Fact-checked by Grok 2 weeks ago

Physical oceanography

Physical oceanography is the discipline within that employs physical laws to analyze the motion, properties, and energy transfers of , encompassing phenomena such as currents, , , and air-sea interactions. It focuses on measurable attributes like , , , and , which govern ocean dynamics and their role in redistributing and across Earth's . Central to the field is the investigation of large-scale circulation patterns, including the driven by density gradients from and variations, which facilitates global transport of and carbon. Key advancements stem from observational via satellites, buoys, and shipboard measurements, enabling models that predict phenomena like El Niño-Southern Oscillation events and their climatic impacts. Controversies arise in interpreting long-term trends, such as attributions, where empirical must be distinguished from model projections influenced by parametric assumptions in coupled atmosphere-ocean systems.

Historical Development

Early Explorations and Observations

Early maritime navigation depended on rudimentary observations of surface currents and winds, with explorers in the 15th and 16th centuries documenting monsoon patterns in the to facilitate trade routes between Europe and Asia. In 1768, , serving as deputy postmaster general for the American colonies, investigated delays in transatlantic mail packets compared to faster whalers and collaborated with his cousin Timothy Folger to produce the first chart of the , incorporating whalers' knowledge and Franklin's temperature measurements to trace its path and warmer boundaries. This chart, distributed to British captains around 1769–1770, highlighted the stream's northward flow from to the , enabling adjustments in sailing courses. Captain James Cook's expeditions from 1768 to 1779 advanced through systematic measurements during his circumnavigations, achieving accuracies of approximately 0.5 feet (15 cm) in height and 0.5 hours in timing at various Pacific and Atlantic sites, which informed early understandings of ranges and lunar influences. Cook's voyages also included hydrographic surveys with line soundings and chronometer-based fixes, revealing coastal and effects on navigation, such as those encountered en route to and . By the mid-19th century, Lieutenant systematized scattered data at the U.S. Naval Observatory's Depot of Charts and Instruments, compiling records from over 1,300 ships between 1842 and 1861 to generate wind and current charts, beginning with the North Atlantic pilot chart in 1847. These charts delineated prevailing surface circulation patterns, including and the Gulf Stream's extensions, reducing average ship transit times across the Atlantic by up to 20–30% through route optimizations. Maury's efforts extended to initial deep-sea soundings, with Sir recording the first modern deep-ocean depth of over 2,500 fathoms in 1840 south of the Islands, laying groundwork for subsurface profiling.

Emergence as a Scientific Discipline

Physical oceanography coalesced as a scientific discipline in the mid-19th century, transitioning from ad hoc naval observations to systematic data compilation and rudimentary physical analysis. , a U.S. officer, pioneered this shift by aggregating thousands of ship logs to produce the first global and current charts in 1847, enabling practical improvements and highlighting patterns like the Gulf Stream's persistence. His 1855 publication, The Physical Geography of the Sea, synthesized these findings into the earliest comprehensive treatise on , emphasizing causes of currents via , differences, and , thus laying foundational principles for treating the as a rather than mere navigational hazard. The HMS Challenger expedition (1872–1876), funded by the British government, marked a pivotal empirical milestone by conducting the first circumnavigating scientific survey, amassing over 492 deep-sea soundings, 362 profiles to abyssal depths, and measurements across the global . These data revealed uniform deep-water properties, a permanent , and abyssal uniformity in and chemistry, challenging prior assumptions of a lifeless, stagnant and providing quantitative evidence for vertical driven by surface and fluxes. Led by Charles Wyville Thomson, the expedition's physical observations—using thermometers, hydrometers, and mechanical current meters—furnished the first large-scale dataset for testing physical theories, though analysis remained largely descriptive until later theoretical integration. By the early 20th century, institutionalization and theoretical advances solidified physical oceanography's disciplinary status, distinct from biological or geological ocean studies. The International Council for the Exploration of the Sea (ICES), established in 1902, coordinated multinational hydrographic surveys in the North Atlantic, standardizing salinity measurements via chlorinity (1902 commission definition as grams of dissolved salts per kilogram of seawater) and fostering data exchange for dynamical inference. Norwegian researchers Bjørn Helland-Hansen and Fridtjof Nansen advanced causal understanding in their 1909 monograph The Norwegian Sea, introducing the dynamic method: leveraging temperature-salinity (T-S) diagrams to compute geostrophic currents from horizontal pressure gradients, assuming hydrostatic and geostrophic balance, which enabled indirect flow estimation without direct velocity measurements. Concurrently, Vagn Walfrid Ekman's 1905 formulation of the Ekman spiral quantified wind-driven surface transport, incorporating Coriolis deflection to explain 90-degree drift from wind direction, bridging empirical data with fluid dynamics equations. Vilhelm Bjerknes' Dynamic Meteorology and Hydrography (1910–1913) further unified the field by applying primitive equations to oceanic circulation, establishing geostrophy as a core principle for large-scale flows. These developments, rooted in verifiable measurements and mathematical rigor, demarcated physical oceanography as a predictive science grounded in conservation laws and balance approximations.

