mean vertical profile averages the molecular and turbulent contributions accounting for their relative occurrence as a function of depth bin. The disproportionate contribution of a few, large turbulent mixing rate estimates is again apparent when comparing the geometric mean vertical profile of Kρ to its arithmetic mean counterpart: the arithmetic mean profile is everywhere one to two orders of magnitude larger. In the upper 200 m of the water column, only 3% of the observations indicate a turbulent density flux. However, the arithmetic mean value of Kρ is typically O(10−7) m2 s−1, two orders of magnitude above the geometric mean value. Below 200-m depth, the arithmetic average of Kρ increases steadily and reaches a maximum of 3.3 × 10−5 m2 s−1 between 335- and 360-m depth. This elevated mean Kρ signal reflects the influence of the near-bottom turbulent patch at the edge of the shelf slope that is seen in the ε and ReB sections (Fig. 4): this turbulent patch is the dominant feature in the Kρ cross-sectional variability (Fig. 7a, right panel). Inside this patch, where 37% of the estimates indicate a turbulent mixing regime (section 4b), the arithmetic mean Kρ value is 4.6 [1.8, 18] × 10−5 m2 s−1; in comparison, outside the patch, where only 5% of mixing estimates are characterized as turbulent, the arithmetic mean value of Kρ is 1.0 [0.86, 1.4] × 10−6 m2 s−1. Further metrics comparing Kρ inside and outside the patch are presented in Table 2.

Mean vertical profiles, in 25-m bins, and horizontal cross sections of (a) Kρ and (b) FH. For Kρ, both geometric mean (blue) and arithmetic mean (black) profiles are shown; error bars indicate the standard error in the mean using the geometric standard deviation in the former and the 95% confidence interval based on bootstrap resampling in the latter. For FH, the arithmetic mean profile is shown, with error bars indicating the 95% confidence interval based on bootstrap resampling. For the cross sections, the horizontal axis, waypoint markers, and white-outlined rectangle identifying the turbulent patch are as in Fig. 4.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

Mean vertical profiles, in 25-m bins, and horizontal cross sections of (a) Kρ and (b) FH. For Kρ, both geometric mean (blue) and arithmetic mean (black) profiles are shown; error bars indicate the standard error in the mean using the geometric standard deviation in the former and the 95% confidence interval based on bootstrap resampling in the latter. For FH, the arithmetic mean profile is shown, with error bars indicating the 95% confidence interval based on bootstrap resampling. For the cross sections, the horizontal axis, waypoint markers, and white-outlined rectangle identifying the turbulent patch are as in Fig. 4.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Mean vertical profiles, in 25-m bins, and horizontal cross sections of (a) Kρ and (b) FH. For Kρ, both geometric mean (blue) and arithmetic mean (black) profiles are shown; error bars indicate the standard error in the mean using the geometric standard deviation in the former and the 95% confidence interval based on bootstrap resampling in the latter. For FH, the arithmetic mean profile is shown, with error bars indicating the 95% confidence interval based on bootstrap resampling. For the cross sections, the horizontal axis, waypoint markers, and white-outlined rectangle identifying the turbulent patch are as in Fig. 4.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
d. Vertical heat fluxes
The Osborn–Cox relation is based on an assumed balance between the turbulent production of thermal fluctuation variance and its dissipation through molecular diffusion. In regions of low ReB, we do not expect significant turbulent production, and the balance assumed in the Osborn–Cox relation is unlikely to hold (e.g., Gregg 1975). Therefore, in estimating the temperature diffusivity, we apply the Osborn–Cox relation [Eq. (8)] only to those observations with ReB ≥ 10. For all remaining observations, we set the temperature diffusivity to the molecular value
5. Discussion: Mixing processes
a. Deep tidal mixing
In addition to geographic variability (section 4a), the deep ε field has a systematic temporal signal. This is seen in the power density spectrum of the ε observations (Fig. 8a), constructed by neglecting spatial variability and treating the glider measurements as a simple time series. Here we analyze a time series, shown in Fig. 8b, constructed by geometrically depth-averaging all ε observations deeper than 100-m depth, interpolating to a 15-min grid, and filtering to remove temporal variability on scales smaller than 2 h. To construct the spectrum, we use Welch’s method using 4-day segments of data, 50% overlapped and Hamming-windowed. The most notable feature in the ε power spectrum is a rounded peak at frequencies spanning 1.3–2.4 cpd, straddling both the M2 tidal frequency, 1.93 cpd, and the local inertial frequency, f = 1.90 cpd. The spectral peak thus suggests that the dominant mode of temporal variability in ε is linked to the semidiurnal tide, inertial forcing, or some combination of both.

