Articles | Volume 21, issue 5
https://doi.org/10.5194/os-21-2149-2025
https://doi.org/10.5194/os-21-2149-2025
Research article
 | 
01 Oct 2025
Research article |  | 01 Oct 2025

Assessment of ocean bottom pressure variations in CMIP6 HighResMIP simulations

Le Liu, Michael Schindelegger, Lara Börger, Judith Foth, and Junyang Gou
Abstract

Ocean bottom pressure (pb) variations from high-resolution climate model simulations under the CMIP6 (Coupled Model Intercomparison Project Phase 6) HighResMIP protocol are potentially useful for oceanographic and space-geodetic research, but the overall signal content and accuracy of these pb estimates have hitherto not been assessed. Here, we compute monthly pb fields from five CMIP6 HighResMIP models at 1/4° grid spacing over both historical and future time spans and compare these data, in terms of temporal variance, against observation-based pb estimates from a 1/4° downscaled GRACE (Gravity Recovery and Climate Experiment) product and 23 bottom pressure recorders, mostly in the Pacific. The model results are qualitatively and quantitatively similar to the GRACE-based pb variances, featuring – aside from eddy imprints – elevated amplitudes on continental shelves and in major abyssal plains of the Southern Ocean. Modeled pb variance in these regions is  10 %–80 % higher and thus overestimated compared to GRACE, whereas underestimation relative to GRACE and the bottom pressure recorders prevails in more quiescent deep-ocean regions. We also form variance ratios of detrended pb signals over 2030–2049 under a high-emission scenario relative to 1980–1999 for three selected models and find statistically significant increases in future pb variance by  30 %–50 % across deep Arctic basins and the southern South Atlantic. The strengthening appears to be linked to projected changes in high-latitude surface winds and, in the case of the South Atlantic, intensified eddy kinetic energy. The study thus points to possibly new pathways for relating observed pb variability from (future) satellite gravimetry missions to anthropogenic climate change.

Share
1 Introduction

Ocean bottom pressure (pb) measures the weight of the total water column exerted on the seafloor, and thus its variations are inherently linked to mass fluxes and redistribution in the Earth system. For example, contemporary increases in ocean mass and thus static pb result from loss of land ice and continued groundwater depletion (Ludwigsen et al.2024). Dynamic pb signals, on the other hand, are largely driven by atmospheric forcing (e.g., Ponte1999; Boening et al.2011; Petrick et al.2014; Fukumori et al.2015) but may also arise internally within the ocean through non-linear processes (Zhao et al.2021). To the extent the ocean is in geostrophic balance, horizontal pb gradients can be used to infer the barotropic transport variability of deep currents (e.g., Makowski et al.2015) or characteristics of the meridional overturning circulation (Hughes et al.2018). Physical oceanography aside, knowledge of pb variability is also essential in geodesy, particularly for de-aliasing purposes in satellite gravimetry processing (Shihora et al.2022; Schindelegger et al.2021). In addition, ocean mass redistributions excite Earth rotation changes (e.g., Harker et al.2021) and cause crustal loading signals that may mask geophysically more relevant effects in station position time series (van Dam et al.2012).

Works in the noted disciplines have been served well by pb estimates on a global scale from time-resolved satellite gravimetry (i.e., the Gravity Recovery and Climate Experiment, GRACE) and numerical ocean models. Pertinent model diagnostics are typically drawn from data-constrained or free simulations of the real ocean (e.g., Piecuch et al.2015; Hughes et al.2018; Androsov et al.2020; Shihora et al.2022; Ponte et al.2024). With a few exceptions (Ponte et al.2002; Bingham and Haines2005; Landerer et al.2007), coupled climate models have so far played no role in studies of pb variability and its implications. However, such synthetic mass change records are potentially useful as background fields in future gravimetry mission simulations to assess the error characteristics and science return of specific satellite constellations (Daras et al.2024). Moreover, climate models can be used to delineate likely changes in surface mass variability that may form target signals for future satellite gravity missions (cf. Jensen et al.2020). Whether or not there are any significant changes in dynamic pb to expect in a warming world is an interesting question in its own right.

We are thus motivated to take a first look at pb estimates from the latest realization of the Coupled Model Intercomparison Project (CMIP), i.e., CMIP6. Specifically, our objectives are to (i) validate the pb diagnostics from CMIP6 experiments by comparing them, in a statistical sense, to observations, (ii) identify source models that could possibly serve as input for end-to-end satellite simulations, and (iii) test for likely future changes in pb variability under a high-emission scenario. Although the coarse-resolution ( 100 km grid spacing) branch of CMIP6 would suffice for most of these purposes, we here analyze the output of eddy-permitting ( 25 km) ocean models under the CMIP6 HighResMIP (CMIP6-HR) protocol. Horizontal model resolution in this range is generally advantageous for representing topographically constrained dynamics (e.g., Harker et al.2021). It also allows transient eddy features, which have their own unique imprint on pb (Bingham and Hughes2008; Hughes et al.2018; Zhao et al.2021). Eddy activity, in turn, is projected to change profoundly in the 21st century (Beech et al.2022) and therefore may strengthen or weaken pb variability in certain locations. In total, and for practical reasons, we consider five out of the seven complete institutional contributions to the CMIP6-HR activity (cf. Haarsma et al.2016; Roberts et al.2020). The specific time periods examined are 1980–2014, representing recent historical climatic conditions, and 2030–2049, i.e., the near future. Given the monthly sampling of the data and the possibility of spurious model drifts (Sen Gupta et al.2013), we are precluded from studying trends and exclusively analyze pb signals at intraseasonal to interannual periods.

2 Data and methods

2.1 Climate model data

CMIP6-HR is an effort dedicated to the investigation of the role of horizontal resolution in CMIP-type climate simulations (Haarsma et al.2016); see, e.g., Docquier et al. (2019), Roberts et al. (2020), or Shi et al. (2021) for evaluations of these simulations with respect to the ocean. We use the model outputs from the coupled historical runs (labeled as “hist-1950” in HighResMIP), complemented by coupled scenario simulations (“highres-future”) in several cases. The hist-1950 simulations cover the period 1950–2014 and provide the initial conditions for highres-future, nominally integrated to the year 2050. External forcing fields comprise time-varying solar, volcanic, anthropogenic aerosol, ozone depletion, and greenhouse gas effects (Haarsma et al.2016). For the highres-future simulations, these forcings follow the Representative Concentration Pathway 8.5 (i.e., high-end emission) scenario with Shared Socioeconomic Pathway 5 (SSP5, O'Neill et al.2016) conditions, abbreviated as SSP5-8.5.

The five CMIP6-HR models examined in our work are ECMWF-IFS-HR (European Centre for Medium-Range Weather Forecasts Integrated Forecasting System), HadGEM3-GC3.1 (Hadley Centre Global Environment Model 3 – Global Coupled version 3.1), CNRM-CM6-1 (Centre National de Recherches Météorologiques Climate Model version 6), CMCC-CM2 (Euro-Mediterranean Centre on Climate Change Coupled Climate Model version 2), and EC-Earth3P; see Table 1 for an overview. The corresponding abbreviations, used throughout the paper, are ECMWF, HadGEM, CNRM, CMCC, and Earth3P, respectively. Given the scope of our study, we focus on the high-resolution (labeled HR, HM, or VHR4) configuration of each model, comprising horizontal grid spacings of 1/4° ( 25 km) for the ocean and 25 to 50 km for the atmosphere. The selected models differ with regard to the atmospheric component but not for the ocean (Table 1); the ocean component in all five cases is NEMO (Nucleus for European Modelling of the Ocean, either version 3.4 or 3.6, Madec2016), a hydrostatic, primitive equation general circulation model. Processes and properties of the ocean are therefore represented very similarly in all five climate models. Hence, any spread in pb variability reported below rather bears on parameter uncertainty, internal variability, or structural uncertainty in the atmospheric component. Values for selected parameters of the ocean component (e.g., viscosity and diffusivity) can be found in Docquier et al. (2019) and Roberts et al. (2020).

Roberts et al. (2018)Roberts et al. (2019)Voldoire et al. (2019)Cherchi et al. (2019)Haarsma et al. (2020)

Table 1Selected specifications of the CMIP6-HR models used in this studya.

a The analysis window is 1980–2014 (hist-1950 experiment) and, in addition, 2030–2049 for the HadGEM, CNRM, and Earth3P highres-future experiments under SSP5-8.5.
b Abbreviations: IFS (Integrated Forecasting System of the ECMWF), UM GA7.0 (Global Atmosphere 7.0 configuration of the Met Office Unified Model), ARPEGE (Action de Recherche Petite Echelle Grande Echelle), and CAM4 (Community Atmosphere Model Version 4).

Download Print Version | Download XLSX

To deduce each model's bottom pressure, we retrieve monthly mean fields for the potential temperature (standard name “thetao”), salinity (“so”), and sea surface height above geoid (“zos”). Alternatively, we could use the models' diagnostic pb output (standard name “pbo”) without invoking assumptions in the calculation of pb; cf. Sect. 2.2. Not all of the tested models provide “pbo” output, though. In any event, differences between the original and our inferred pb values in terms of their temporal RMS (root-mean-square) are very small (Fig. S1 in the Supplement). For the hist-1950 experiments, we focus on the last 35 simulation years (1980–2014). The time span considered for the highres-future simulations is 2030–2049. These choices were made in view of manageable data volumes, data availability, and the requirement for extended overlap with the pb time series from satellite gravimetry (Sect. 2.3). For complementary analyses and tentative interpretation of the pb results, we additionally process monthly eastward and northward seawater velocities (u, v; standard names “uo”, “vo”) and the eastward and northward components of the near-surface atmospheric wind (“uas”, “vas”, usually in reference to an altitude of 10 m).

2.2 Calculation of ocean bottom pressure anomalies

A mathematical expression for pb, obtained by integrating the hydrostatic balance over full depth along the vertical coordinate z (positive upward), is (e.g., Gill and Niller1973; Ponte1999):

(1) p b = - H 0 ρ g d z + 0 η ρ g d z - H 0 ρ g d z + ρ 0 g η ,

where ρ=ρ(z) is the (in situ) density of seawater with surface value ρ0, η represents the sea level anomaly, H is the local water depth, and g is the acceleration due to gravity. While atmospheric pressure contributes to pb in nature, it is not part of the exchange variables between atmospheric and oceanic components of CMIP6-HR and therefore neglected in Eq. (1). We perform the vertical integration on the native model grid and specifically derive the value of ρ at a given location from the modeled potential temperature and salinity using the Gibbs SeaWater Oceanographic Toolbox (McDougall and Barker2011). The exact evaluation of Eq. (1) from CMIP6-HR data would also require each model's internal representation of the bottom topography, e.g., in the form of a three-dimensional array of wet and dry cell fractions (or partial cell thicknesses; see Adcroft et al.1997; Bernard et al.2006). Given that these fractions are not provided in the CMIP6 archive, we calculate approximate versions of them based on given layer thicknesses and the ETOPO1 bathymetry dataset (Amante and Eakins2009), interpolated to the horizontal model grid. Although the information on the provenance of model bathymetries in CMIP6-HR is uncertain, the HadGEM ocean component description (Storkey et al.2018) explicitly points to ETOPO1, and that is indeed a standard choice for NEMO simulations on the ORCA025 grid (cf. Börger et al.2023).