Post-World War II Advances and Computational Era

Following , physical oceanography benefited from wartime technological developments such as improved and bathythermographs, which enabled more precise measurements of propagation and gradients in the , initially driven by antisubmarine warfare needs but adapted for peacetime research. Cold War-era military funding further prioritized deep-sea acoustics and circulation studies, fostering institutional growth at centers like and . A pivotal theoretical advance came in 1948 when Stommel proposed the mechanism for western boundary currents, demonstrating through vortex stretching how Earth's rotation concentrates intense, narrow flows like the along western margins, contrasting with broader eastern returns; this inertial boundary layer theory resolved longstanding inconsistencies in wind-driven circulation models. The International Geophysical Year (IGY) from July 1957 to December 1958 marked a collaborative surge, with 67 nations contributing ship-based expeditions that amassed extensive hydrographic data on temperature, salinity, and currents, revealing basin-scale patterns and advancing the conceptualization of the global overturning circulation. IGY efforts quantified meridional heat transports and validated Sverdrup's interior balance while highlighting the role of boundary dynamics, setting the stage for integrated observational networks. Instrumentation evolved with the 1969 invention of the conductivity-temperature-depth (CTD) profiler by Neil Brown, which provided rapid, in situ measurements of salinity via conductivity alongside temperature and pressure, supplanting slower bottle-based sampling and enabling high-resolution mapping of density structures like the pycnocline. The computational era dawned in the late 1960s with the advent of digital computers, allowing numerical solutions to the governing ocean flow. Kirk Bryan's 1969 model was the first to simulate steady-state global circulation, incorporating realistic continental outlines and bottom topography on a , predicting wind-driven gyres and their intensification at western boundaries consistent with Stommel's theory. That same year, Bryan collaborated with on the inaugural coupled ocean-atmosphere , demonstrating oceanic heat uptake's role in modulating equatorial climates and foreshadowing studies. By the , these foundations supported primitive equation models resolving baroclinic modes and eddy variability, transitioning physical oceanography from empirical descriptions to predictive simulations of thermohaline and wind-driven regimes.

Fundamental Physical Properties

Ocean Basin Geometry and Setting

The ocean basins form the dominant topographic features of 's surface, covering 71% of the planet and holding 1.37 billion cubic kilometers of , which constitutes 97% of all water on . These basins exhibit profound variations in depth, averaging 3,682 meters globally, with the deepest regions exceeding 11,000 meters due to tectonic zones. The geometry reflects ongoing , including divergent ridges where new crust forms and convergent trenches where it is recycled, alongside passive continental margins. Major ocean basins include the Pacific, the largest at approximately 165 million square kilometers or 46% of the oceanic surface, featuring an average depth of 4,000 meters and the Mariana Trench's at 10,994 meters; the , an S-shaped basin bisected by the ; the , bounded by , , and ; the , the shallowest and smallest; and the encircling . The Pacific's expansive basin hosts the global system's most active segments and encircles subduction zones forming the . In contrast, the Atlantic's ridge rises prominently, dividing the basin into eastern and western segments with abyssal plains accumulating sediments over basaltic crust. Continental margins transition from shelves at depths less than 200 , averaging 80 kilometers in width, to steep slopes descending 2-5 kilometers, followed by gentler rises leading to abyssal plains at 4,500-6,000 depth, which comprise much of the floors and represent the flattest expanses on . Mid-ocean ridges, totaling over 64,000 kilometers in length, elevate to average depths of 2,500 , facilitating at rates of 1-10 centimeters per year. Submarine trenches, primarily in the Pacific, plunge to depths greater than 6,000 , hosting the planet's deepest points and influencing asymmetry through crustal recycling. These features dictate water mass distribution, circulation patterns, and sediment deposition, underpinning physical oceanographic processes.

Temperature Distributions and Profiles

The vertical temperature profile in the typically consists of a warm, well-mixed surface layer overlying a —a zone of rapid decrease—followed by a deep layer of nearly uniform cold water. In the surface , which extends from the surface to depths of about 50–200 meters depending on wind forcing and seasonal heating, are relatively uniform due to turbulent mixing. Global surface average approximately 17°C, ranging from near-freezing values of -2°C in polar regions to maxima exceeding 30°C in equatorial zones or enclosed during summer. The begins below the , often at depths of 100–400 meters in subtropical regions, where gradients sharpen to drops of 15–20°C over several hundred meters, transitioning from surface values around 20°C to intermediate depths near 5°C. This layer's depth and intensity vary geographically: shallower and stronger in the (200–300 meters) due to persistent , and deeper or absent in polar areas where cold surface waters extend downward. Below the , in the deep ocean (typically >1,000 meters), stabilize at 1–4°C, with a global volume-weighted average of about 3.5°C, maintained by slow vertical mixing and the downward propagation of cold and . Horizontally, distributions follow latitudinal patterns driven by insolation, with annual mean surface peaking at 25–28°C near the and declining poleward to below 5°C at high latitudes, modulated by ocean currents that transport heat equatorward in western boundary flows like the . Seasonal variations in profiles are pronounced in mid-latitudes (30–60°), where summer heating warms and deepens the to 100 meters or more, eroding the temporarily, while winter cooling and shoal it; equatorial and polar regions exhibit minimal seasonal change due to consistent insolation or cover. These profiles influence stratification, as dominates in the upper ocean, with colder, denser deep waters reflecting long-term isolation from surface heat fluxes.