(a) Power density spectrum of deep ε, constructed using Welch’s method and 4-day segments of data. Gray shading indicates the 95% confidence interval. The M2 tidal frequency and f are indicated by vertical lines. (b) The ε time series used to construct the power density spectrum. The series is made from the geometric cast averages of ε for all depths greater than 100 m and is interpolated to a 15-min grid. Variability on scales smaller than 2 h has been removed.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

(a) Power density spectrum of deep ε, constructed using Welch’s method and 4-day segments of data. Gray shading indicates the 95% confidence interval. The M2 tidal frequency and f are indicated by vertical lines. (b) The ε time series used to construct the power density spectrum. The series is made from the geometric cast averages of ε for all depths greater than 100 m and is interpolated to a 15-min grid. Variability on scales smaller than 2 h has been removed.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
(a) Power density spectrum of deep ε, constructed using Welch’s method and 4-day segments of data. Gray shading indicates the 95% confidence interval. The M2 tidal frequency and f are indicated by vertical lines. (b) The ε time series used to construct the power density spectrum. The series is made from the geometric cast averages of ε for all depths greater than 100 m and is interpolated to a 15-min grid. Variability on scales smaller than 2 h has been removed.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Because of the close proximity of the frequencies of semidiurnal tides and the frequency of inertial currents at these latitudes, distinguishing internal tidal waves from inertial waves in this region is difficult. A number of results, however, suggest that tides are the dominant process setting the deep ε variability seen in Fig. 8. These include first the finding that no analogous peak in the ε power spectrum exists for observations shallower than 100-m depth (not shown); that is, the signal at the inertial and M2 frequencies is only prevalent in the deeper measurements. It seems unlikely that the forcing of this deep signal originates at the surface (cf. Lincoln et al. 2016). Second, acoustic Doppler current profiler (ADCP) measurements from a nearby mooring (ArcticNet 2018 mooring CA08; Fig. 1b) show a slow and steady modulation of the barotropic (here depth-averaged between 10- and 170-m depth) velocity amplitude (Fig. 9a), suggesting that local current variability is predominantly tidal and not wind driven. Third, rotary spectra of the depth-averaged current velocities decomposed into tidal and nontidal components using tidal harmonic analysis (Pawlowicz et al. 2002) indicate that energy in both the clockwise and counterclockwise components centered at the f and M2 frequencies is dominated by the energy in the tidal signal (Fig. 9b). Our analysis of the spatial distribution of turbulent dissipation (Fig. 5) further suggests a link between enhanced turbulence and topographic features, and thus a credible role of tidal modulation. Last, the directionality of both the deep tidal flow and the low-frequency residual current suggest that local tidally driven mixing is plausibly an important process at this site. Decomposing the deep currents into high-frequency and residual flows using a scale separation of 1.3 cpd, we find that the high-frequency flow (dominated by the tides, and accounting for 23% of the total variance) is predominantly aligned with the major axis of the Amundsen Gulf, as is the low-frequency residual flow that exceeds the tidal flow in magnitude (Fig. 9c). This setup is significant, because it implies that our study site is downstream of a region of complex topography offshore of the southern tip of Banks Island (Fig. 1a) roughly 2 times per day (i.e., the semidiurnal M2 frequency).

(a) Depth-averaged current velocity components U (gray) and V (black) measured by ArcticNet mooring CA08 between 10- and 170-m depths. The light-gray shading indicates the period of the glider deployment. (b) Rotary spectra of the U and V records, as well as their decomposition into tidal and nontidal components, as indicated in the legend. The M2 tidal frequency and f are indicated as in Fig. 8. (c) Polar histograms with current speeds of the U and V records, decomposed into high-frequency and residual components. High frequencies are defined as those greater than 1.3 cpd and are dominated by the M2 tide. The approximate orientation of Amundsen Gulf’s major axis, azimuth 305°, is indicated in each histogram by the yellow line. The percentage on each histogram’s perimeter is the tick label for the radial axis (relative occurrence).
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

(a) Depth-averaged current velocity components U (gray) and V (black) measured by ArcticNet mooring CA08 between 10- and 170-m depths. The light-gray shading indicates the period of the glider deployment. (b) Rotary spectra of the U and V records, as well as their decomposition into tidal and nontidal components, as indicated in the legend. The M2 tidal frequency and f are indicated as in Fig. 8. (c) Polar histograms with current speeds of the U and V records, decomposed into high-frequency and residual components. High frequencies are defined as those greater than 1.3 cpd and are dominated by the M2 tide. The approximate orientation of Amundsen Gulf’s major axis, azimuth 305°, is indicated in each histogram by the yellow line. The percentage on each histogram’s perimeter is the tick label for the radial axis (relative occurrence).
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
(a) Depth-averaged current velocity components U (gray) and V (black) measured by ArcticNet mooring CA08 between 10- and 170-m depths. The light-gray shading indicates the period of the glider deployment. (b) Rotary spectra of the U and V records, as well as their decomposition into tidal and nontidal components, as indicated in the legend. The M2 tidal frequency and f are indicated as in Fig. 8. (c) Polar histograms with current speeds of the U and V records, decomposed into high-frequency and residual components. High frequencies are defined as those greater than 1.3 cpd and are dominated by the M2 tide. The approximate orientation of Amundsen Gulf’s major axis, azimuth 305°, is indicated in each histogram by the yellow line. The percentage on each histogram’s perimeter is the tick label for the radial axis (relative occurrence).
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Further support for the interpretation that ε is tidally modulated comes from the results of past studies in the region. We expect semidiurnal currents in the region to be strongly influenced by the locally generated baroclinic tide (Kulikov et al. 2010); the shelf slope north of nearby Cape Bathurst has previously been identified as a likely region of strong internal tide generation (Kulikov et al. 2004). A detailed analysis of tidal currents in the region by Kulikov et al. (2004) concluded that over 70% of the total energy of semidirunal currents was associated with baroclinic coherent tidal currents, as compared with only ~7% in association with inertial components. Furthermore, although the latitude of our study is southward of the critical latitude for the M2 tide, it has been suggested that an M2 internal tide generated in this region becomes resonantly trapped between the continent and the critical latitude (Kulikov et al. 2004, 2010). Ultimately, internal tides in this region are expected to dissipate near their generation site (Morozov and Pisarev 2002; Kulikov et al. 2004, 2010). We further note that observations linking tides and topography to mixing have been recently reported for the broader Arctic Ocean (Rippeth et al. 2015, 2017).
b. Double diffusion
Even when the mechanical energy inputs to mixing do not support enhanced turbulent mixing of buoyancy (section 4b), double diffusion can act to produce turbulent transports of temperature and salt (Radko 2013). The susceptibility of a water column to the diffusive regime of double diffusion can be characterized by the gradient density ratio Rρ that is defined in section 4c. Empirically, diffusive convection in the Arctic Ocean is most commonly observed when 1 < Rρ < 7; it is also sometimes seen when 7 < Rρ < 10, but it is not typically observed when Rρ > 10 (Shibley et al. 2017). In the central Canada Basin’s warm halocline, Rρ is typically 6.3 ± 1.4, and coherent double diffusive staircases are observed over horizontal scales exceeding 1000 km (Timmermans et al. 2008a; Shibley et al. 2017).