We reduce the full bottom pressure to dynamic anomalies (also denoted as pb) by subtracting the time-mean pressure and trend values at each grid point, along with the monthly varying spatial average of pb. Global ocean mass changes, which may have physical and non-physical causes in numerical ocean models (e.g., Sen Gupta et al.2013), are therefore not part of the analyzed pb variability. To convey some sense of how this variability is distributed across frequencies while also keeping the description of the results compact, we consider (i) full time series of detrended dynamic pb and (ii) non-seasonal anomalies obtained by removing least-squares fits of the annual and semiannual oscillations from the full time series. Variant (ii) thus emphasizes intraseasonal and interannual pb signals. The seasonal cycle, which has been the subject of several recent GRACE and ocean model studies (e.g., Cheng et al.2021; Chen et al.2023; Ponte et al.2024), is treated only in passing. Note that pb has standard SI units (Pa) in our calculations but is expressed as centimeters of equivalent water height (EWH) below, assuming the same reference density (1025 kg m−3) as in the GRACE data.

All comparisons to observations are conducted based on measures of signal variance, i.e., the temporal RMS or quantities derived therefrom. As will be shown below (Sect. 3.5), the pb RMS at a given location is generally a non-stationary statistic that fluctuates under both constant and evolving climatic conditions. These fluctuations may render modeled and observed pb variability statistically indistinguishable from each other. We indicate such cases in our presentation by testing whether the GRACE-based pb RMS is inside the 10th–90th percentile range of an approximate distribution of model RMS values at the considered grid point. For each model, the distribution is constructed by computing RMS values in rolling 15-year windows, progressively shifted forward by one year from 1980–1994 to 1990–2014. We thus have 21 RMS estimates, drawn from windows that are similar in length to (although somewhat shorter than) the period of the adopted GRACE data. In figures pertaining to the historical simulations, we always plot the median RMS of each model's distribution, which closely agrees with the RMS estimate over the full 1980–2014 period (typically within 0.2 cm).

2.3 GRACE data

The primary dataset for validating the CMIP6-HR bottom pressures is a high-resolution satellite-based pb product (Gou et al.2025), obtained by downscaling available monthly pb anomalies from GRACE/-Follow On (Tapley et al.2019) beyond their typical effective resolution of  3°. This product, called GRACE-DS hereinafter, offers pb information from April 2002 to December 2020 on a 1/4°×1/4° grid. The downscaling was realized using a self-supervised deep-learning algorithm, guided by pb diagnostics from two eddy-permitting ocean reanalyses based on the NEMO model (Lellouche et al.2013; Zuo et al.2017). Spatial detail on sub-GRACE grid scales was obtained by maximizing similarities between the downscaled product and the reanalysis output while constraining the solution to the original GRACE values at larger scales (Gou et al.2025). The data fusion algorithm was further enhanced by supervision signals based on input features, comprising, e.g., wind stress and bathymetry. From the three downscaled solutions provided by Gou et al. (2024), we use the one based on GRACE mascons from the Center for Space Research (CSR, RL06.2, Save et al.2016).

A visual comparison of the GRACE-DS and CSR fields in terms of the RMS of dynamic, detrended pb anomalies is provided in Fig. 1. As per design, the deep-learning algorithm retains the signal content of the CSR parent solution at wavelengths of a few hundred kilometers and longer. At the same time, short-scale features are added throughout the ocean, including marked pb increases in eddy-active regions (e.g., Southern Ocean, western boundary current regions). The downscaling approach also enhances near-shore amplitudes on several continental shelves, usually by 20 %–30 % (e.g., around Australia, China, the Atlantic coast of South America, or the Siberian Shelf). As these spatial details are mainly derived from the NEMO-based ocean reanalyses, the model-data comparison in eddying regions and near continental boundaries should be treated with caution. In the supplement to this article, we thus repeat parts of our model assessment using (i) the CSR RMS map in Fig. 1b and (ii) another pb mascon solution from the Jet Propulsion Laboratory (JPL, RL06.3M, Watkins et al.2015; Wiese et al.2023). Whereas CSR employs the same 1/4°×1/4° grid as GRACE-DS, the JPL mascons are discretized in the form of 3°×3° equal-area caps. Prior to comparing to the JPL data, we average the model pb fields to the exact geometry of these caps (cf. Ponte et al.2024). Such coarsening suppresses most of the mesoscale imprints on pb, which are, in any case, difficult to assess. For the presentation in the main text, we focus on GRACE-DS, as its characteristics are most commensurate with the CMIP6-HR models across scales and from deep to shallow regions.

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f01

Figure 1RMS of monthly pb anomalies (in cm of EWH) from (a) GRACE-DS, obtained by downscaling 1° CSR mascons (Save et al.2016) to a 1/4°×1/4° grid. The RMS of the CSR solution is illustrated in panel (b). Analyzed time series are from April 2002 to December 2020, with trends and spatial mean pb signals removed. White triangles in (b) indicate the 23 BPR stations used in this study, labeled by increasing longitude. The dashed black polygon in the Andaman Sea marks the region excluded from the comparison in Sect. 3 due to the presence of seismic deformation signals in the GRACE fields. Lettered keys in panel (a) denote the (A) Gulf of Thailand, (B) Australian Antarctic Basin, (C) Gulf of Carpentaria, (D) Kuroshio Current, (E) Ross Sea, (F) Canadian Arctic Archipelago, (G) Bellingshausen Basin, (H) Drake Passage, (I) Gulf Stream Extension, (J) Argentine Basin/Zapiola Gyre, (K) Weddell Basin, (L) Celtic Sea, (M) Cape Basin/Agulhas leakage, (N) Enderby Basin, and (O) Agulhas Return Current.

Sources of inconsistency between CMIP6-HR and the GRACE(-DS) data, among other effects, include the absence of barometric pressure-driven pb variability in the models (cf. Piecuch et al.2022 0.5 cm in places) and seismic gravity/deformation signals associated with major earthquakes in the gravity field solutions (e.g., Zhang et al.2020). The largest such contribution to pb, apparent in Fig. 1 around the Malay Peninsula, stems from the 2004 Sumatra–Andaman and 2012 Indian Ocean earthquakes. We approximately delimit the affected region based on insight from literature (De Linage et al.2009) and enhanced ( 2 cm) open-ocean pb values and exclude it from all direct comparisons to the model fields.

2.4 In situ observations

We also compare the CMIP6-HR pb diagnostics with direct seafloor pressure measurements from the DART® (Deep-ocean Assessment and Reporting of Tsunamis) program (Bernard and Meinig2011). DART is a US-led tsunami early warning system comprising more than 50 active deep-ocean bottom pressure recorders (BPRs) located mainly along the Circum-Pacific belt. The DART data are disseminated in meters of EWH with a nominal sampling of 15 min. From the full network, we extract 23 stations with nearly continuous observations over an 11-year time span from 2013 to 2023. The year 2013 is a reasonable start date, as it marks the completion of a major upgrade to the DART technology (Bernard and Meinig2011). Each of the 23 time series is a concatenation of recordings from successive deployments in slightly different locations, typically differing by O(10 m) in the overlying pressure. We process each of these segments individually, removing (i) 3 d of observations at the start and end of the time series, (ii) a fitted second-order polynomial (to account for sensor drift, cf. Poropat et al.2018), (iii) occasional spikes, and (iv) signals of the 16 largest tidal constituents (Hart-Davis et al.2021, including fortnightly, monthly, and semiannual constituents). Steps (i)–(iv) are applied to hourly sampled versions of the segments, which we subsequently splice together and average into a monthly pb series per station.

For consistency with our treatment of the CMIP6-HR fields (Sect. 2.1), the DART data should also be cleaned from the global mean pb signals, reflecting both net atmospheric pressure variations over the ocean and true ocean mass changes (Ponte1993; Johnson and Chambers2013). While the former is readily corrected for using monthly sea level pressures from a modern atmospheric reanalysis (Hersbach et al.2020), precise estimates for the mean ocean mass change over 2013–2023 are harder to come by. However, apart from a trend, these changes mainly consist of an annual oscillation, reaching  0.5 cm in amplitude (Johnson and Chambers2013). We thus remove from the detrended DART data a synthetic time series of the seasonal cycle in ocean mass, constructed from pb output (1992–2017) of an ocean state estimate (the Estimating the Circulation and Climate of the Ocean, Version 4 Release 4b, ECCOv4rb, Forget et al.2015). ECCOv4rb is deemed suitable for this purpose, as it is constrained to most available oceanographic data, including GRACE-based pb and observed global ocean mass changes.

Our final network, illustrated in Fig. 1b, comprises 23 BPRs featuring a mean cumulative record length of 10.8 years, with monthly averages typically built from 260 to 350 d of valid observations per year. The in situ time is, therefore, much shorter than the CMIP6-HR records but sufficiently long to constrain variability at the annual and higher frequencies (cf. Ray et al.2021, for the case of sea level). Interannual pb variability, which is less well represented by these  10-year records, accounts for about 30 % of the variance in the full monthly DART time series. The network also heavily emphasizes the Pacific, with only two stations in the western North Atlantic and one sensor located in the Indian Ocean (North Australian Basin). We nevertheless prefer to perform our analysis based on a homogeneous, consistently processed in situ dataset and refrain from adding other BPR time series, which are usually short in duration ( 1–4 years, Poropat et al.2018; Androsov et al.2020; Schindelegger et al.2021) or confined to special locations (e.g., the Arctic Beaufort Gyre, Kemp et al.2005).

3 Results

3.1 Overview of modeled pb variability

Figure 2 (left column) illustrates the RMS of monthly dynamic pb anomalies, calculated for each of the five CMIP6-HR models over the period 1980–2014 (as the median of 21 RMS values in rolling 15-year windows). The model results are qualitatively and quantitatively very similar to the pb variability in the GRACE-DS product (Fig. 1a), ocean state reconstructions (Piecuch et al.2015; Gou et al.2025), and ocean forward models forced by atmospheric analysis fields (e.g., Hughes et al.2018; Poropat et al.2018; Androsov et al.2020). Among the defining features are a general increase in pb variability from temperate to polar latitudes, elevated amplitudes in several marginal seas (e.g., Hudson Bay, Mediterranean Sea, Red Sea), and the largest signals near continental boundaries. Shallow depths facilitate a vigorous barotropic response to the imposed wind stress (Vinogradova et al.2007), leading to peak values of 10–15 cm in Fig. 2 over, e.g., the Siberian Shelf, Chukchi Sea, Gulf of Thailand, Gulf of Carpentaria, and Baltic Sea. In some of these places, coastally trapped waves contribute to the spreading of locally generated mass anomalies (e.g., Oliver and Thompson2011; Fukumori et al.2015).