Salinity Variations and Controls

, defined as the mass fraction of dissolved in , is conventionally measured in practical salinity units (PSU), where 1 PSU approximates 1 gram of per of . The global average sea surface is approximately 35 PSU, though it ranges from about 32 PSU in regions of high freshwater input to 37 PSU in areas of net evaporation. Spatial variations in sea surface primarily reflect the imbalance between and . Subtropical gyre centers exhibit maxima exceeding 36.5 PSU due to dominant exceeding by up to 1 meter per year, concentrating salts in surface waters. In contrast, equatorial zones and high-latitude regions show minima around 32-34 PSU, driven by excess (often 2-3 meters annually) and river discharge diluting surface layers. River outflows, such as from the (adding ~0.3 million cubic meters per second of freshwater) and , create localized low- plumes extending hundreds of kilometers offshore, while enclosed basins like the Mediterranean reach salinities over 38 PSU from minimal freshwater input and high rates. The principal controls on salinity operate through surface freshwater fluxes and phase changes of . removes pure , elevating salinity by 0.1-0.5 PSU per month in arid subtropical zones, whereas and river runoff introduce freshwater, reducing salinity by comparable amounts in humid . formation in polar regions rejects salts into the underlying water, increasing near-surface salinity by up to 2 PSU during winter brine rejection, while summer melting dilutes it; this process contributes to the formation of dense with salinities near 34.6 PSU. Subsurface and mixing redistribute these surface signals, but vertical profiles typically show fresher surface layers in low latitudes ( deepening to 100-200 meters) and saltier deep waters in high latitudes due to brine exclusion. Temporal variations occur on diurnal to decadal scales, modulated by seasonal cycles of evaporation-precipitation and . Seasonal swings of 0.5-2 PSU are observed in the upper 350 meters globally, with maxima in dry seasons (e.g., northern summer in ) and minima during wet periods, as confirmed by float and satellite data from 2005-2020. Interannual changes, linked to modes like El Niño (which freshens the eastern Pacific by 0.2-0.5 PSU via enhanced rainfall), amplify these patterns, while long-term trends show amplification of extremes—fresher fresh waters and saltier salty waters—consistent with intensified hydrological cycles under warming climates.

Density Stratification and Pycnocline Dynamics

Ocean results from spatial variations in , , and , governed by the nonlinear for , ρ = ρ(S, T, p), where colder temperatures, higher salinities, and greater pressures increase . This , empirically derived and valid for temperatures from -2°C to 35°C, salinities of 2 to 42, and pressures up to 10,000 dbar, captures thermosteric effects from and halosteric effects from , with compressibility adding dependence. manifests as lighter water overlying denser water, stabilizing the against vertical displacement. The pycnocline denotes the zone of rapid increase with depth, typically spanning 100–1,000 meters globally, acting as a barrier to turbulent mixing by enhancing gravitational . In low latitudes, a permanent pycnocline persists year-round at depths around 100–200 meters, driven by consistent heating and that maintain surface lightness. Higher latitudes feature a seasonal pycnocline overlying the permanent one, forming in summer via surface warming and precipitation-induced freshening, with thicknesses averaging 50–150 meters in subtropical regions. Pycnocline dynamics respond to surface forcing: winter convection from cooling and storms deepens the , eroding the seasonal pycnocline and homogenizing to depths exceeding 200 meters in subpolar zones; spring restratification rebuilds gradients through flux. Wind-driven mixing and activity modulate pycnocline depth, with mesoscale eddies displacing it by tens of meters, particularly in the upper 200–300 meters where cores align. Observations indicate summertime upper-ocean pycnocline has intensified since 1970 at rates of 10^{-6} to 10^{-5} s^{-2} per decade across basins, linked to surface warming that amplifies contrasts. Regionally, pycnocline depth shoals toward the , averaging shallower in the (∼150 ) and deepening poleward, with global upper-ocean pycnocline thickness remaining relatively uniform at ∼100 despite depth variations from 50 to 300 . In the , sea ice-ocean interactions contribute to internal pycnocline formation by exporting freshwater, enhancing stratification between 200–1,500 . These dynamics limit vertical exchanges, confining most mixing to the surface layer and influencing trapping below, though submesoscale processes and internal waves enable localized across the .