We calculate Rρ from our measurements using background gradients filtered to exclude vertical scales smaller than 5 m (Fig. 10). We find that 1 < Rρ < 10 in 21% of observations; 19% are in the range 7–10, and 2% are in the range 1–7. Instances in which 1 < Rρ < 10 are almost exclusively in a band near the top of the Atlantic Water layer: in the potential density anomaly band σ0 = 28.5–29.5 kg m−3, corresponding approximately to the depth range ~200–335 m, 70% of Rρ observations are in the range 1–10. There is also a notable number of small Rρ values in the eddy, where 16% of Rρ observations are in the range 1–10.

Geometric mean vertical profile and horizontal cross section of the density ratio Rρ. In the cross section, data are discretized into three regimes: susceptible to double diffusion (red: Rρ ≤ 7), marginally susceptible (yellow: 7 < Rρ ≤ 10), and not susceptible (purple: Rρ > 10). The value Rρ =10 is shown in the mean profile by the yellow vertical line.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

Geometric mean vertical profile and horizontal cross section of the density ratio Rρ. In the cross section, data are discretized into three regimes: susceptible to double diffusion (red: Rρ ≤ 7), marginally susceptible (yellow: 7 < Rρ ≤ 10), and not susceptible (purple: Rρ > 10). The value Rρ =10 is shown in the mean profile by the yellow vertical line.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Geometric mean vertical profile and horizontal cross section of the density ratio Rρ. In the cross section, data are discretized into three regimes: susceptible to double diffusion (red: Rρ ≤ 7), marginally susceptible (yellow: 7 < Rρ ≤ 10), and not susceptible (purple: Rρ > 10). The value Rρ =10 is shown in the mean profile by the yellow vertical line.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
Despite conditions near the top of the AW layer that suggest that the density structure there could be favorable to double diffusion, we do not observe pervasive double diffusive staircases like those observed in the central Canada Basin’s thermocline, although we do see sporadically dispersed individual temperature steps that are likely related to double diffusive processes. It appears, therefore, that double diffusion does not play a substantial role in the broader vertical transport of heat or density out of the thermocline in this region. This finding is potentially surprising at first because it is often thought that the absence of a double diffusive staircase implies energetic turbulent mixing (e.g., Guthrie et al. 2017; Shibley and Timmermans 2019), and the turbulent mixing estimates from our dataset are typically weak.
The absence of staircases could be due to a number of reasons. For example, more intense mixing upstream or earlier in time could plausibly explain staircase absence, especially given the region of complex topography upstream of the survey site and the variability of the low-frequency flow recorded by the CA08 mooring. It is also plausible that Rρ is generally too large to lead to double-diffusive staircases, even in such a low-mixing environment. This situation is well documented in a number of deep lakes, where levels of
c. Pacific water mesoscale and smaller features
One of the most striking features in our observations is the large variability in the temperature structure of the Pacific Water layer, visible in an enlarged view of the Conservative Temperature cross section (Fig. 11a). The most obvious feature here is the anticyclonic warm-core mesoscale eddy between along-track distances of s = 53 km and s = 100 km and depths of 40 and 110 m. The eddy has an approximate height of 70 m and an approximate diameter of 38 km; the latter dimension was estimated using the glider flight model of Merckelbach et al. (2019) to estimate the distance the glider traveled through water, and by assuming that the glider intersected the eddy’s central core (see the appendix for details). With a maximum Conservative Temperature of −0.1°C observed at ~50-m depth, the eddy is about 1.3°C warmer than the ambient water at its core. It appears to have at least one outer tendril, transected by the glider twice at s = 104 km and s = 139 km along-track distance.

(a) An enlarged view of the spatial cross section of Θ showing the cold halocline and Pacific Water layers, highlighting the eddy as well as smaller, O(1) km, temperature anomalies. The vertical dashed white lines correspond, from left to right, to the three Θ–SA lines shown in the three diagrams in (b). (b) The Θ–SA diagrams for the three vertical profiles indicated (a). Gray dots are all of the data shown in (a). Dotted lines are potential density anomaly contours.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1

(a) An enlarged view of the spatial cross section of Θ showing the cold halocline and Pacific Water layers, highlighting the eddy as well as smaller, O(1) km, temperature anomalies. The vertical dashed white lines correspond, from left to right, to the three Θ–SA lines shown in the three diagrams in (b). (b) The Θ–SA diagrams for the three vertical profiles indicated (a). Gray dots are all of the data shown in (a). Dotted lines are potential density anomaly contours.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
(a) An enlarged view of the spatial cross section of Θ showing the cold halocline and Pacific Water layers, highlighting the eddy as well as smaller, O(1) km, temperature anomalies. The vertical dashed white lines correspond, from left to right, to the three Θ–SA lines shown in the three diagrams in (b). (b) The Θ–SA diagrams for the three vertical profiles indicated (a). Gray dots are all of the data shown in (a). Dotted lines are potential density anomaly contours.
Citation: Journal of Physical Oceanography 51, 1; 10.1175/JPO-D-20-0057.1
The origin of the eddy is unknown, but its large Conservative Temperature anomaly and Absolute Salinity range of 32–33 g kg−1 (Fig. 11b, right panel) suggest that it is composed of summer Bering Seawater (sBSW) that has been modified on the Chukchi shelf (Timmermans et al. 2014). Although the pathways of Chukchi shelf waters into the Beaufort Sea are highly variable, sBSW can form a component of the eastward shelfbreak jet (von Appen and Pickart 2012). These authors note that it is unlikely for sBSW to enter the Canadian Arctic Archipelago as a well-defined jet; rather it is expected that sBSW will enter in the form of mesoscale eddies derived from the jet through baroclinic and barotropic instabilities, consistent with what we observe. This life history suggests that the eddy may have originated from as far as along the shelf-break north of Alaska. Other known locations of mesoscale eddy formation in the Pacific Water layer include Cape Bathurst (Williams and Carmack 2008; Sévigny et al. 2015) and Mackenzie Canyon (Williams et al. 2006).