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f02

Figure 2Plots on the left show the RMS (cm) of detrended pb time series from five CMIP6-HR models, computed as the median of the RMS distribution over 1980–2014; see Sect. 2.2 for details. Plots on the right show each model's variance ratio, R, relative to the pb variance from GRACE-DS (i.e., R>1 indicates higher bottom pressure variability in the model). Stippling marks 3° × 3° cells where the RMS of GRACE-DS (Fig. 1a) is within the 10th–90th percentile range of the model RMS for at least a third of the contained 1/4° grid points. The black solid line is the 1000 m isobath, and the region of the Andaman Sea anomaly seen in GRACE-DS is masked out.

In the deep ocean, known regions of relatively high barotropic variability are the Arctic Ocean, the subpolar Northwest Pacific (Petrick et al.2014), and the three major abyssal plains in the Southern Ocean (Bellingshausen, Australian–Antarctic, and Weddell–Enderby basins). These plains are encased by nearly closed contours of planetary potential vorticity, thus trapping and amplifying forced barotropic circulations (e.g., Fukumori et al.1998; Weijer and Gille2005; Ponte and Piecuch2014; Weijer2015). The Australian–Antarctic basin is the most energetic of the three regions, with RMS values of  4–6 cm in all five models. Bottom topography also plays a key role in generating the characteristic, near-uniform fluctuation in pb (up to 4 cm, Fig. 2) in the interconnected deep basins of the Arctic Ocean and Nordic Seas. In detail, wind stress along the continental slope results in cross-slope Ekman transport, which creates pb anomalies of opposite sign between the shallow shelf and the adjacent deep ocean (Fukumori et al.2015). The deep-ocean mass anomaly propagates away from its source regions as a trapped Kelvin wave, separates from its shallow counterpart at straits and sills, and eventually equilibrates across the deep Arctic basins that are bounded by gradients of potential vorticity.

Forced barotropic dynamics aside, the CMIP6-HR models also contain an active mesoscale field, which can modulate the pb diagnostics on local to basin-wide scales at intraseasonal and longer periods (Hughes et al.2018; Zhao et al.2021, 2023). Localized imprints of eddies and instabilities are most obvious in western boundary current regions (Agulhas Retroreflection, Kuroshio, Gulf Stream) and broadly along the path of the Antarctic Circumpolar Current (Androsov et al.2020). A stand-out feature in all models, perhaps with the exception of CMCC (Fig. 2d), is a  8 cm G-shaped pb structure in the Argentine Basin. The pattern is the signature of the Zapiola Anticyclone, a well-known recirculation supported by high eddy activity in the western parts of the basin and interactions of that eddy field with bottom topography (de Miranda et al.1999; Hughes et al.2007). The GRACE-DS product suggests a similarly G-shaped structure around the Zapiola Drift (Fig. 1a), albeit inherited from the utilized ocean reanalyses and not the GRACE gravity fields themselves (Fig. 1b).

3.2 Comparison with GRACE: broadband variability

For a more quantitative analysis of the agreement between modeled and observation-based pb fluctuations, we compute the variance ratios, R, defined as

(2) R = σ mod 2 σ obs 2 ,

where σmod2 represents the pb variance of a chosen model and σobs2 is the observed variance at a given location, calculated from GRACE-DS, CSR or JPL mascons, or the BPRs.

Figure 2 (right column) illustrates estimates of R for the full (i.e., unfiltered) pb model time series relative to GRACE-DS, deduced after interpolating σmod2 from the native model grid to the GRACE-DS grid. Despite differences in detail, the five maps have several features in common. In particular, they suggest a tendency for excess model variance (R>1) in areas of enhanced pb amplitudes (Fig. 2, left column), while underestimation by the models (R<1) prevails in less energetic regions. One may argue that the latter is simply due to noise in the chosen GRACE product, dominating the low and mid-latitude RMS in Fig. 1a. However, random errors in a single monthly gravity field solution are typically estimated to be 3–4 cm (e.g., Kvas et al.2019), which limits the noise component in the temporal RMS in Fig. 1 to 0.3 cm (inferred from a pessimistic variance propagation over 192 monthly pb fields). The prevalence of low variance ratios outside energetic regions is therefore more likely the result of systematic effects in the GRACE-DS product (e.g., imperfect corrections for low-degree zonal spherical harmonics, Gałdyn and Sośnica2024) and structural errors in the CMIP6-HR models. Thoughts on possible model limitations are offered in Sect. 3.4. Out of the five cases considered, HadGEM and Earth3P compare most favorably to GRACE-DS; global median variance ratios, R, deduced from all grid points with a model RMS of  3 cm and at depths greater than 1000 m, are 0.81 (HadGEM) and 0.90 (Earth3P). These numbers are somewhat closer to 1 than for the other three models (R 0.72, Table 2).

Table 2Global median variance ratios, R, of the CMIP6-HR models relative to GRACE-DS and BPRsa.

a For each model, we list the global median value of R, relative to GRACE-DS, in four regions, two regions each in deep water (depths > 1000 m) and shallow water (depths < 1000 m). The partitioning in the deep ocean is by model RMS (>3 cm or ≤3 cm), while shelf regions are separated by latitude ϕ (higher or lower than 60°). Values in brackets are the same estimates for the non-seasonal signal content. The last column shows the median variance ratios relative to BPRs.
b When compared to BPRs, GRACE-DS has a median variance ratio of 1.30 for the full time series and 0.92 for non-seasonal pb changes.

Download Print Version | Download XLSX

Shifting the focus to the red (i.e., R>1) patches in Fig. 2, we find that all models have higher pb variance than GRACE-DS in the Northwest Pacific, the Australian–Antarctic Basin, and (to some extent) the Weddell–Enderby abyssal plain. Excess values are least pronounced in CNRM and largest in CMCC, evident also from the global median variance ratios taken over deep energetic regions (R= 1.15 vs. R= 1.75; see Table 2). Excessive pb variance in the CMIP6-HR models is also seen for most continental shelf regions – in subpolar and polar latitudes more so than in the tropics (cf., e.g., the Gulf of Thailand and Indonesian Seas). This geographical dependence is further brought out in Table 2, where we list global median variance ratios in shallow water (H< 1000 m) for latitudes both higher and lower than 60° (a somewhat arbitrarily chosen cut-off). The overestimation of shallow pb variations within 60° from the Equator is rather subtle, bounded by R= 1.30 for HadGEM. The HadGEM value increases to R= 1.95 in polar latitudes, followed by R 1.40 in the other models. Common regions of overly strong pb variations are the East Siberian Shelf, the Bering Sea, and the Antarctic continental shelf.

As for the pb imprints of mesoscale eddies, areas of high and low variance ratios tend to change from model to model; see, e.g., the Drake Passage or the Agulhas Retroreflection and its leakage to the west. However, most eddy-related variability in the GRACE-DS product is propagated through from ocean reanalysis output (Sect. 2.3Gou et al.2025). This limits the validity of the comparison in Fig. 2 at the mesoscale. We nevertheless find a common patch of low variances ratios (R 0.3) in the Northwest Atlantic south of Grand Banks, where the pb signature of eddies in the GRACE-DS product spreads considerably further into the Atlantic interior (Fig. 1a) than in the CMIP6-HR models. The situation bears close resemblance to Fig. 8 in Bernard et al. (2006), which suggests that data-free NEMO simulations on the ORCA025 grid typically underestimate the magnitude and spatial spread of eddy kinetic energy in the Gulf Stream Extension.

The quantitative results above can be tested for robustness by replacing GRACE-DS with either the CSR or JPL mascons as reference data (Figs. S2–S3, Tables S1–S2). These analyses suggest a spatial partitioning of variance ratios that is very similar to Fig. 2. We again observe a tendency for R<1 in many quiet deep-ocean regions and Southeast Asian seas, whereas values of R>1 are common in the Southern Ocean, in the Northwest Pacific, and on continental shelves in polar latitudes. Among the more prominent deviations from Fig. 2 (GRACE-DS) and Fig. S2 (CSR) are increased variance ratios with JPL throughout most of the Atlantic. These differences hint at uncertainties in the gravity field recovery in regions of small signal levels. In addition, the patches of red color are more saturated in the evaluation against CSR mascons, which lack signal content on sub-grid GRACE scales (Fig. 1). Hence, the estimates of R from CSR in energetic deep regions in Table S1 cannot be taken at face value. Variance ratios of the coarsened CMIP6-HR fields relative to the 3° × 3° JPL data offer a cleaner form of comparison, and that analysis does indeed confirm many of the numerical values and model rankings noted in relation to Table 2 (column “deep”). On shelf regions, the situation is mostly reversed, that is, the finer-resolved CSR mascons are better suited as alternative reference data than JPL. Here, Table S1 (column “shelf”) again reveals no substantial changes from the corresponding R values with GRACE-DS in Table 2. Given these successful cross-checks, we employ only the downscaled product for the remaining model-data comparisons.

3.3 Non-seasonal pb variability

Modeled pb variability with annual and semiannual oscillations subtracted is briefly assessed for three of the five available models. We select CNRM and Earth3P, which compare favorably to GRACE-DS for energetic deep-ocean regions in Fig. 2, and HadGEM – a somewhat arbitrary choice but still a counterpoint to the other two models. Global median variance ratios at non-seasonal periods for all five models are compiled in Table 2. As evident from Fig. 3, removal of the seasonal cycle reduces RMS values in deep waters of temperate latitudes to  1 cm. At these magnitudes, comparisons with GRACE-DS are likely limited by data noise and subtle systematic errors (cf. Sect. 3.2). These issues inhibit more solid inferences as to why variance ratios in quiet deep-ocean regions decrease relative to the full time series (Table 2, Fig. 2f–j vs. Fig. 3d–f).

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f03

Figure 3Same as in Fig. 2 but restricted to three models (HadGEM, CNRM, Earth3P) and non-seasonal signals.

Non-seasonal signal levels in many of the more energetic areas in Fig. 3 remain comparable to the full variability, and so do the variance ratios over GRACE-DS. Obvious examples are the Arctic Ocean and Nordic Seas, the Bellingshausen Basin, and several eddy-active regions (e.g., Zapiola Drift, Agulhas leakage, Gulf Stream Extension). By contrast, RMS values for non-seasonal variations are clearly smaller than their broadband counterparts (Fig. 2) in the Gulf of Thailand, Northwest Pacific, Australian–Antarctic Basin, and Weddell–Enderby Abyssal Plain. These changes go along with an improvement of the variance ratios relative to GRACE-DS toward R 1 (even for HadGEM), suggesting that wind-driven variability in abyssal-plain regions is indeed too energetic at seasonal periods in CMIP6-HR. Over mid- and low-latitude continental shelves, estimates of R are largely unaltered as we switch from full to non-seasonal time series, and that behavior is also reflected in the global median statistics (Table 2).