Governing Physical Principles

Fluid Mechanics and Momentum Balances

The Navier-Stokes equations govern the momentum balance in oceanic flows, expressing Newton's second law for a continuum of incompressible, viscous fluid parcels. In their general form, they balance the rate of change of momentum—comprising local temporal acceleration and advective terms—with pressure gradient forces, viscous diffusion, and body forces such as gravity. For seawater, modeled as a Newtonian fluid with density ρ ≈ 1025 kg/m³ and dynamic viscosity μ ≈ 10^{-3} Pa·s at typical temperatures, the equations are ∂u/∂t + (u·∇)u = - (1/ρ) ∇p + ν ∇²u + g, where u is the velocity vector, ν = μ/ρ is kinematic viscosity (≈10^{-6} m²/s), p is pressure, and g is gravitational acceleration (≈9.8 m/s²). These derive from integrating stresses over infinitesimal fluid volumes, assuming no-slip boundaries at solid surfaces and continuity of velocity across fluid interfaces. Oceanic applications invoke the Boussinesq approximation, treating as constant (ρ₀) in all terms except buoyancy forcing via g' = g (ρ - ρ₀)/ρ₀, where density anomalies δρ/ρ₀ < 0.5% arise primarily from temperature (α ≈ 2×10^{-4} °C^{-1}) and salinity (β ≈ 7.5×10^{-4} psu^{-1}) variations. This filter eliminates sound waves irrelevant to slow oceanic circulations (timescales > hours) while preserving internal gravity waves and stratification effects, reducing computational demands in models spanning global basins. The approximation holds because oceanic numbers are <<1 and aspect ratios (depth/scale) <<1, with error bounds <1% for typical flows. Vertical momentum reduces to hydrostatic balance, ∂p/∂z = -ρ g, as vertical accelerations (<<10^{-3} m/s²) pale against gravity, justified by Rossby numbers Ro = U/fL ≈ 0.1-1 (U horizontal velocity ~0.1-1 m/s, f Coriolis ~10^{-4} s^{-1}, L ~10^2-10^4 km) and small vertical velocities W ~ (H/L) U ~10^{-3}-10^{-2} m/s (H ~km). This decouples vertical structure, enabling pressure computation from integrated density via the equation of state σ_T(T,S,p) ≈ ρ(T,S) - 1000 kg/m³, where T is temperature and S salinity. Horizontal equations retain nonlinear self-advection (u·∇)u, dominant in mesoscale eddies (scales 10-100 km, speeds 0.1-0.5 m/s) where local Rossby deformation radii L_R = NH/f ≈ 10-50 km (N buoyancy frequency ~10^{-2} s^{-1}). Viscous terms ν ∇²u, negligible molecularly (Ekman numbers Ek = ν/f H² <<10^{-10}), are upscaled via Reynolds averaging to eddy viscosities A_h ~10-100 m²/s horizontally and A_v ~10^{-4}-10^{-5} m²/s vertically, parameterizing turbulence from unresolved scales. These sustain balances in boundary layers, such as Ekman depths δ_E = √(2 A_v / f) ≈ 10-100 m, where friction counters wind stress τ ~0.01-0.1 N/m². In primitive equation models standard since the 1980s, these form the core alongside mass conservation ∇·u = 0 (incompressibility) and tracer equations for T and S, enabling simulations of currents like the (speeds >1 m/s, widths ~100 km). Full Navier-Stokes resolution remains infeasible for basins due to grid requirements <1 m amid Reynolds numbers Re ~10^8-10^{10}, necessitating subgrid closures informed by large-eddy simulations or observations.

Coriolis Effect and Geostrophic Balance

The Coriolis effect manifests as an apparent deflection of fluid motion in a rotating reference frame, arising from the conservation of angular momentum in the Navier-Stokes equations adapted for Earth's rotation. This fictitious force is expressed as \mathbf{F}_c = -2 \boldsymbol{\Omega} \times \mathbf{u}, where \boldsymbol{\Omega} is Earth's angular velocity vector with magnitude \Omega = 7.292 \times 10^{-5} rad/s directed along the rotation axis, and \mathbf{u} is the fluid velocity. For horizontal flows in oceanography, the vertical component dominates, yielding the Coriolis parameter f = 2 \Omega \sin \phi, where \phi denotes latitude; f > 0 in the , causing deflection to the right of motion, and f < 0 in the Southern Hemisphere, causing leftward deflection. This parameter varies from near zero at the equator (\phi = 0^\circ) to approximately $1.46 \times 10^{-4} s^{-1} at the poles, rendering the effect negligible within about 2° of the equator where |\sin \phi| \approx 0. Geostrophic balance describes the dominant steady-state equilibrium for large-scale oceanic flows, where the Coriolis force precisely counters the horizontal pressure gradient force, neglecting acceleration and friction. The governing equations in a Cartesian coordinate system (x eastward, y northward) are -f v_g = -\frac{1}{\rho} \frac{\partial p}{\partial x} and f u_g = -\frac{1}{\rho} \frac{\partial p}{\partial y}, yielding geostrophic velocities u_g = -\frac{1}{\rho f} \frac{\partial p}{\partial y} and v_g = \frac{1}{\rho f} \frac{\partial p}{\partial x} for f > 0. This balance implies currents flow parallel to isobars (constant pressure surfaces), with higher pressure to the right in the , and is valid when the Ro = \frac{U}{f L} \ll 1, where U is a characteristic velocity and L a horizontal length scale—typically holding for gyres and mid-ocean currents with L > 100 km and U < 1 m/s. In practice, oceanic pressure gradients often stem from sea surface height variations \eta, approximated via hydrostatic balance as \frac{\partial p}{\partial x} \approx \rho g \frac{\partial \eta}{\partial x} near the surface, allowing geostrophic currents to be inferred from satellite altimetry measurements of \eta. The f-plane approximation assumes constant f over the domain, simplifying models for mid-latitude studies but introducing errors near the equator or for meridional flows spanning significant latitudes; the β-plane extension incorporates \beta = \frac{\partial f}{\partial y} \approx \frac{2 \Omega \cos \phi}{a} (with Earth radius a \approx 6371 km) to capture latitudinal variation effects like planetary vorticity gradients. Deviations from geostrophy occur in boundary layers, eddies, or high-Ro regimes (e.g., tropical currents), where ageostrophic components introduce acceleration or frictional terms, but global assessments confirm geostrophic dominance in approximately 80-90% of extratropical open-ocean momentum balances based on reanalysis data.