Irrespective of its possible origin, the eddy represents a significant amount of heat for the Amundsen Gulf. Assuming that the eddy is symmetric about a central axis, we estimate its total heat content to be 180 PJ (1 PJ = 1015 J) relative to the ambient water, taken to be Θ = −1.37°C (see the appendix for details). When spread over the approximate area of Amundsen Gulf (taken as 400 km × 150 km), the heat from this eddy alone would be capable of producing a 1.0-cm depth of ice melt (see section 3), corresponding to 0.5%–1.7% of the late spring sea ice pack (Peterson et al. 2008). However, our observations suggest that turbulence in the Pacific Water layer is extremely weak and is generally insufficient to effect the upward turbulent transport of this heat to the sea ice cover above (section 4d). The absence of turbulence at these depths is reasonably expected to lead to the long-lived persistence of temperature anomalies in the PW layer such as the mesoscale eddy observed here.
Equally striking as the eddy is the presence of substantial temperature variability on horizontal scales of O(1) km seen throughout the PW layer. This variability can be seen in Fig. 11a as a series of blotches superimposed on the ambient PW outside of the influence of the eddy. Temperature–salinity characteristics of these smaller-scale structures (Fig. 11b, middle panel) appear to be sufficiently dissimilar from those of the eddy that we suggest that they are distinct features, rather than tendrils of the eddy. It is unclear how these smaller structures were created. However, in light of recent results by Sévigny et al. (2015), who linked horizontal temperature structure above 100-m depth in Amundsen Gulf to submesoscale frontal formation and isopycnal outcropping at Cape Bathurst, it is possible that we are observing remnant features of nearby submesoscale dynamics.
6. Conclusions
Using a unique, near-continuous, 11-day glider-based sampling of turbulence and hydrography in Amundsen Gulf, a detailed quantification of turbulence and mixing led us to the following conclusions.
Consistent with previous studies in the region, we find that turbulence is typically very weak. In our dataset, the turbulent kinetic energy dissipation rate, ε, has a geometric mean value of 2.8 × 10−11 W kg−1 and is less than 1 × 10−10 W kg−1 in 68% of observations.
Further, we find that weak turbulence and strong stratification combine to predominantly inhibit turbulent diapycnal mixing of buoyancy in the region. The presence of turbulence strong enough to drive vertical overturns and an enhanced buoyancy flux above that expected from molecular diffusion, defined by the criteria ReB ≥ 10, is found in only 7% of observations.
Despite typically weak turbulent levels, our dataset also shows that ε is highly variable, despite the limited time and space spanned by our measurements. Variability in ε in our observations spans five orders of magnitude.
An important implication of this variability is that a small number of strongly turbulent mixing events are disproportionately important in determining net buoyancy fluxes. These rare energetic turbulent events are responsible for enhancing the arithmetic mean diffusivity of density by three orders of magnitude above molecular levels.
The tightly resolved measurements of turbulence and hydrography also provided insight into the relevance of various possible turbulence-forcing mechanisms. We conclude the following.
Turbulent dissipation below the Pacific Water layer is modulated in time at the semidiurnal frequency and shows structure fixed in space across repeat transects over the continental shelf slope, suggesting that the deep turbulent field is tidally forced. Regional bathymetry and currents support the plausibility of this mechanism.
Despite weak turbulence levels and a density structure that is potentially favorable to double diffusion at the top of the Atlantic Water layer, we do not observe well-formed double diffusive structures. Thus, double diffusion does not appear to play a substantial role in the vertical transport of heat in this region at this time. However, the magnitude of the turbulent heat fluxes we estimate are comparable to those in the Canada Basin diffusive staircases, of order 0.1 W m−2.
Significant heat is present in the Pacific Water layer in the form of a mesoscale eddy and smaller scale structures; however, turbulent levels appear to be insufficient to cause a significant turbulent transport of this stored heat to shallower depths.
Given that the dominant forcing mechanism of deep turbulent mixing in the gulf has been proposed to occur through tide–topography interaction, it is relevant to pose the question of how representative our characterization of turbulence and mixing is for broader space and time scales than those spanned by our observations. The spatial variability of turbulence evident in our observations points to a possibly greater role played by regions of complex topography in setting the net mixing rate in the region, raising the question of whether broader scale measurements are needed to achieve a spatial sampling that leads to representative mixing statistics. Although the glider sampling provides improved statistics of the intermittent turbulent field, we also expect significant variability over longer time scales due to changes in wind and tidal forcing, residual current strength, ice conditions, and stratification, all of which have been shown to modulate internal wave energy and turbulence in the broader region (e.g., Dosser et al. 2014; Chanona et al. 2018; Cole et al. 2018). Low turbulence levels that these observations document are consistent with finescale estimates of diffusivity in the Beaufort Sea in spring and summer over multiple years reported by Guthrie et al. (2013), but appear up to two orders of magnitude weaker than those measured in the gulf in early winter (Bourgault et al. 2011), the prime season for ventilating the mixed layer.