3.4 Intermediate discussion

What are potential reasons for the documented differences between modeled and satellite-based pb variability? As alluded to in Sect. 3.2, the dominance of R<1 in quiescent ocean regions may be attributable to a mixture of data noise, residual systematic effects in the GRACE-DS fields, and mismodeled or entirely absent pb signals in the simulations. One specific process disregarded in CMIP6-HR is the effects of gravitational attraction and loading (GAL, Vinogradova et al.2011). GAL induces ocean mass changes on the order of 0.5 cm, particularly at the annual frequency and near continental boundaries (Ponte et al.2024). Parts of the higher variance in GRACE-DS in the Northern Indian Ocean and off the Amazon Shelf, which adjoin regions of large terrestrial water storage variations, may be due to GAL. We also note that the low model pb amplitudes relative to GRACE-DS in the deep-water tropics and some eastern boundary regions (Fig. 2f–j) coincide with areas where pb variability is influenced by non-linearities and coupling between baroclinic and barotropic modes (Piecuch et al.2015; Hughes et al.2018; Zhao et al.2021). Proper depiction of these processes in the models may pose a challenge and may depend on spatial resolution or certain parameter settings concerning, e.g., the viscosity of the flow.

Deep energetic regions, on the other hand, show a propensity for too vigorous pb signals (R>1). Given the importance of wind stress curl and bottom topography in shaping the dynamics of these deep basins (cf. Sect. 3.1), misrepresentation of either of these factors could reduce the fidelity of the modeled barotropic pb variability. For example, both the horizontal and vertical model resolution may be too coarse to produce the necessary topographic steering and dissipation by form stress (Weijer and Gille2005). Other suspects are imperfect wind stress parameterizations or overly strong surface winds as simulated by each model's atmosphere. Similarly, the enhanced pb variations on several continental shelves (e.g., in the Arctic) at both seasonal and non-seasonal periods point to issues regarding model topography and dissipation mechanisms. In particular, as the barotropic ocean response to wind stress forcing scales with the inverse of both bed friction and water depth (Csanady1982), large pb variations in shelf regions can occur if the models are too shallow or not frictional enough. These processes may also contribute to the inflated modeled pb variability in the shallow parts of the Weddell and Ross Seas (Figs. 23). An arguably more critical factor, though, is the blocking of Antarctic ice shelf cavities in CMIP6-HR. This modeling choice constricts the solution to a narrower body of water than in the real ocean and necessarily distorts the dynamics of the region.

3.5 Comparison with in situ observations

Comparisons against BPRs in terms of signal levels (RMS), variance ratios, and median variance ratios over all 23 sites are illustrated in Fig. 4 and Table 2. The analysis complements the previous validation against GRACE-DS in relatively quiet (RMS  3 cm) deep-ocean regions, emphasizing in particular the Pacific; cf. the BPR locations in Fig. 1b. For additional context, we also include (but do not extensively discuss) point-wise statistics from GRACE-DS in Fig. 4. A first general observation is that the BPRs have signal levels very similar to those present in the model-based and GRACE-DS pb estimates at the tested locations. Exceptions to this behavior are rather nuanced and also depend on whether unfiltered or non-seasonal signals (Fig. 4a–b) are analyzed. Specifically, we can discern a tendency for higher in situ than model pb variance for several sites near the Equator (IDs 1–5 and 19), i.e., latitudes where baroclinic and partly non-linear dynamics govern bottom pressure changes (Piecuch2015; Hughes et al.2018). The phenomenon appears to be common to both the full and filtered time series at most of the noted stations.

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f04

Figure 4Comparison of the CMIP6-HR and GRACE-DS pb variability against BPR observations at 23 locations (see Fig. 1b) in terms of RMS (left panels) and variance ratios (right panels, R=1 indicated by horizontal gray line). For context, the left panels also include the 10th–90th percentile interval of the RMS from one model (Earth3P, red vertical bars). The top row shows the statistics for the full broadband pb variations, while the bottom row is for non-seasonal signals.

Download

Somewhat more puzzling, at least at first glance, is that for all sites in the North and Northeast Pacific (IDs 7–16, and partly elsewhere), the model variance ratios hover around 1–1.5 with the seasonal cycle included (Fig. 4c) but drop well below R=1 once it is eliminated (Fig. 4d). This contrast points to systematic differences between BPRs and models in their representation of seasonal pb signals. At the points analyzed here, seasonal oscillations evidently make up a larger fraction of the overall pb variance in the models than they do in the DART data. We suspect that at a resolution of 1/4°, the model bathymetry is too coarse (or too smooth) to capture the damping effects of small-scale topography on the pb response to seasonal wind stress forcing; cf. Chen et al. (2023) for an in-depth study of the problem in the North Pacific Ocean. Accordingly, the apparent agreement between CMIP6-HR and the BPRs in the broadband case (Fig. 4a) may be partly coincidental; that is, anomalous seasonal variability in the models matches the variance due to non-seasonal local pb fluctuations in the in situ series. Twin experiments with NEMO in the ORCA025 configuration, employing either a standard 1/4° bathymetry or smoothed versions of it, could be used to test this hypothesis.

Adopting a more quantitative view, median variance ratios R over all BPRs range from 0.77 (CMCC) to 1.28 (Earth3P) and decrease to R= 0.47–0.74 after removal of the seasonal cycle (Table 2). The comparison of CMIP6-HR with GRACE-DS on a global scale suggests, in fact, very similar ranges and model rankings. This is a reassuring result that alleviates some of our concerns as to the quality of the satellite-based pb solutions at low signal levels (Sect. 3.3). Consistency in detail is particularly seen along the North and East Pacific margins (IDs 9–20), where both GRACE-DS and the BPRs attribute the lowest variance ratios to either CNRM and CMCC and the highest variance ratios to Earth3P (Figs. 24). Thus, the DART series offer a useful point of comparison, despite their limited spatial coverage and the possible presence of short-scale signals in the records. Joint consideration of the lumped statistics in Table 2 and the GRACE-DS variance ratio maps (Figs. 23) nevertheless suggests that Earth3P outperforms the other models in most dynamical regimes. CNRM is least affected by the common issue of excess pb variance in deep energetic regions, whereas HadGEM comes close to Earth3P in quiescent areas. We proceed with these three models for the remainder of the analysis and note that while other choices would have been possible, the overall pb characteristics in all five models are still very similar.

3.6 Projected changes in pb variability

We now examine whether the CMIP6-HR models suggest any notable changes in pb variability under anthropogenic climate change (Fig. 5). To that end, we compute variance ratios, R, of the detrended pb series from the three previously selected models over the time span 2030–2049 (highres-future) with respect to 1980–1999 (hist-1950 simulations). The mid-points of the two windows are therefore 50 years apart, representing the maximum separation possible with the data at hand. We additionally consider 2030–2049 vs. 1980–1999 variance ratios from the control run (Fig. 5, left column), which was integrated for 150 years with fixed forcing (Haarsma et al.2016). These estimates give an approximate idea of the magnitude of regional pb changes that emerge internally within the simulations in the absence of time-evolving climate conditions. Evidently, the control run variance ratios from the three models have very little in common and are mostly restricted to the range R= 0.7–1.3, with only a few broader statistically significant changes (e.g., in the Bellingshausen Basin for HadGEM). These findings provide useful context for the discussion of climate-driven changes in the simulated pb variance (see below). A region that needs to be interpreted with caution, though, is the deep Arctic Ocean, which sees strengthened pb variability in the control run of all three models. However, that increase is statistically significant only in CNRM.

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f05

Figure 5Ratio of pb variance over the time period 2030–2049 relative to 1980–1999 for HadGEM (a, d), CNRM (b, e), and Earth3P (c, f). Left panels (a)(c) present the variance ratios from each model's control simulation, whereas the right panels (d)(f) show the ratios formed from the scenario and historical simulations. Stipples indicate 3° × 3° cells where the variance ratio is significantly different from unity at 90 % confidence for at least a third of the contained 1/4° grid points. To determine these confidence levels, we used an F-test based on the effective degrees of freedom of the two pb series compared at each grid point. Black contours in all panels encase areas where the model RMS in Fig. 2, smoothed to 200 km length scales, exceeds 3 cm. Note that the logarithmic color scale ranges from log(0.5)=0.693 to log(2.0)= 0.693, with labels indicating values of R.

The maps of R formed using the CMIP6-HR historical and scenario runs (Fig. 5, right column) paint a relatively complex picture of future changes in pb variance. Amplitude decreases typically alternate with increases on sub-basin scales, involving changes in regions with distinct dynamical regimes. Although the three maps show many differences in detail, we can still discern a few common, statistically significant features. One region of interest is the connected deep basin of the Arctic Ocean and Nordic Seas, with widespread variance increases of 40 %–60 %, partly also extending to adjacent continental shelves. The increase in CNRM, though, is smaller than in the control run (Fig. 5b and e) and therefore calls into question the results from HadGEM and Earth3P. To examine the matter further – and given the tight connection of pb fluctuations in the Arctic region to wind forcing (Fukumori et al.2015) – we compute variance ratios RV for the near-surface wind speed from the three CMIP6-HR simulations (RV=σ2030–20492/σ1980–19992). These estimates are shown for the global ocean in Fig. 6 and in a more focused manner for the Arctic Ocean and its neighboring seas in Fig. S4.

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f06

Figure 6Same as in Fig. 5 but for the variance ratio RV of the near-surface wind speed (i.e., the magnitude of the horizontal wind vector). Ratios are again formed for the period 2030–2049 relative to 1980–1999. Time-mean wind speeds do not contribute to the picture. To help pinpoint wind speed changes near the continental slope, the 1000 m isobath is included.