Friction and Turbulent Processes

In oceanic flows, friction predominantly occurs through turbulent processes rather than molecular viscosity, owing to Reynolds numbers exceeding 10^6, which render laminar regimes negligible. Turbulent friction dissipates kinetic energy across scales from millimeters to hundreds of kilometers, influencing momentum transfer, mixing, and large-scale circulation balances. Bottom friction, concentrated in the benthic boundary layer, parameterizes as quadratic drag \tau_b = \rho C_d |u_b| u_b, where C_d typically ranges from 2.0 \times 10^{-3}) to 3.0 \times 10^{-3}) for rough seafloors, derived from empirical fits to velocity profiles over continental shelves and abyssal plains. This friction extracts energy from geostrophic currents and tides, acting as a vorticity sink that weakens gyre circulations, as evidenced in models of the where enhanced bottom drag reduces cyclonic vorticity by up to 20%. Over rough topography, such as mid-ocean ridges, friction amplifies dissipation, with drag coefficients increasing by factors of 2-5 compared to smooth basins. At the surface, wind-induced shear and wave breaking inject turbulent kinetic energy into the ocean surface boundary layer (OSBL), typically 10-50 m thick under moderate winds of 10 m/s. Wave breaking generates plumes with dissipation rates \epsilon peaking at 10^{-6} to 10^{-4} , \mathrm{W/kg}) near the surface, decaying exponentially with depth over a scale of 10-20 m, as measured by microstructure profilers in fetch-limited conditions. This turbulence entrains momentum downward, sustaining , while rain or convection can enhance mixing rates by 10-100 times in localized patches. Turbulent processes drive diapycnal mixing, quantified by eddy diffusivities K_\rho of 0.1-1 \times 10^{-4} , \mathrm{m^2/s}) in the stratified interior, rising to 10^{-3} , \mathrm{m^2/s}) near rough topography or eddy edges due to internal wave breaking. Global estimates from Argo floats and gliders indicate enhanced K_\rho over seamounts and ridges, contributing 20-50% of interior dissipation, essential for upwelling in thermohaline circulation. In ocean models, these are parameterized via schemes linking \epsilon to buoyancy frequency N, with Richardson number criteria (Ri < 0.25) triggering shear instability. Observational campaigns, such as DIMES in the Southern Ocean, confirm topographic enhancement of mixing by factors of 5-10, underscoring friction's role in meridional overturning.

Large-Scale Circulation

Wind-Driven Surface Currents and Gyres

Wind-driven surface currents arise from the tangential stress exerted by atmospheric winds on the sea surface, transferring momentum to the upper ocean layer through molecular and turbulent friction, with typical penetration depths of 50-100 meters in the absence of strong mixing. These currents constitute approximately 10% of the total ocean volume, as they are largely confined to the upper 100-200 meters where direct wind influence dominates over density-driven flows. The magnitude of wind stress, \tau = \rho_a C_d |U_{10}|^2, depends on air density \rho_a, drag coefficient C_d (typically 1-2 \times 10^{-3}), and wind speed at 10 meters U_{10}, resulting in surface velocities that are roughly 3% of wind speed under steady conditions. In the ocean interior, away from boundaries, the large-scale circulation integrates this wind forcing via the Sverdrup balance, which equates the planetary vorticity input from Ekman pumping to the wind stress curl: \beta V = \frac{\curl \tau}{\rho_0}, where \beta is the meridional gradient of the Coriolis parameter, V is the meridional transport, \tau is wind stress, and \rho_0 is reference density. Positive wind stress curl in subtropical regions (driven by equatorward trade winds transitioning to poleward westerlies) induces downward Ekman pumping, fostering clockwise-rotating gyres in the Northern Hemisphere and counterclockwise in the Southern, with total meridional transports scaling to 20-50 Sverdrups (Sv; 1 Sv = $10^6 m³/s) across major basins. This relation accurately predicts the volume transport and rotational sense of observed gyres, validating the quasi-geostrophic approximation for wind-forced, frictionally arrested flows. The five primary subtropical gyres—one each in the North and South Atlantic, North and South Pacific, and Indian Oceans—span thousands of kilometers and are bounded by zonal equatorial currents, eastern cool return flows, western warm boundary currents, and northern/southern transverse currents. For instance, the North Pacific Gyre encompasses the North Equatorial Current (flowing westward at ~0.5-1 m/s), Kuroshio Extension (with transports up to 60-70 Sv), and California Current, enclosing a vast subtropical region of convergent flow that influences nutrient upwelling and biogeochemical cycles. Similarly, the North Atlantic Gyre integrates the Gulf Stream system, achieving peak speeds exceeding 2.5 m/s in its western intensification phase, though interior speeds remain subdued at 0.1-0.3 m/s due to geostrophic balance. Subpolar gyres, such as those in the northern and , contrast by rotating oppositely under negative wind stress curl from prevailing westerlies and polar easterlies, promoting divergence and upwelling with transports on the order of 10-20 Sv. These systems interface with subtropical gyres at frontal zones like the subpolar-subtropical front, where baroclinic instabilities generate mesoscale eddies that redistribute momentum and heat. Empirical observations from satellite altimetry and Argo floats confirm that gyre-scale transports align closely with Sverdrup predictions, with discrepancies attributable to eddy variability or remote boundary influences rather than fundamental flaws in the wind-driven paradigm. Seasonal wind variations modulate gyre intensities, with stronger trades amplifying equatorial divergence and westerlies enhancing mid-latitude convergence, as quantified in reanalysis datasets like those from the ECMWF.