The role of turbulent mixing and its space–time variability in setting the vertical nitrate flux is important for Arctic Ocean primary production (e.g., Bourgault et al. 2011; Tremblay et al. 2015; Randelhoff and Guthrie 2016) and is thus particularly important to understand in the habitat of the Cape Bathurst Polynya (Stirling 1980, 1997) given the low turbulent mixing rates characterizing this region. Randelhoff and Guthrie (2016) report that regional differences in turbulent mixing and stratification lead to varying responses of primary production to climate change. We suggest that future glider-based sampling of turbulence in Amundsen Gulf and the Arctic Ocean more broadly has the potential to deliver a better understanding of time and space variability in mixing rates and its impacts, from which important insights with regard to the functioning of Arctic ecosystems in a changing climate may be gained.
Acknowledgments
The authors thank Lucas Merckelbach, Keith Lévesque, Dan Kelley, and the crew of the Canadian Coast Guard Ship Amundsen. They also thank Achim Randelhoff and two anonymous reviewers for critically reading the paper and making suggestions that led to significant improvements and clarifications. This work was supported by the National Science and Engineering Research Council of Canada (NSERC) through the Canadian Arctic GEOTRACES program—supported by the Climate Change and Atmospheric Research program (NSERC RGPCC 433848-12)—and the Discovery Grant Program (NSERC-2015-04866); the Helmholtz Foundation through the Polar Regions and Coasts in the Changing Earth System II program; the Alfred P. Sloan Foundation; the University of British Columbia; the Killam Doctoral Scholarships program; the Vanier Canada Graduate Scholarships program; the UBC Four Year Fellowship Program; the Northern Scientific Training Program; and the NSERC Michael W. Smith Foreign Study Supplement. Logistical support was provided by the Ocean Tracking Network (OTN); the Marine Environmental Observation, Prediction and Response (MEOPAR) Network; and the Amundsen Science Program, which are supported by the Canada Foundation for Innovation (CFI) and NSERC. Mooring data were collected on board the Canadian research icebreaker CCGS Amundsen for the Long-Term Ocean Observatory, a project of ArcticNet. MEOPAR and ArcticNet belong to the Networks of Centres of Excellence Program of the Government of Canada.
Data availability statement
ArcticNet data are publicly available on the Polar Data Catalogue. Replication data for this study are available online (https://doi.org/10.14288/1.0368671).
APPENDIX
Calculation of Eddy Geometry and Heat Content
REFERENCES
Aagaard, K., and P. Greisman, 1975: Towards new mass and heat budgets for the Arctic Ocean. J. Geophys. Res., 80, 3821–3827, https://doi.org/10.1029/JC080i027p03821.
ArcticNet, 2018: Amundsen Science Mooring Data Collection. Long-Term Ocean Observatory (LTOO) oceanographic mooring data collected in the Canadian Arctic Beaufort Sea. Accessed 28 August 2018, www.polardata.ca.
Armi, L., D. Hebert, N. Oakey, J. Price, P. L. Richardson, T. Rossby, and B. Ruddick, 1988: The history and decay of a Mediterranean salt lens. Nature, 333, 649–651, https://doi.org/10.1038/333649a0.
Arrigo, K. R., and G. L. van Dijken, 2004: Annual cycles of sea ice and phytoplankton in Cape Bathurst polynya, southeastern Beaufort Sea, Canadian Arctic. Geophys. Res. Lett., 31, L08304, https://doi.org/10.1029/2003GL018978.
Baker, M. A., and C. H. Gibson, 1987: Sampling turbulence in the stratified ocean: Statistical consequences of strong intermittency. J. Phys. Oceanogr., 17, 1817–1836, https://doi.org/10.1175/1520-0485(1987)017<1817:STITSO>2.0.CO;2.
Bouffard, D., and L. Boegman, 2013: A diapycnal diffusivity model for stratified environmental flows. Dyn. Atmos. Oceans, 61-62, 14–34, https://doi.org/10.1016/j.dynatmoce.2013.02.002.
Bourgault, D., C. Hamel, F. Cyr, J. É. Tremblay, P. S. Galbraith, D. Dumont, and Y. Gratton, 2011: Turbulent nitrate fluxes in the Amundsen Gulf during ice-covered conditions. Geophys. Res. Lett., 38, L15602, https://doi.org/10.1029/2011GL047936.
Carmack, E., and R. W. MacDonald, 2002: Oceanography of the Canadian shelf of the Beaufort Sea: A setting for marine life. Arctic, 55, 29–45, https://doi.org/10.14430/arctic733.
Carmack, E., R. W. MacDonald, and J. E. Papadakis, 1989: Water mass structure and boundaries in the Mackenzie shelf estuary. J. Geophys. Res., 94, 18 043–18 055, https://doi.org/10.1029/JC094iC12p18043.
Carmack, E., and Coauthors, 2015: Toward quantifying the increasing role of oceanic heat in sea ice loss in the new Arctic. Bull. Amer. Meteor. Soc., 96, 2079–2105, https://doi.org/10.1175/BAMS-D-13-00177.1.
Chanona, M., S. Waterman, and Y. Gratton, 2018: Variability of internal wave-driven mixing and stratification in Canadian Arctic shelf and shelf-slope waters. J. Geophys. Res. Oceans, 123, 9178–9195, https://doi.org/10.1029/2018JC014342.
Cole, S. T., J. M. Toole, L. Rainville, and C. M. Lee, 2018: Internal waves in the Arctic: Influence of ice concentration, ice roughness, and surface layer stratification. J. Geophys. Res. Oceans, 123, 5571–5586, https://doi.org/10.1029/2018JC014096.
Dickson, D. L., and H. G. Gilchrist, 2002: Status of marine birds of the southeastern Beaufort Sea. Arctic, 55, 46–58, https://doi.org/10.14430/arctic734.
Dosser, H. V., L. Rainville, and J. M. Toole, 2014: Near-inertial internal wave field in the Canada Basin from ice-tethered profilers. J. Phys. Oceanogr., 44, 413–426, https://doi.org/10.1175/JPO-D-13-0117.1.