A first pass of the RV maps indicates that while the three models attribute variance gains and reductions to somewhat different geographic locations, the Arctic Ocean and northern North Atlantic are, in general, leading regions for future changes in near-surface wind speed variability. To link these changes to the modeled increase in Arctic pb variance, a look at the results in Fukumori et al. (2015) is enlightening. Their Figs. 9 and 10 reveal the continental slope circling around Northeast Greenland/Canada all the way to the East Siberian Shelf as a spatially connected forcing region for the near-uniform Arctic pb fluctuation. Similarly, the fluctuation can be driven by along-slope winds south of Iceland and the Celtic Sea, but such remote contributions counteract the basin-wide bottom pressure signals excited inside the Arctic domain (Fukumori et al.2015). This type of cancellation may indeed occur in the CNRM scenario simulation (Fig. 5e), which suggests increases in wind speed variance both in the Celtic Sea and over the continental slope off Northeast Greenland (Fig. 6e, Fig. S4e). In contrast, HadGEM, Earth3P, and the CNRM control run see little change in near-surface wind speed variability over remote forcing regions. Rather, patches of RV>1 sit over relevant continental slope sections inside the Arctic region, i.e., northeast of Greenland in Earth3P and CNRM-control, and north of the Canadian Arctic Archipelago in HadGEM. Enhanced variability of the associated wind stress could trigger more frequent or more energetic trapped Kelvin waves (Sect. 3.1), thus amplifying the basin-wide Arctic pb fluctuation in the three models (Fig. 5b, d, f).

Along with some interesting cases of diminishing amplitudes (e.g., Northwest Pacific, Gulf Stream region), we also find an intensification of pb variability under SSP5-8.5 in the western tropical Pacific. North of New Guinea, values of R peak at R 1.3–1.5. The area is characterized by low-magnitude pb signals (Figs. 1 and 2), which are forced by winds, involve a mix of barotropic and baroclinic processes, and are most active at annual and semiannual frequencies (e.g., Qu et al.2008; Piecuch et al.2015; Cheng et al.2021). Indeed, the patch of intensified pb amplitudes in the western tropical Pacific vanishes when variance ratios are formed based on non-seasonal model time series (Fig. S5). We also note that in Fig. 5, both HadGEM and Earth3P show increases in R (non-significant) over large parts of the tropical Pacific Ocean. The patterns could hint at broader changes in the wind-driven component of the seasonal cycle around the Inter-Tropical Convergence Zone. A more definite diagnosis of the processes at work is beyond our scope, though.

The third salient feature in Fig. 5, seen across all three models, is a mid-21st century intensification of  30 %–60 % in pb variance in the southern South Atlantic. The patch of red color sits between the Zapiola Gyre and Cape Basin (cf. Fig. 1a) and is accompanied in the eastward direction by a decrease in pb variability over the Agulhas Return Current. Closer inspection of the corresponding RV maps (Fig. 6) fails to reveal a common, larger-scale change in wind speed variance that would hint at an origin of the pb modifications in forced barotropic dynamics. In particular, it is unclear how the east–west dipole in pb changes centered around the dip of South Africa could arise from variations in surface winds and associated quantities (e.g., wind stress curl).

Instead, we suspect that the pb structures in Fig. 5 around the junction of the Atlantic and Indian oceans largely bear on the projected changes in the regions's eddy kinetic energy (EKE, Beech et al.2022; Wang et al.2024). This is not an unreasonable connection to make, as previous model analyses (Hughes et al.2018; Zhao et al.2021) suggest a clear imprint of mesoscale variability on pb in the areas in question. In terms of processes, global warming is expected to strengthen Southern Hemisphere westerlies and progressively shift them to the south (Deng et al.2022, see also Fig. S6). This change in the large-scale wind forcing will in turn reduce volume transport and EKE in the Agulhas Current while also leading to increases in Agulhas leakage and mesoscale activity in the southern South Atlantic (Biastoch et al.2009; van Sebille et al.2009; Beech et al.2022). In addition, climate change invigorates eddy activity in the western part of the basin (Beech et al.2022), and that too could contribute to the enhanced future pb variance seen in Fig. 5.

We have examined the three CMIP6-HR models for some of the described processes and specifically deduced estimates for the depth-integrated EKE (Ni et al.2023, Sect. S1) in the region. Figure 7 illustrates the resultant changes in EKE by 2030–2049 relative to 1980–1999. As evident from the control run (left column), the full-depth EKE is subjected to considerable changes in distribution and magnitude even under constant climatic conditions. The response under SSP5-8.5 (right column) is nevertheless more pronounced, exhibiting also a good degree of consistency among the three models. Shared characteristics are an increase in eddy activity in and west of the Argentine Basin (with signs of a southward shift in HadGEM and Earth3P) and reduced full-depth EKE along the path of the Agulhas Return Current. Agulhas leakage clearly intensifies by the middle of the 21st century in HadGEM and Earth3P but has not yet emerged in CNRM. Overall, the EKE changes produced by CNRM are rather subdued, and so are the changes in the model's pb variance in the analyzed region (Fig. 5e). In comparing the variations in oceanic EKE and pb, note that there is no prerequisite for an exact spatial correspondence between the two variables, as the mesoscale imprint on pb is generally spread out wider than the eddies themselves (Zhao et al.2021). We leave the analysis with these pointers and conclude that the southern South Atlantic and Agulhas Return Current region are an interesting laboratory for linking long-term changes in pb variability with anthropogenic impacts on the ocean circulation.

https://os.copernicus.org/articles/21/2149/2025/os-21-2149-2025-f07

Figure 7Changes in depth-integrated time-mean EKE (J m−2), by 2030–2049 relative to 1980–1999, around the southern South Atlantic. As in Fig. 5, we present results from HadGEM (a, d), CNRM (b, e), and Earth3P (c, f). Left panels (a)(c) show the variance ratios from each model's control simulation, whereas the right panels (d)(f) show the ratios formed from the scenario and historical simulations. The black solid line is the 1000 m isobath. Note the non-linear color scale.

4 Summary and conclusions

We have run basic and mostly global assessments of bottom pressure signals in CMIP6-HR simulations, motivated in large part by the potential use of these fields in satellite gravimetry simulation studies. The climate models capture all relevant pb phenomena that can be expected to emerge at the given resolution and under the encoded physics, e.g., eddy structures and wind-driven variability on continental shelves and in the deep ocean. Model accuracy, as evaluated through comparisons to a downscaled GRACE product and BPR measurements, varies with latitude, depth, frequency, and, more generally, the specific dynamical regime (Figs. 2 and 3, Table 2). In particular, the CMIP6-HR models tested here are susceptible to overestimation of topographically constrained pb variability, most notably in the Australian–Antarctic and Weddell–Enderby Abyssal plains and at seasonal frequencies. Similar issues are apparent in NEMO-based ocean reanalyses on the ORCA025 grid (cf. Fig. 3 in Gou et al.2025), pointing to structural errors in that particular ocean model setup. As suggested in Sect. 3.4, the excess variance may have diverse causes, including imperfect momentum transfer schemes, coarse vertical spacing of deep layers, and the artificial blocking of ice shelf cavities. In this light, analysis of pb variability may be useful to guide model improvements or at least complement the standard tests of near-surface variables and meridional overturning (e.g., Zuo et al.2017; Roberts et al.2019, 2020).

We have exercised a degree of caution in our comparisons of the CMIP6-HR pb fields against GRACE-DS and BPRs. Neither reference dataset is free of error, whereas the models only approximate reality and thus disregard several legitimate contributions to the observed pb changes (e.g., GAL, barometric pressure-driven dynamics, sub grid-scale variability). These issues can compromise the model-data comparison, particularly in very quiet deep-ocean regions (Fig. 3) and at individual BPRs (Fig. 4). The spatial median statistics in Table 2 are nonetheless conclusive and internally consistent enough to discriminate more accurate models from less realistic pb representations. Specifically, in satellite gravimetry simulations, where one would want to portray oceanic mass changes across all spatial and temporal scales, the HighResMIP contribution of Earth3P appears to be a very sensible choice. One minor caveat is that, compared to the models (RMS distribution analyzed over 1980–2014), the BPR and GRACE-DS variance estimates are taken over different time periods, i.e., 2013–2023 and  2002–2020. However, when narrowing the model time series to 2002–2014 (Table S3), our basic conclusions drawn from Table 2 remain unchanged, except for low-latitude shelf regions. The agreement is not unexpected, as the climatic conditions over the CMIP6-HR late historical period evolve less drastically than they do in the scenario extension to 2049, implying temporal changes in pb variance that are well below those shown in Fig. 5d–f.

Our model-based diagnosis of future climate impacts on the strength of pb fluctuations has added a new facet to the existing bottom pressure literature. While these impacts are spread over regions characterized by very different dynamics (i.e., Arctic, tropical Pacific, Agulhas leakage), an origin in altered wind forcing is likely common to all three cases. Dedicated modeling work, such as stand-alone ocean simulations under distinct forcing assumptions, will be required to shed light on the exact processes at work. More generally, the reported strengthening of regional pb variability awaits confirmation through analysis of the output from other climate models. Consideration of a larger ensemble of simulations would also help better distinguish climate-induced pb changes from model internal variability (cf. the case of CNRM in Sect. 3.6). At some point, it might even be possible to detect the projected pb changes in actual observations. Here, we have made our inferences based on variance metrics in bi-decadal windows 50 years apart, but the noted pb amplitude increases in the Arctic, tropical Pacific, and parts of the South Atlantic also emerge when using a reference period shifted forward by 15 years (1995–2014, Fig. S7). Thus, given the planned continuity of space-based mass change monitoring (Daras et al.2024), climate-driven changes in large-scale, dynamic pb variability may be soon detectable with satellite gravimetry.

Data availability

The datasets used in this study are available from the following links: ECMWF CMIP6 HighResMIP output (https://doi.org/10.22033/ESGF/CMIP6.2463, Roberts et al.2017), HadGEM CMIP6 HighResMIP output (https://doi.org/10.22033/ESGF/CMIP6.5989, Roberts2019), CNRM CMIP6 HighResMIP output (https://doi.org/10.22033/ESGF/CMIP6.1375, Voldoire2018), CMCC CMIP6 HighResMIP output (https://doi.org/10.22033/ESGF/CMIP6.1367, Scoccimarro et al.2017), Earth3P CMIP6 HighResMIP output (https://doi.org/10.22033/ESGF/CMIP6.2322, EC-Earth Consortium (EC-Earth)2019), GRACE-DS (https://doi.org/10.3929/ethz-b-000686843, Gou et al.2024), CSR monthly mass grids (https://doi.org/10.15781/cgq9-nh24, Save2020), JPL mascon data (https://doi.org/10.5067/TEMSC-3JC634, Wiese et al.2023), DART® historical data (https://www.ndbc.noaa.gov/historical_data.shtml, National Oceanic and Atmospheric Administration (NOAA)2023), and ECCOv4r4b (https://doi.org/10.5067/ECG5M-OBP4B, ECCO Consortium et al.2021). Monthly pb fields from all five CMIP6-HR models analyzed in this study have been made available at https://doi.org/10.5281/zenodo.14886818 (Liu et al.2025).

Supplement

The supplement related to this article is available online at https://doi.org/10.5194/os-21-2149-2025-supplement.

Author contributions

LL: data curation, formal analysis, investigation, methodology, software, visualization, writing – original draft, writing – review & editing. MS: conceptualization, funding acquisition, methodology, supervision, validation, writing – original draft, writing – review & editing. LB: software, writing – review & editing. JF: formal analysis, software, writing – review & editing. JG: methodology, writing – review & editing.