Ekman Transport and Spiral

The Ekman spiral characterizes the ageostrophic component of wind-forced ocean currents in the surface mixed layer, arising from the interaction of the Coriolis force with vertical turbulent momentum fluxes. Swedish oceanographer Vagn Walfrid Ekman formulated the theory in 1902 to explain observations of surface drift velocities deflected approximately 20-45 degrees to the right of the wind in the Northern Hemisphere, as noted during Arctic expeditions. The model assumes steady, horizontally uniform, barotropic flow under constant wind stress, with momentum balanced solely by Coriolis deflection and linear frictional drag parameterized by a constant eddy viscosity K. Horizontal pressure gradients are absent in the simplest formulation, isolating frictional effects. In the Northern Hemisphere, the steady-state momentum equations reduce to f v = \frac{\partial}{\partial z} (K \frac{\partial u}{\partial z}) and -f u = \frac{\partial}{\partial z} (K \frac{\partial v}{\partial z}), where u and v are eastward and northward velocities, f = 2 \Omega \sin \phi is the Coriolis parameter (\Omega Earth's rotation rate, \phi latitude), and z increases upward from a no-slip bottom boundary. Solving with a surface wind stress \tau_x, \tau_y yields a helical velocity profile: surface flow at 45 degrees to the wind, rotating clockwise with depth and decaying exponentially as e^{z/D}, where the Ekman depth D = \sqrt{2K/f} marks the scale over which velocities drop to $1/e of surface values. Typical oceanic values yield D \approx 50-100 meters, depending on turbulence levels in the mixed layer. Ekman transport refers to the vertically integrated mass flux through this layer, M_x = \int_{-\infty}^0 v \, dz = -\tau_y / (\rho f) and M_y = \int_{-\infty}^0 u \, dz = \tau_x / (\rho f), directed 90 degrees to the right of the wind stress vector (\tau_x, \tau_y) in the Northern Hemisphere (left in the Southern). This net advection, with magnitude \tau / (\rho f) independent of K or spiral details, arises from the orthogonal alignment of successive layers' deflections./07:Basis_of_Wind-Driven_Circulation-_Ekman_spiral_and_transports) The transport scales inversely with latitude via f, vanishing at the equator, and drives divergence or convergence that induces Ekman pumping velocities w_E = \nabla \cdot \mathbf{M} / \rho of order 10-100 meters per year, linking surface winds to interior geostrophic flow. Observational validation comes from current meter arrays and drifters, which confirm the 90-degree transport direction but often show muted spirals due to stratification, variable eddy viscosity, and wave-induced turbulence altering frictional closure. For instance, analyses of mid-latitude data reveal effective D modulated by seasonal mixing, with spirals more evident under steady trades than in stormy conditions. The theory underpins Sverdrup's interior ocean balance but overpredicts spiral rigidity in stratified settings, where buoyancy suppresses vertical momentum transfer.

Thermohaline Circulation and Deep Overturning

The thermohaline circulation comprises the density-driven component of global ocean circulation, wherein differences in water density arising from variations in temperature and salinity induce vertical motion and large-scale meridional flows. This process contrasts with wind-driven surface currents by operating predominantly on centennial timescales and penetrating to abyssal depths. Dense water masses form through surface cooling and brine rejection during sea ice production, leading to convective overturning that exports water from the surface layer into the deep ocean. Deep water formation occurs in select high-latitude regions where extreme cooling and salinity increases elevate density sufficiently for sinking to abyssal levels. Primary sites include the Labrador Sea and Greenland-Norwegian Seas for North Atlantic Deep Water (NADW), the Weddell and Ross Seas for Antarctic Bottom Water (AABW), and overflow contributions from the Mediterranean Sea via dense outflows over sills. In the North Atlantic, winter convection can reach depths exceeding 2,000 meters in the Labrador Sea, with annual formation rates estimated at 10-20 Sverdrups (Sv; 1 Sv = 10^6 m³/s) of NADW. AABW production in the Southern Ocean involves open-ocean polynyas and shelf processes, contributing denser waters that fill the global deep basins. The resulting deep overturning manifests as a meridional circulation cell, particularly prominent in the Atlantic as the (AMOC), where northward surface flow of warm, saline water is balanced by southward deep return of NADW at approximately 15-18 Sv at 26°N latitude. This northward heat transport, equivalent to about 1 petawatt, moderates European climates and ventilates the deep ocean with oxygen and nutrients. In the global context, the "conveyor belt" links basins: NADW flows southward, mixes with , upwells primarily in the Southern Ocean via wind-driven Ekman suction and diapycnal mixing, then returns northward in the Pacific and Indian Oceans as less dense intermediate waters. Observational evidence from hydrographic sections, neutrally buoyant floats, and geochemical tracers confirms this interbasin exchange, with deep Pacific waters exhibiting ages of 500-1000 years based on radiocarbon deficits. Variability in thermohaline circulation arises from surface flux anomalies, such as freshwater input from Arctic ice melt or precipitation, which can weaken deep convection by stabilizing the water column. Instrumental records from the RAPID array since 2004 indicate AMOC fluctuations of 2-3 Sv interannually, linked to wind stress and buoyancy forcing, though long-term trends remain debated due to sparse pre-1990s deep observations. Peer-reviewed syntheses emphasize that while models project potential AMOC slowdown under high-emission scenarios from increased Greenland meltwater, paleoclimate proxies like sediment cores reveal past collapses tied to abrupt freshwater pulses, underscoring sensitivity to northern density gradients. Sustained monitoring via arrays like OSNAP and deep Argo floats is essential to discern anthropogenic signals amid natural decadal variability.