Emery, W. J., W. G. Lee, and L. Magaard, 1984: Geographic and seasonal distributions of Brunt–Väisälä frequency and Rossby radii in the North Pacific and North Atlantic. J. Phys. Oceanogr., 14, 294–317, https://doi.org/10.1175/1520-0485(1984)014<0294:GASDOB>2.0.CO;2.
Fer, I., A. K. Peterson, and J. E. Ullgren, 2014: Microstructure measurements from an underwater glider in the turbulent Faroe Bank Channel Overflow. J. Atmos. Oceanic Technol., 31, 1128–1150, https://doi.org/10.1175/JTECH-D-13-00221.1.
Fine, E. C., J. A. MacKinnon, M. H. Alford, and J. B. Mickett, 2018: Microstructure observations of turbulent heat fluxes in a warm-core Canada Basin eddy. J. Phys. Oceanogr., 48, 2397–2418, https://doi.org/10.1175/JPO-D-18-0028.1.
Gargett, A., T. Osborn, and P. Nasmyth, 1984: Local isotropy and the decay of turbulence in a stratified fluid. J. Fluid Mech., 144, 231–280, https://doi.org/10.1017/S0022112084001592.
Grebmeier, J. M., and Coauthors, 2006: A major ecosystem shift in the northern Bering Sea. Science, 311, 1461–1464, https://doi.org/10.1126/science.1121365.
Gregg, M. C., 1975: Microstructure and intrusions in the California Current. J. Phys. Oceanogr., 5, 253–278, https://doi.org/10.1175/1520-0485(1975)005<0253:MAIITC>2.0.CO;2.
Gregg, M. C., 1987: Diapycnal mixing in the thermocline: A review. J. Geophys. Res., 92, 5249–5286, https://doi.org/10.1029/JC092iC05p05249.
Gregg, M. C., 1999: Uncertainties and limitations in measuring ε and χT. J. Atmos. Oceanic Technol., 16, 1483–1490, https://doi.org/10.1175/1520-0426(1999)016<1483:UALIMA>2.0.CO;2.
Gregg, M. C., M. H. Alford, H. Kontoyiannis, V. Zervakis, and D. Winkel, 2012: Mixing over the steep side of the Cycladic Plateau in the Aegean Sea. J. Mar. Syst., 89, 30–47, https://doi.org/10.1016/j.jmarsys.2011.07.009.
Guthrie, J. D., J. H. Morison, and I. Fer, 2013: Revisiting internal waves and mixing in the Arctic Ocean. J. Geophys. Res. Oceans, 118, 3966–3977, https://doi.org/10.1002/jgrc.20294.
Guthrie, J. D., I. Fer, and J. H. Morison, 2017: Thermohaline staircases in the Amundsen Basin: Possible disruption by shear and mixing. J. Geophys. Res. Oceans, 122, 7767–7782, https://doi.org/10.1002/2017JC012993.
Harwood, L. A., and I. Stirling, 1992: Distribution of ringed seals in the southeastern Beaufort Sea during late summer. Can. J. Zool., 70, 891–900, https://doi.org/10.1139/z92-127.
Ivey, G. N., K. B. Winters, and J. R. Koseff, 2008: Density stratification, turbulence, but how much mixing? Annu. Rev. Fluid Mech., 40, 169–184, https://doi.org/10.1146/annurev.fluid.39.050905.110314.
Jackson, J. M., E. C. Carmack, F. A. McLaughlin, S. E. Allen, and R. G. Ingram, 2010: Identification, characterization, and change of the near-surface temperature maximum in the Canada Basin, 1993-2008. J. Geophys. Res., 115, C05021, https://doi.org/10.1029/2009JC005265.
Jakobsson, M., and Coauthors, 2012: The International Bathymetric Chart of the Arctic Ocean (IBCAO) Version 3.0. Geophys. Res. Lett., 39, L12609, https://doi.org/10.1029/2012GL052219.
Jones, E. P., 2001: Circulation in the Arctic Ocean. Polar Res., 20, 139–146, https://doi.org/10.1111/j.1751-8369.2001.tb00049.x.
Khon, V. C., I. I. Mokhov, M. Latif, V. A. Semenov, and W. Park, 2010: Perspectives of Northern Sea Route and Northwest Passage in the twenty-first century. Climatic Change, 100, 757–768, https://doi.org/10.1007/s10584-009-9683-2.
Kirkwood, T. B. L., 1979: Geometric means and measures of dispersion. Biometrics, 35, 908–909.
Kulikov, E. A., A. B. Rabinovich, and E. Carmack, 2004: Barotropic and baroclinic tidal currents on the Mackenzie shelf break in the southeastern Beaufort Sea. J. Geophys. Res., 109, C05020, https://doi.org/10.1029/2003JC001986.
Kulikov, E. A., A. B. Rabinovich, and E. Carmack, 2010: Variability of baroclinic tidal currents on the Mackenzie Shelf, the southeastern Beaufort Sea. Cont. Shelf Res., 30, 656–667, https://doi.org/10.1016/j.csr.2009.11.006.
Lincoln, B. J., T. P. Rippeth, Y.-D. Lenn, M.-L. Timmermans, W. J. Williams, and S. Bacon, 2016: Wind-driven mixing at intermediate depths in an ice-free Arctic Ocean. Geophys. Res. Lett., 43, 9749–9756, https://doi.org/10.1002/2016GL070454.
Lueck, R. G., F. Wolk, and H. Yamazaki, 2002: Oceanic velocity microstructure measurements in the 20th century. J. Oceanogr., 58, 153–174, https://doi.org/10.1023/A:1015837020019.
Merckelbach, L., A. Berger, G. Krahmann, M. Dengler, and J. R. Carpenter, 2019: A dynamic flight model for Slocum gliders and implications for turbulence microstructure measurements. J. Atmos. Oceanic Technol., 36, 281–296, https://doi.org/10.1175/JTECH-D-18-0168.1.