Competing interests

The contact author has declared that none of the authors has any competing interests.

Disclaimer

Publisher’s note: Copernicus Publications remains neutral with regard to jurisdictional claims made in the text, published maps, institutional affiliations, or any other geographical representation in this paper. While Copernicus Publications makes every effort to include appropriate place names, the final responsibility lies with the authors.

Acknowledgements

This study represents a contribution to the New Refined Observations of Climate Change from Spaceborne Gravity Missions (NEROGRAV) project. We are grateful to the two reviewers for their helpful comments. This publication was supported by the Open Access Publication Fund of the University of Bonn.

Financial support

This research has been supported by the Deutsche Forschungsgemeinschaft (grant nos. 388296632 and 459392861).

Review statement

This paper was edited by John M. Huthnance and reviewed by Christopher Piecuch and one anonymous referee.

References

Adcroft, A., Hughes, C., and Marshall, J.: Representation of topography by shaved cells in height coordinate ocean model, Mon. Weather Rev., 125, 2293–2315, https://doi.org/10.1175/1520-0493(1997)125<2293:ROTBSC>2.0.CO;2, 1997. a

Amante, C. and Eakins, B. W.: ETOPO1 1 Arc-Minute Global Relief Model: Procedures, Data Sources and Analysis, NOAA Technical Memorandum NESDIS NGDC-24, NCEI [data set], https://doi.org/10.7289/V5C8276M, 2009. a

Androsov, A., Boebel, O., Schröter, J., Danilov, S., Macrander, A., and Ivanciu, I.: Ocean bottom pressure variability: Can it be reliably modeled?, J. Geophys. Res.-Oceans, 125, e2019JC015469, https://doi.org/10.1029/2019JC015469, 2020. a, b, c, d

Beech, N., Rackow, T., Semmler, T., Danilov, S., Wang, Q., and Jung, T.: Long-term evolution of ocean eddy activity in a warming world, Nat. Clim. Change, 12, 910–917, https://doi.org/10.1038/s41558-022-01478-3, 2022. a, b, c, d

Bernard, B., Madec, G., Penduff, T., Molines, J.-M., Treguier, A.-M., Le Sommer, J., Beckmann, A., Biastoch, A., Böning, C., Dengg, J., Derval, C., Durand, E., Gulev, S., Remy, E., Talandier, C., Theeten, M., McClean, J., and De Cuevas, B.: Impact of partial steps and momentum advection schemes in a global ocean circulation model at eddy-permitting resolution, Ocean Dynam., 56, 543–567, https://doi.org/10.1007/s10236-006-0082-1, 2006. a, b

Bernard, E. N. and Meinig, C.: History and future of deep-ocean tsunami measurements, in: Proceedings of Oceans'11 MTS/IEEE Kona, 6106894, IEEE, 1–7, https://doi.org/10.23919/OCEANS.2011.6106894, 2011. a, b

Biastoch, A., Böning, C. W., Schwarzkopf, F. U., and Lutjeharms, J. R. E.: Increase in Agulhas leakage due to poleward shift of Southern Hemisphere westerlies, Nature, 462, 495–498, https://doi.org/10.1038/nature08519, 2009. a

Bingham, R. and Hughes, C.: The relationship between sea-level and bottom pressure variability in an eddy-permitting ocean model, Geophys. Res. Lett., 35, L03602, https://doi.org/10.1029/2007GL032662, 2008. a

Bingham, R. J. and Haines, K.: Decadal ocean bottom pressure variability and its associated gravitational effects in a coupled ocean-atmosphere model, in: Gravity, Geoid and Space Missions: GGSM 2004 IAG International Symposium Porto, Springer, 298–303, https://doi.org/10.1007/3-540-26932-0_52, 2005. a

Boening, C., Lee, T., and Zlotnicki, V.: A record-high ocean bottom pressure in the South Pacific observed by GRACE, Geophys. Res. Lett., 38, L04602, https://doi.org/10.1029/2010GL046013, 2011. a

Börger, L., Schindelegger, M., Dobslaw, H., and Salstein, D.: Are ocean reanalyses useful for Earth rotation research?, Earth Space Sci., 10, e2022EA002700, https://doi.org/10.1029/2022EA002700, 2023. a

Chen, L., Yang, J. Y., and Wu, L. X.: Topography effects on the seasonal variability of ocean bottom pressure in the North Pacific Ocean, J. Phys. Oceanogr., 53, 929–941, https://doi.org/10.1175/JPO-D-22-0140.1, 2023. a, b

Cheng, X., Ou, N. S., Chen, J. J., and Huang, R. X.: On the seasonal variations of ocean bottom pressure in the world oceans, Geosci. Lett., 8, 29, https://doi.org/10.1186/s40562-021-00199-3, 2021. a, b

Cherchi, A., Fogli, P. G., Lovato, T., Peano, D., Iovino, D., Gualdi, S., Masina, S., Scoccimarro, E., Materia, S., Bellucci, A., and Navarra, A.: Global mean climate and main patterns of variability in the CMCC-CM2 coupled model, J. Adv. Model. Earth Sy., 11, 185–209, https://doi.org/10.1029/2018MS001369, 2019. a

Csanady, G. T.: Circulation in the Coastal Ocean, Springer, https://doi.org/10.1007/978-94-017-1041-1, 1982. a

Daras, I., March, G., Pail, R., Hughes, C. W., Braitenberg, C., Güntner, A., Eicker, A., Wouters, B., Heller-Kaikov, B., Pivetta, T., and Pastorutti, A.: Mass-change And Geosciences International Constellation (MAGIC) expected impact on science and applications, Geophys. J. Int., 236, 1288–1308, https://doi.org/10.1093/gji/ggad472, 2024. a, b

De Linage, C., Rivera, L., Hinderer, J., Boy, J.-P., Rogister, Y., Lambotte, S., and Biancale, R.: Separation of coseismic and postseismic gravity changes for the 2004 Sumatra–Andaman earthquake from 4.6 yr of GRACE observations and modelling of the coseismic change by normal-modes summation, Geophys. J. Int., 176, 695–714, https://doi.org/10.1111/j.1365-246X.2008.04025.x, 2009. a

de Miranda, A. P., Barnier, B., and Dewar, W. K.: On the dynamics of the Zapiola Anticyclone, J. Geophys. Res.-Oceans, 104, 21137–21149, https://doi.org/10.1029/1999JC900042, 1999. a

Deng, K., Azorin-Molina, C., Yang, S., Hu, C., Zhang, G., Minola, L., and Chen, D.: Changes of Southern Hemisphere westerlies in the future warming climate, Atmos. Res., 270, 106040, https://doi.org/10.1016/j.atmosres.2022.106040, 2022. a

Docquier, D., Grist, J. P., Roberts, M. J., Roberts, C. D., Semmler, T., Ponsoni, L., Massonnet, F., Sidorenko, D., Sein, D. V., Iovino, D., Bellucci, A., and Fichefet, T.: Impact of model resolution on Arctic sea ice and North Atlantic Ocean heat transport, Clim. Dynam., 53, 4989–5017, https://doi.org/10.1007/s00382-019-04840-y, 2019. a, b

EC-Earth Consortium (EC-Earth): EC-Earth-Consortium EC-Earth3P model output prepared for CMIP6 HighResMIP, Earth System Grid Federation [data set], https://doi.org/10.22033/ESGF/CMIP6.2322, 2019. a

ECCO Consortium, Fukumori, I., Wang, O., Fenty, I., Forget, G., Heimbach, P., and Ponte, R. M.: ECCO Ocean Bottom Pressure - Monthly Mean 0.5 Degree (Version 4 Release 4 B), Ver. V4r4b, PO.DAAC [data set], CA, USA, https://doi.org/10.5067/ECG5M-OBP4B, 2021. a

Forget, G., Campin, J.-M., Heimbach, P., Hill, C. N., Ponte, R. M., and Wunsch, C.: ECCO version 4: an integrated framework for non-linear inverse modeling and global ocean state estimation, Geosci. Model Dev., 8, 3071–3104, https://doi.org/10.5194/gmd-8-3071-2015, 2015. a

Fukumori, I., Raghunath, R., and Fu, L. L.: Nature of global large-scale sea level variability in relation to atmospheric forcing: A modeling study, J. Geophys. Res.-Oceans, 103, 5493–5512, https://doi.org/10.1029/97JC02907, 1998. a

Fukumori, I., Wang, O., Llovel, W., Fenty, I., and Forget, G.: A near-uniform fluctuation of ocean bottom pressure and sea level across the deep ocean basins of the Arctic Ocean and the Nordic Seas, Prog. Oceanogr., 134, 152–172, https://doi.org/10.1016/j.pocean.2015.01.013, 2015. a, b, c, d, e, f

Gałdyn, F. and Sośnica, K.: Impact of the combination and replacement of SLR-based low-degree gravity field coefficients in GRACE solutions, Prog. Earth Planet. Sci., 11, 7, https://doi.org/10.1186/s40645-024-00608-z, 2024. a

Gill, A. E. and Niller, P. P.: The theory of the seasonal variability in the ocean, Deep-Sea Res. Oceanogr. Abstr., 20, 141–177, https://doi.org/10.1016/0011-7471(73)90049-1, 1973. a

Gou, J., Börger, L., Schindelegger, M., and Soja, B.: SeDA-OBP: Global ocean bottom pressure anomaly fields with a spatial resolution of 0.25 degrees from self-supervised data assimilation, ETH Research Collection [data set], https://doi.org/10.3929/ethz-b-000686843, 2024. a, b

Gou, J., Börger, L., Schindelegger, M., and Soja, B.: Downscaling GRACE-derived ocean bottom pressure anomalies using self-supervised data fusion, J. Geodesy, 99, 19, https://doi.org/10.1007/s00190-025-01943-9, 2025. a, b, c, d, e

Haarsma, R., Acosta, M., Bakhshi, R., Bretonnière, P.-A., Caron, L.-P., Castrillo, M., Corti, S., Davini, P., Exarchou, E., Fabiano, F., Fladrich, U., Fuentes Franco, R., García-Serrano, J., von Hardenberg, J., Koenigk, T., Levine, X., Meccia, V. L., van Noije, T., van den Oord, G., Palmeiro, F. M., Rodrigo, M., Ruprich-Robert, Y., Le Sager, P., Tourigny, E., Wang, S., van Weele, M., and Wyser, K.: HighResMIP versions of EC-Earth: EC-Earth3P and EC-Earth3P-HR – description, model computational performance and basic validation, Geosci. Model Dev., 13, 3507–3527, https://doi.org/10.5194/gmd-13-3507-2020, 2020. a