Boundary Currents and Western Intensification

Boundary currents are narrow, swift flows that hug the western and eastern edges of major ocean basins, primarily driven by wind patterns in subtropical gyres. Western boundary currents, such as the in the North Atlantic and the in the North Pacific, are characteristically intense, transporting large volumes of water poleward while eastern boundary currents, like the and , are broader and weaker. The , for instance, attains surface speeds of up to 2.5 meters per second, spans approximately 100 kilometers in width, and extends to depths exceeding 1,000 meters, carrying over 100 million cubic meters of water per second northward. Similarly, the reaches velocities of 0.5 to 3 meters per second near the surface and flows at depths around 400 meters, forming the western limb of the North Pacific Gyre. Western intensification refers to the observed asymmetry in gyre circulation, wherein western boundary currents are narrower, faster, and deeper than eastern ones to achieve equivalent meridional transport volumes required by in the basin interior. This phenomenon arises because the planetary vorticity gradient, or (the latitudinal variation of the f = 2 \Omega \sin \phi, where \beta = \partial f / \partial y \approx 2 \times 10^{-11} m^{-1} s^{-1} at mid-latitudes), deflects interior flows equatorward on the eastern side, necessitating a concentrated, friction-dominated return flow along the western boundary to conserve potential vorticity and close the gyre. Eastern boundaries, conversely, feature sluggish, wide flows where topographic constraints and weaker relative vorticity gradients allow broader dissipation without intensification. Henry Stommel's seminal 1948 model provided the foundational explanation, modeling a homogeneous rectangular ocean under steady wind stress with linear bottom friction and no lateral viscosity. In this framework, the governs the interior vorticity balance (\beta v = \nabla \times \tau / \rho H, where v is meridional velocity, \tau wind stress, \rho density, and H depth), but boundary layers emerge to rectify the mismatch: a thin eastern layer balances relative vorticity, while a thicker western layer incorporates planetary vorticity changes and friction to enable northward transport against southward interior advection. Stommel's solution predicts an exponentially decaying western boundary current scale \delta \approx \sqrt{2 r / \beta U}, where r is friction coefficient and U interior speed, yielding widths of tens to hundreds of kilometers consistent with observations. Subsequent refinements, such as Walter Munk's 1950 inclusion of lateral eddy viscosity, reinforced the role of nonlinear instabilities in sustaining meanders and rings, but the core remains dominant. Empirical data from moored arrays and satellite altimetry confirm western intensification's robustness across basins, with the exhibiting transport variability of 30-150 Sverdrups (1 Sv = 10^6 m^3/s) and the showing comparable poleward heat fluxes exceeding 10^15 W, influencing regional climates and meridional overturning. Deviations occur in high-latitude or buoyancy-driven regimes, but wind-forced subtropical gyres universally display this pattern, underscoring the primacy of Earth's rotation in shaping large-scale oceanic momentum balances.

Air-Sea Interactions

Momentum Flux and Ocean-Atmosphere Coupling

The momentum flux across the air-sea interface refers to the transfer of horizontal momentum from the atmosphere to the ocean, primarily through surface wind stress, which drives upper-ocean currents and contributes to large-scale circulation patterns. This flux is directed downward under typical conditions, with global mean magnitudes on the order of 0.05 to 0.1 N/m² corresponding to average wind speeds of 6-8 m/s. The stress arises from shear in the atmospheric boundary layer interacting with ocean surface waves and currents, quantified via the bulk aerodynamic parameterization \vec{\tau} = \rho_a C_d |\vec{U}_{10} - \vec{u}_s| (\vec{U}_{10} - \vec{u}_s), where \rho_a \approx 1.2 kg/m³ is air density, C_d is the dimensionless drag coefficient, \vec{U}_{10} is wind velocity at 10 m height, and \vec{u}_s is surface current velocity. When surface currents are small relative to winds, this approximates \vec{\tau} \approx \rho_a C_d U_{10}^2. The drag coefficient C_d encapsulates effects of surface roughness, stability, and sea state, exhibiting a global mean of approximately $1.25 \times 10^{-3} but varying systematically with wind speed: it increases from near $1.0 \times 10^{-3} at low speeds to saturation around $2.5 \times 10^{-3} in high winds exceeding 20 m/s. Empirical measurements from field campaigns, such as those using direct covariance flux techniques with sonic anemometers during experiments like and , confirm this wind-speed dependence, with C_d formulations like incorporating neutral-stability corrections (C_d = (0.017 U_{10N} - 0.005) \times 10^{-3}, where U_{10N} is neutral 10-m wind). Wave influences further modulate C_d; short wind-driven waves enhance roughness and stress by up to 25%, while opposing swells can reduce it by 15% in mixed seas, as validated against buoy and aircraft data. Ocean-atmosphere coupling manifests in the bidirectional nature of momentum exchange: while atmospheric winds impose stress on the ocean, surface currents feedback by altering relative wind speeds, typically reducing stress magnitude by 3-10% in regions of strong currents like the Gulf Stream. This relative-wind effect generates stress curls from crosswind current shears, amplifying atmospheric wind responses over mesoscale ocean features such as eddies, thereby enhancing energy transfer to the ocean by up to 30% locally. In coupled models, neglecting current-relative formulations leads to biases in simulated surface currents and circulation strength, underscoring the causal role of oceanic motion in modulating atmospheric forcing. Wave-current interactions further couple the systems, with parameterized wave-based adjustments improving flux estimates in mixed sea states and reducing errors in ocean heat content trends.