Morozov, E., and S. V. Pisarev, 2002: Internal tides at the Arctic latitudes (numerical experiments). Oceanology, 42, 165–173.
Niemi, A., J. Johnson, A. Majewski, H. Melling, J. Reist, and W. Williams, 2012: State of the ocean report for the Beaufort Sea Large Ocean Management Area. Canadian Manuscript Rep. of Fisheries and Aquatic Sciences Rep. 2977, 60 pp.
Ono, N., 1967: Specific heat and heat of fusion of sea ice. Phys. Snow Ice: Proc., 1, 599–610.
Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 83–89, https://doi.org/10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2.
Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3, 321–345, https://doi.org/10.1080/03091927208236085.
Padman, L., and T. M. Dillon, 1987: Vertical heat fluxes through the Beaufort Sea thermohaline staircase. J. Geophys. Res., 92, 10 799–10 806, https://doi.org/10.1029/JC092iC10p10799.
Palmer, M. R., G. R. Stephenson, M. E. Inall, C. Balfour, A. Düsterhus, and J. A. M. Green, 2015: Turbulence and mixing by internal waves in the Celtic Sea determined from ocean glider microstructure measurements. J. Mar. Syst., 144, 57–69, https://doi.org/10.1016/j.jmarsys.2014.11.005.
Pawlowicz, R., B. Beardsley, and S. Lentz, 2002: Classical tidal harmonic analysis including error estimates in MATLAB using T TIDE. Comput. Geosci., 28, 929–937, https://doi.org/10.1016/S0098-3004(02)00013-4.
Peterson, A. K., and I. Fer, 2014: Dissipation measurements using temperature microstructure from an underwater glider. Methods Oceanogr., 10, 44–69, https://doi.org/10.1016/j.mio.2014.05.002.
Peterson, I. K., S. J. Prinsenberg, and J. S. Holladay, 2008: Observations of sea ice thickness, surface roughness and ice motion in Amundsen Gulf. J. Geophys. Res., 113, C060164, https://doi.org/10.1029/2007JC004456.
Post, E., and Coauthors, 2009: Ecological dynamics across the Arctic associated with recent climate change. Science, 325, 1355–1358, https://doi.org/10.1126/science.1173113.
Prowse, T. D., C. Furgal, R. Chouinard, H. Melling, D. Milburn, and S. L. Smith, 2009: Implications of climate change for economic development in northern Canada: Energy, resource, and transportation sectors. Ambio, 38, 272–281, https://doi.org/10.1579/0044-7447-38.5.272.
Radko, T., 2013: Double-Diffusive Convection. Cambridge University Press, 342 pp.
Rainville, L., and P. Winsor, 2008: Mixing across the Arctic Ocean: Microstructure observations during the Beringia 2005 Expedition. Geophys. Res. Lett., 35, L08606, https://doi.org/10.1029/2008GL033532.
Rainville, L., C. Lee, and R. Woodgate, 2011: Impact of wind-driven mixing in the Arctic Ocean. Oceanography, 24, 136–145, https://doi.org/10.5670/oceanog.2011.65.
Randelhoff, A., and J. D. Guthrie, 2016: Regional patterns in current and future export production in the central Arctic Ocean quantified from nitrate fluxes. Geophys. Res. Lett., 43, 8600–8608, https://doi.org/10.1002/2016GL070252.
Rippeth, T. P., B. J. Lincoln, Y.-D. Lenn, J. A. M. Green, A. Sundfjord, and S. Bacon, 2015: Tide-mediated warming of Arctic halocline by Atlantic heat fluxes over rough topography. Nat. Geosci., 8, 191–194, https://doi.org/10.1038/ngeo2350.
Rippeth, T. P., V. Vlasenko, N. Stashchuk, B. D. Scannell, J. A. M. Green, B. J. Lincoln, and S. Bacon, 2017: Tidal conversion and mixing poleward of the critical latitude (an Arctic case study). Geophys. Res. Lett., 44, 12 349–12 357, https://doi.org/10.1002/2017GL075310.
Ruddick, B., A. Anis, and K. Thompson, 2000: Maximum likelihood spectral fitting: The Batchelor spectrum. J. Atmos. Oceanic Technol., 17, 1541–1555, https://doi.org/10.1175/1520-0426(2000)017<1541:MLSFTB>2.0.CO;2.
Scheifele, B., S. Waterman, L. Merckelbach, and J. R. Carpenter, 2018: Measuring the dissipation rate of turbulent kinetic energy in strongly stratified, low-energy environments: A case study from the Arctic Ocean. J. Geophys. Res. Oceans, 123, 5459–5480, https://doi.org/10.1029/2017JC013731.
Schultze, L. K. P., L. M. Merckelbach, and J. R. Carpenter, 2017: Turbulence and mixing in a shallow shelf sea from underwater gliders. J. Geophys. Res. Oceans, 122, 9092–9109, https://doi.org/10.1002/2017JC012872.
Sévigny, C., Y. Gratton, and P. S. Galbraith, 2015: Frontal structures associated with coastal upwelling and ice-edge subduction events in southern Beaufort Sea during the Canadian Arctic Shelf Exchange Study. J. Geophys. Res. Oceans, 120, 2523–2539, https://doi.org/10.1002/2014JC010641.
Shaw, W. J., and T. P. Stanton, 2014: Vertical diffusivity of the Western Arctic Ocean halocline. J. Geophys. Res. Oceans, 119, 5017–5038, https://doi.org/10.1002/2013JC009598.
Shaw, W. J., T. P. Stanton, M. G. McPhee, J. H. Morison, and D. G. Martinson, 2009: Role of the upper ocean in the energy budget of Arctic sea ice during SHEBA. J. Geophys. Res., 114, C06012, https://doi.org/10.1029/2008JC004991.