Haarsma, R. J., Roberts, M. J., Vidale, P. L., Senior, C. A., Bellucci, A., Bao, Q., Chang, P., Corti, S., Fučkar, N. S., Guemas, V., von Hardenberg, J., Hazeleger, W., Kodama, C., Koenigk, T., Leung, L. R., Lu, J., Luo, J.-J., Mao, J., Mizielinski, M. S., Mizuta, R., Nobre, P., Satoh, M., Scoccimarro, E., Semmler, T., Small, J., and von Storch, J.-S.: High Resolution Model Intercomparison Project (HighResMIP v1.0) for CMIP6, Geosci. Model Dev., 9, 4185–4208, https://doi.org/10.5194/gmd-9-4185-2016, 2016. a, b, c, d

Harker, A. A., Schindelegger, M., Ponte, R. M., and Salstein, D. A.: Modeling ocean-induced rapid Earth rotation variations: An update, J. Geodesy, 95, 110, https://doi.org/10.1007/s00190-021-01555-z, 2021. a, b

Hart-Davis, M. G., Piccioni, G., Dettmering, D., Schwatke, C., Passaro, M., and Seitz, F.: EOT20: a global ocean tide model from multi-mission satellite altimetry, Earth Syst. Sci. Data, 13, 3869–3884, https://doi.org/10.5194/essd-13-3869-2021, 2021. a

Hersbach, H., Bell, B., Berrisford, P., Hirahara, S., Horányi, A., Muñoz-Sabater, J., Nicolas, J., Peubey, C., Radu, R., Schepers, D., Simmons, A., Soci, C., Abdalla, S., Abellan, X., Balsamo, G., Bechtold, P., Biavati, G., Bidlot, J., Bonavita, M., De Chiara, G., Dahlgren, P., Dee, D., Diamantakis, M., Dragani, R., Flemming, J., Forbes, R., Fuentes, M., Geer, A., Haimberger, L., Healy, S., Hogan, R. J., Hólm, E., Janisková, M., Keeley, S., Laloyaux, P., Lopez, P., Lupu, C., Radnoti, G., de Rosnay, P., Rozum, I., Vamborg, F., Villaume, S., and Thépaut, J.-N.: The ERA5 global reanalysis, Q. J. Roy. Astron. Soc., 146, 1999–2049, https://doi.org/10.1002/qj.3803, 2020. a

Hughes, C. W., Stepanov, V. N., Fu, L. L., Barnier, B., and Hargreaves, G. W.: Three forms of variability in Argentine Basin ocean bottom pressure, J. Geophys. Res.-Oceans, 112, C01011, https://doi.org/10.1029/2006JC003679, 2007. a

Hughes, C. W., Williams, J., Blaker, A., Coward, A., and Stepanov, V.: A window on the deep ocean: The special value of ocean bottom pressure for monitoring the large-scale, deep-ocean circulation, Prog. Oceanogr., 161, 19–46, https://doi.org/10.1016/j.pocean.2018.01.011, 2018. a, b, c, d, e, f, g, h

Jensen, L., Eicker, A., Dobslaw, H., and Pail, R.: Emerging changes in terrestrial water storage variability as a target for future satellite gravity missions, Remote Sens., 12, 3898, https://doi.org/10.3390/rs12233898, 2020. a

Johnson, G. C. and Chambers, D. P.: Ocean bottom pressure seasonal cycles and decadal trends from GRACE Release-05: Ocean circulation implications, J. Geophys. Res.-Oceans, 118, 4228–4240, https://doi.org/10.1002/jgrc.20307, 2013. a, b

Kemp, J., Newhall, K., Ostrom, W., Krishfield, R., and Proshutinsky, A.: The Beaufort Gyre observing system 2004: Mooring recovery and deployment operations in pack ice, Woods Hole Oceanographic Institution, Woods Hole Oceanographic Institution, https://www2.whoi.edu/site/beaufortgyre/methods/deployment-procedures/ (last accss: 15 July 2024), 2005. a

Kvas, A., Behzadpour, S., Ellmer, M., Klinger, B., Strasser, S., Zehentner, N., and Mayer-Gürr, T.: ITSG-Grace2018: Overview and evaluation of a new GRACE-only gravity field time series, J. Geophys. Res.-Sol. Ea., 124, 9332–9344, https://doi.org/10.1029/2019JB017415, 2019. a

Landerer, F. W., Jungclaus, J. H., and Marotzke, J.: Ocean bottom pressure changes lead to a decreasing length-of-day in a warming climate, Geophys. Res. Lett., 34, L06307, https://doi.org/10.1029/2006GL029106, 2007. a

Lellouche, J.-M., Le Galloudec, O., Drévillon, M., Régnier, C., Greiner, E., Garric, G., Ferry, N., Desportes, C., Testut, C.-E., Bricaud, C., Bourdallé-Badie, R., Tranchant, B., Benkiran, M., Drillet, Y., Daudin, A., and De Nicola, C.: Evaluation of global monitoring and forecasting systems at Mercator Océan, Ocean Sci., 9, 57–81, https://doi.org/10.5194/os-9-57-2013, 2013. a

Liu, L., Schindelegger, M., and Börger, L.: Monthly ocean bottom pressure anomalies 1980–2014 and 2030–2049 from CMIP6 HighResMIP models, Zenodo [data set], https://doi.org/10.5281/zenodo.14886818, 2025. a

Ludwigsen, C. B., Andersen, O. B., Marzeion, B., Malles, J.-H., Müller Schmied, H., Döll, P., Watson, C., and King, M. A.: Global and regional ocean mass budget closure since 2003, Nat. Commun., 15, 1416, https://doi.org/10.1038/s41467-024-45726-w, 2024. a

Madec, G.: NEMO reference manual 3.6_STABLE: NEMO ocean engine, Institut Pierre-Simon Laplace (IPSL), France, http://hdl.handle.net/10013/epic.46840.d001 (last access: 10 July 2024), 2016. a

Makowski, J. K., Chambers, D. P., and Bonin, J. A.: Using ocean bottom pressure from the gravity recovery and climate experiment (GRACE) to estimate transport variability in the southern Indian Ocean, J. Geophys. Res.-Oceans, 120, 4245–4259, https://doi.org/10.1002/2014JC010575, 2015. a

McDougall, T. J. and Barker, P. M.: Getting started with TEOS-10 and the Gibbs Seawater (GSW) oceanographic toolbox, Scor/Iapso WG, 127, 1–28, ISBN 978-0-646-55621-5, https://www.teos-10.org/pubs/gsw/v3_04/pdf/Getting_Started.pdf (last access: 12 June 2024), 2011. a

National Oceanic and Atmospheric Administration (NOAA): DART Bottom Pressure Recorder (BPR) Historical Data, NOAA National Data Buoy Center (NDBC), https://www.ndbc.noaa.gov/historical_data.shtml (last access: 11 September 2025), 2023. a

Ni, Q., Zhai, X., LaCasce, J. H., Chen, D., and Marshall, D. P.: Full-depth eddy kinetic energy in the global ocean estimated from altimeter and Argo observations, Geophys. Res. Lett., 50, e2023GL103114, https://doi.org/10.1029/2023GL103114, 2023. a

Oliver, E. C. J. and Thompson, K. R.: Sea level and circulation variability of the Gulf of Carpentaria: Influence of the Madden-Julian Oscillation and the adjacent deep ocean, J. Geophys. Res.-Oceans, 116, C02019, https://doi.org/10.1029/2010JC006596, 2011. a

O'Neill, B. C., Tebaldi, C., van Vuuren, D. P., Eyring, V., Friedlingstein, P., Hurtt, G., Knutti, R., Kriegler, E., Lamarque, J.-F., Lowe, J., Meehl, G. A., Moss, R., Riahi, K., and Sanderson, B. M.: The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6, Geosci. Model Dev., 9, 3461–3482, https://doi.org/10.5194/gmd-9-3461-2016, 2016. a

Petrick, C., Dobslaw, H., Bergmann-Wolf, I., Schön, N., Matthes, K., and Thomas, M.: Low-frequency ocean bottom pressure variations in the North Pacific in response to time-variable surface winds, J. Geophys. Res.-Oceans, 119, 5190–5202, https://doi.org/10.1002/2013JC009635, 2014. a, b

Piecuch, C. G.: Bottom-pressure signature of annual baroclinic Rossby waves in the northeast tropical Pacific Ocean, J. Geophys. Res.-Oceans, 120, 2449–2459, https://doi.org/10.1002/2014JC010667, 2015. a

Piecuch, C. G., Fukumori, I., Ponte, R. M., and Wang, O.: Vertical structure of ocean pressure variations with application to satellite-gravimetric observations, J. Atmos. Ocean. Tech., 32, 603–613, https://doi.org/10.1175/JTECH-D-14-00156.1, 2015. a, b, c, d

Piecuch, C. G., Fukumori, I., Ponte, R. M., Schindelegger, M., Wang, O., and Zhao, M.: Low-frequency dynamic ocean response to barometric-pressure loading, J. Phys. Oceanogr., 52, 2627–2641, https://doi.org/10.1175/JPO-D-22-0090.1, 2022. a

Ponte, R. M.: Variability in a homogeneous global ocean forced by barometric pressure, Dynam. Atmos. Oceans, 18, 209–234, https://doi.org/10.1016/0377-0265(93)90010-5, 1993. a

Ponte, R. M.: A preliminary model study of the large-scale seasonal cycle in bottom pressure over the global ocean, J. Geophys. Res.-Oceans, 104, 1289–1300, https://doi.org/10.1029/1998JC900028, 1999. a, b

Ponte, R. M. and Piecuch, C. G.: Interannual bottom pressure signals in the Australian–Antarctic and Bellingshausen basins, J. Phys. Oceanogr., 44, 1456–1465, https://doi.org/10.1175/JPO-D-13-0223.1, 2014. a

Ponte, R. M., Rajamony, J., and Gregory, J. M.: Ocean angular momentum signals in a climate model and implications for Earth rotation, Clim. Dynam., 19, 181–190, https://doi.org/10.1007/s00382-001-0216-6, 2002. a

Ponte, R. M., Zhao, M., and Schindelegger, M.: How well do we know the seasonal cycle in ocean bottom pressure?, Earth Space Sci., 11, e2024EA003661, https://doi.org/10.1029/2024EA003661, 2024. a, b, c, d

Poropat, L., Dobslaw, H., Zhang, L., Macrander, A., Boebel, O., and Thomas, M.: Time variations in ocean bottom pressure from a few hours to many years: In situ data, numerical models, and GRACE satellite gravimetry, J. Geophys. Res.-Oceans, 123, 5612–5623, https://doi.org/10.1029/2018JC014108, 2018. a, b, c

Qu, T., Gan, J., Ishida, A., Kashino, Y., and Tozuka, T.: Semiannual variation in the western tropical Pacific Ocean, Geophys. Res. Lett., 35, L16602, https://doi.org/10.1029/2008GL035058, 2008. a