Shibley, N. C., and M. L. Timmermans, 2019: The formation of double-diffusive layers in a weakly turbulent environment. J. Geophys. Res. Oceans, 124, 1445–1458, https://doi.org/10.1029/2018JC014625.
Shibley, N. C., M. L. Timmermans, J. R. Carpenter, and J. M. Toole, 2017: Spatial variability of the Arctic Ocean’s double-diffusive staircase. J. Geophys. Res. Oceans, 122, 980–994, https://doi.org/10.1002/2016JC012419.
Shih, L. H., J. R. Koseff, G. N. Ivey, and J. H. Ferziger, 2005: Parameterization of turbulent fluxes and scales using homogeneous sheared stably stratified turbulence simulations. J. Fluid Mech., 525, 193–214, https://doi.org/10.1017/S0022112004002587.
Shroyer, E., 2012: Turbulent kinetic energy dissipation in Barrow Canyon. J. Phys. Oceanogr., 42, 1012–1021, https://doi.org/10.1175/JPO-D-11-0184.1.
Smyth, W., J. Nash, and J. Moum, 2005: Differential diffusion in breaking Kelvin-Helmholtz billows. J. Phys. Oceanogr., 35, 1004–1022, https://doi.org/10.1175/JPO2739.1.
Stillinger, D. C., K. N. Helland, and C. W. Van Atta, 1983: Experiments on the transition of homogeneous turbulence to internal waves in a stratified fluid. J. Fluid Mech., 131, 91–122, https://doi.org/10.1017/S0022112083001251.
Stirling, I., 1980: The biological importance of polynyas in the Canadian Arctic. Arctic, 33, 303–315, https://doi.org/10.14430/arctic2563.
Stirling, I., 1997: The importance of polynyas, ice edges, and leads to marine mammals and birds. J. Mar. Syst., 10, 9–21, https://doi.org/10.1016/S0924-7963(96)00054-1.
Timco, G., and R. Frederking, 1996: A review of sea ice density. Cold Reg. Sci. Technol., 24 (1), 1–6, https://doi.org/10.1016/0165-232X(95)00007-X.
Timmermans, M.-L., J. Toole, R. Krishfield, and P. Winsor, 2008a: Ice-Tethered Profiler observations of the double-diffusive staircase in the Canada Basin thermocline. J. Geophys. Res., 113, C00A02, https://doi.org/10.1029/2008JC004829.
Timmermans, M.-L., J. Toole, A. Proshutinsky, R. Krishfield, and A. Plueddemann, 2008b: Eddies in the Canada Basin, Arctic Ocean, observed from ice-tethered profilers. J. Phys. Oceanogr., 38, 133–145, https://doi.org/10.1175/2007JPO3782.1.
Timmermans, M.-L., and Coauthors, 2014: Mechanisms of Pacific summer water variability in the Arctic’s central Canada Basin. J. Geophys. Res. Oceans, 119, 7523–7548, https://doi.org/10.1002/2014JC010273.
Tremblay, J.-E., L. G. Anderson, P. Matrai, P. Coupel, S. Belanger, C. Michel, and M. Reigsta, 2015: Global and regional drivers of nutrient supply, primary production and CO2 drawdown in the changing Arctic Ocean. Prog. Oceanogr., 139, 171–196, https://doi.org/10.1016/j.pocean.2015.08.009.
Turner, J. S., 2010: The melting of ice in the Arctic Ocean: The influence of double-diffusive transport of heat from below. J. Phys. Oceanogr., 40, 249–256, https://doi.org/10.1175/2009JPO4279.1.
von Appen, W.-J., and R. Pickart, 2012: Two configurations of the Western Arctic shelfbreak current in summer. J. Phys. Oceanogr., 42, 329–351, https://doi.org/10.1175/JPO-D-11-026.1.
Wassmann, P., 2011: Arctic marine ecosystems in an era of rapid climate change. Prog. Oceanogr., 90, 1–17, https://doi.org/10.1016/j.pocean.2011.02.002.
Wassmann, P., 2015: Overarching perspectives of contemporary and future ecosystems in the Arctic Ocean. Prog. Oceanogr., 139, 1–12, https://doi.org/10.1016/j.pocean.2015.08.004.
Williams, W. J., and E. Carmack, 2008: Combined effect of wind-forcing and isobath divergence on upwelling at Cape Bathurst, Beaufort Sea. J. Mar. Res., 66, 645–663, https://doi.org/10.1357/002224008787536808.
Williams, W. J., E. Carmack, K. Shimada, H. Melling, K. Aagaard, R. W. Macdonald, and R. G. Ingram, 2006: Joint effects of wind and ice motion in forcing upwelling in Mackenzie Trough, Beaufort Sea. Cont. Shelf Res., 26, 2352–2366, https://doi.org/10.1016/j.csr.2006.06.012.
Wüest, A., T. Sommer, M. Schmid, and J. R. Carpenter, 2012: Diffusive-type of double diffusion in lakes—A review. Environmental Fluid Mechanics: Memorial Volume in Honour of Prof. Gerhard H. Jirka, W. Rodi and M. Uhlmann, Eds., IAHR Monographs, CRC Press, 271–284, https://doi.org/10.1201/b12283.
Note that the geometric mean is defined only for nonzero values, so we set below-detection-limit values to the detection limit (2.0 × 10−12 W kg−1) for geometric mean calculations throughout. We note that arithmetic mean and median values are not sensitive to the treatment of below-detection-limit estimates: these values are unchanged (to the number of significant figures reported) if these estimates are set to zero or the detection limit. In contrast, the geometric mean value can be sensitive to the treatment of below-detection-limit values: for example, the geometric mean value of ε for the full dataset reduces from 2.8 [2.7, 2.8] × 10−11 to 1.7 [1.6, 1.7] × 10−11 W kg−1 if these estimates are set to be the detection limit and an order of magnitude smaller than the detection limit value, respectively.