Ray, R. D., Loomis, B. D., and Zlotnicki, V.: The mean seasonal cycle in relative sea level from satellite altimetry and gravimetry, J. Geodesy, 95, 80, https://doi.org/10.1007/s00190-021-01529-1, 2021. a

Roberts, C. D., Senan, R., Molteni, F., Boussetta, S., Keeley, S.: ECMWF ECMWF-IFS-LR model output prepared for CMIP6 HighResMIP, Earth System Grid Federation [data set], https://doi.org/10.22033/ESGF/CMIP6.2463, 2017. a

Roberts, C. D., Senan, R., Molteni, F., Boussetta, S., Mayer, M., and Keeley, S. P. E.: Climate model configurations of the ECMWF Integrated Forecasting System (ECMWF-IFS cycle 43r1) for HighResMIP, Geosci. Model Dev., 11, 3681–3712, https://doi.org/10.5194/gmd-11-3681-2018, 2018. a

Roberts, M.: MOHC HadGEM3-GC31-MM model output prepared for CMIP6 HighResMIP highres-future, Earth System Grid Federation [data set], https://doi.org/10.22033/ESGF/CMIP6.5989, 2019. a

Roberts, M. J., Baker, A., Blockley, E. W., Calvert, D., Coward, A., Hewitt, H. T., Jackson, L. C., Kuhlbrodt, T., Mathiot, P., Roberts, C. D., Schiemann, R., Seddon, J., Vannière, B., and Vidale, P. L.: Description of the resolution hierarchy of the global coupled HadGEM3-GC3.1 model as used in CMIP6 HighResMIP experiments, Geosci. Model Dev., 12, 4999–5028, https://doi.org/10.5194/gmd-12-4999-2019, 2019. a, b

Roberts, M. J., Jackson, L. C., Roberts, C. D., Meccia, V., Docquier, D., Koenigk, T., Ortega, P., Moreno-Chamarro, E., Bellucci, A., Coward, A., Drijfhout, S., Exarchou, E., Gutjahr, O., Hewitt, H., Iovino, D., Lohmann, K., Putrasahan, D., Schiemann, R., Seddon, J., Terray, L., Xu, X., Zhang, Q., Chang, P., Yeager, S. G., Castruccio, F. S., Zhang, S., and Wu, L.: Sensitivity of the Atlantic Meridional Overturning Circulation to model resolution in CMIP6 HighResMIP simulations and implications for future changes, J. Adv. Model. Earth Sy., 12, e2019MS002014, https://doi.org/10.1029/2019MS002014, 2020. a, b, c, d

Save, H.: CSR GRACE and GRACE-FO RL06 Mascon Solutions v02, Center for Space Research [data set], https://doi.org/10.15781/cgq9-nh24, 2020. a

Save, H., Bettadpur, S., and Tapley, B. D.: High-resolution CSR GRACE RL05 mascons, J. Geophys. Res.-Sol. Ea., 121, 7547–7569, https://doi.org/10.1002/2016JB013007, 2016. a, b

Schindelegger, M., Harker, A. A., Ponte, R. M., Dobslaw, H., and Salstein, D. A.: Convergence of daily GRACE solutions and models of submonthly ocean bottom pressure variability, J. Geophys. Res.-Oceans, 126, e2020JC017031, https://doi.org/10.1029/2020JC017031, 2021. a, b

Scoccimarro, E., Bellucci, A., Peano, D.: CMCC CMCC-CM2-VHR4 model output prepared for CMIP6 HighResMIP, Earth System Grid Federation [data set], https://doi.org/10.22033/ESGF/CMIP6.1367, 2017. a

Sen Gupta, A., Jourdain, N. C., Brown, J. N., and Monselesan, D.: Climate drift in the CMIP5 models, J. Climate, 26, 8597–8615, https://doi.org/10.1175/JCLI-D-12-00521.1, 2013. a, b

Shi, J. R., Talley, L. D., Xie, S. P., Peng, Q. H., and Liu, W.: Ocean warming and accelerating Southern Ocean zonal flow, Nat. Clim. Change, 11, 1090–1097, https://doi.org/10.1038/s41558-021-01212-5, 2021. a

Shihora, L., Balidakis, K., Dill, R., Dahle, C., Ghobadi-Far, K., Bonin, J., and Dobslaw, H.: Non-tidal background modeling for satellite gravimetry based on operational ECWMF and ERA5 reanalysis data: AOD1B RL07, J. Geophys. Res.-Sol. Ea., 127, e2022JB024360, https://doi.org/10.1029/2022JB024360, 2022. a, b

Storkey, D., Blaker, A. T., Mathiot, P., Megann, A., Aksenov, Y., Blockley, E. W., Calvert, D., Graham, T., Hewitt, H. T., Hyder, P., Kuhlbrodt, T., Rae, J. G. L., and Sinha, B.: UK Global Ocean GO6 and GO7: a traceable hierarchy of model resolutions, Geosci. Model Dev., 11, 3187–3213, https://doi.org/10.5194/gmd-11-3187-2018, 2018. a

Tapley, B. D., Watkins, M. M., Flechtner, F., Reigber, C., Bettadpur, S., Rodell, M., Sasgen, I., Famiglietti, J. S., Landerer, F. W., Chambers, D. P., Reager, J. T., Gardner, A. S., Himanshu, S., Ivins, E. R., Swenson, S. C., Boening, C., Dahle, C., Wiese, D. N., Dobslaw, H., Tamisiea, M. E., and Velicogna, I.: Contributions of GRACE to understanding climate change, Nat. Clim. Change, 9, 358–369, https://doi.org/10.1038/s41558-019-0456-2, 2019. a

van Dam, T., Collilieux, X., Wuite, J., Altamimi, Z., and Ray, J.: Nontidal ocean loading: amplitudes and potential effects in GPS height time series, J. Geodesy, 86, 1043–1057, https://doi.org/10.1007/s00190-012-0564-5, 2012. a

van Sebille, E., Biastoch, A., Van Leeuwen, P., and De Ruijter, W.: A weaker Agulhas Current leads to more Agulhas leakage, Geophys. Res. Lett., 36, L03601, https://doi.org/10.1029/2008GL036614, 2009. a

Vinogradova, N. T., Ponte, R, M., and Stammer, D.: Relation between sea level and bottom pressure and the vertical dependence of oceanic variability, Geophys. Res. Lett., 34, L03608, https://doi.org/10.1029/2006GL028588, 2007. a

Vinogradova, N. T., Ponte, R. M., Tamisiea, M. E., Quinn, K. J., Hill, E. M., and Davis, J. L.: Self-attraction and loading effects on ocean mass redistribution at monthly and longer time scales, J. Geophys. Res.-Oceans, 116, C08041, https://doi.org/10.1029/2011JC007037, 2011. a

Voldoire, A.: CNRM-CERFACS CNRM-CM6-1 model output prepared for CMIP6 CMIP, Earth System Grid Federation [data set], https://doi.org/10.22033/ESGF/CMIP6.1375, 2018. a

Voldoire, A., Saint-Martin, D., Sénési, S., Decharme, B., Alias, A., Chevallier, M., Colin, J., Guérémy, J.-F., Michou, M., Moine, M.-P., Nabat, P., Roehrig, R., Salas y Mélia, D., Séférian, R., Valcke, S., Beau, I., Belamari, S., Berthet, S., Cassou, C., Cattiaux, J., Deshayes, J., Douville, H., Ethé, C., Franchistéguy, L., Geoffroy, O., Lévy, C., Madec, G., Meurdesoif, Y., Msadek, R., Ribes, A., Sanchez-Gomez, E., Terray, L., and Waldman, R.: Evaluation of CMIP6 DECK experiments with CNRM-CM6-1, J. Adv. Model. Earth Sy., 11, 2177–2213, https://doi.org/10.1029/2019MS001683, 2019. a

Wang, S., Jing, Z., Wu, L., Sun, S., Chen, Z., Ma, X., and Gan, B.: A more quiescent deep ocean under global warming, Nat. Clim. Change, 14, 961–967, https://doi.org/10.1038/s41558-024-02075-2, 2024. a

Watkins, M. M., Wiese, D. N., Yuan, D.-N., Boening, C., and Landerer, F. W.: Improved methods for observing Earth's time variable mass distribution with GRACE using spherical cap mascons, J. Geophys. Res.-Sol. Ea., 120, 2648–2671, https://doi.org/10.1002/2014JB011547, 2015. a

Weijer, W.: Modal variability in the Southeast Pacific Basin: Energetics of the 2009 event, Deep-Sea Res. Pt. II, 114, 3–11, https://doi.org/10.1016/j.dsr2.2012.10.002, 2015.  a

Weijer, W. and Gille, S. T.: Adjustment of the Southern Ocean to wind forcing on synoptic time scales, J. Phys. Oceanogr., 35, 2076–2089, https://doi.org/10.1175/JPO2801.1, 2005. a, b

Wiese, D. N., Yuan, D.-N., Boening, C., Landerer, F. W., and Watkins, M. M.: JPL GRACE and GRACE-FO Mascon Ocean, Ice, and Hydrology Equivalent Water Height CRI Filtered RL06.3Mv04, Ver. RL06.3Mv04, PO.DAAC [data set], CA, USA, https://doi.org/10.5067/TEMSC-3JC634, 2023. a, b

Zhang, L., Tang, H., Chang, L., and Sun, W.: Performance of GRACE mascon solutions in studying seismic deformations, J. Geophys. Res.-Sol. Ea., 125, e2020JB019510, https://doi.org/10.1029/2020JB019510, 2020. a

Zhao, M. N., Ponte, R. M., Penduff, T., Close, S., Llovel, W., and Molines, J. M.: Imprints of ocean chaotic intrinsic variability on bottom pressure and implications for data and model analyses, Geophys. Res. Lett., 48, e2021GL096341, https://doi.org/10.1029/2021GL096341, 2021. a, b, c, d, e, f

Zhao, M. N., Ponte, R. M., and Penduff, T.: Global-scale random bottom pressure fluctuations from oceanic intrinsic variability, Sci. Adv., 9, eadg0278, https://doi.org/10.1126/sciadv.adg0278, 2023. a

Zuo, H., Balmaseda, M. A., and Mogensen, K.: The new eddy-permitting ORAP5 ocean reanalysis: description, evaluation and uncertainties in climate signals, Clim. Dynam., 49, 791–811, https://doi.org/10.1007/s00382-015-2675-1, 2017. a, b

Download
Short summary
As seawater is moved about by the different types of ocean flow, the pressure at the ocean bottom changes with time and location. We show that such bottom pressure variations are represented reasonably well by high-resolution climate models and that in some regions, like the Arctic Ocean, the intensity of the pressure fluctuations will likely increase under global warming. These insights are useful for the design of future satellite missions that will track mass variations in the Earth system.
Share