Heat Flux Mechanisms and Storage

The air-sea heat flux represents the exchange of thermal energy between the ocean surface and the overlying atmosphere, comprising radiative and non-radiative components that govern the ocean's role in Earth's energy balance. Radiative fluxes include incoming shortwave solar radiation, which penetrates the upper ocean layers, and net longwave radiation, where the ocean emits infrared radiation modulated by atmospheric back-radiation. Non-radiative or turbulent fluxes consist of sensible heat transfer driven by air-sea temperature gradients via conduction and convection, and latent heat transfer associated with evaporation and condensation processes. Shortwave radiation typically dominates incoming energy, with global annual averages of approximately 160-170 W/m² absorbed after accounting for albedo (ocean reflectivity around 0.06-0.10), varying by latitude and cloud cover. Net longwave loss from the ocean surface averages 50-60 W/m², as surface emission (governed by , σT⁴ where σ=5.67×10⁻⁸ W/m²K⁴ and T≈288 K) exceeds downward atmospheric radiation. Sensible heat flux, smaller in magnitude at 10-20 W/m² globally, follows bulk aerodynamic formulas depending on wind speed and temperature differences (ΔT), while latent heat flux, often 80-100 W/m², is tied to evaporation rates parameterized by wind-driven moisture gradients and the latent heat of vaporization (2.5×10⁶ J/kg). These components are quantified through in-situ measurements (e.g., buoys) and satellite-derived products, with uncertainties reduced by closure experiments balancing fluxes against observed ocean heat content changes. In the present climate, the net surface heat flux into the ocean averages about 0.5-1 W/m² globally, reflecting an imbalance where radiative gains exceed turbulent and longwave losses, primarily due to anthropogenic greenhouse gas forcing that enhances atmospheric trapping of outgoing radiation. This net uptake drives oceanic warming, with regional variations: subtropical gyres exhibit net heat loss via enhanced evaporation, while equatorial and polar regions show net gain or loss tied to upwelling and ice dynamics, respectively. Air-sea fluxes couple with ocean circulation to redistribute heat meridionally, but local flux anomalies (e.g., from wind shifts) can trigger events like marine heatwaves through surface trapping. The ocean's heat storage capacity stems from water's high specific heat capacity (4186 J/kg·K, over four times that of air), enabling it to absorb and retain approximately 90% of Earth's excess anthropogenic heat since the mid-20th century without proportional temperature rises. Full-depth ocean heat content reached a record 452 ± 77 zettajoules (10²¹ J) by 2024 relative to 1960 levels, with the upper 2000 m accounting for most of the increase (about 90% of total OHC anomaly). Heat is stored primarily in the mixed layer (0-100 m) for seasonal cycles via vertical mixing and entrainment, while subduction and diffusion convey it to the thermocline and deep ocean over decadal to millennial timescales, modulated by circulation strength. Observations from Argo floats and ship-based profiles confirm accelerating uptake rates, with 2024 OHC 15 ± 9 ZJ higher than 2023, underscoring the ocean's buffering role against atmospheric warming.

Freshwater Flux and Its Implications

Freshwater flux refers to the net transfer of water across the ocean-atmosphere interface and continental boundaries, primarily comprising (P - E) over the ocean surface plus riverine and ice sheet runoff (R). This flux acts as a boundary condition for ocean salinity, influencing density gradients through dilution or concentration effects. Globally, the ocean dominates the , accounting for 86% of evaporation and 78% of precipitation, with net fluxes balancing near zero over long timescales but varying regionally. Spatial patterns exhibit strong latitudinal dependence: subtropical gyres experience net evaporation (E > P), exporting freshwater and increasing surface , while equatorial convergence zones and high latitudes receive net or runoff, importing freshwater. River discharge adds approximately 40,181 km³ annually between 90°N and 60°S, concentrated in coastal margins like the and , further freshening shelf seas. Recent estimates from 2020–2025 indicate amplified flux variability due to enhanced hydrological cycling, with inputs rising from melt and thaw, though global net oceanic influx remains small at around 0.1–0.5 × 10³ km³/year after balancing land-ocean exchanges. Excess freshwater input reduces surface , enhancing vertical by lightening upper layers relative to deeper saline waters, which suppresses convective mixing in polar regions. This inhibits the formation of dense water masses essential for deep overturning, potentially destabilizing ; model simulations show that anomalous Arctic freshwater pulses weaken the Atlantic Meridional Overturning Circulation (AMOC) by 10–20% through reduced production. In the , and runoff dominate , maintaining a that isolates cold deep waters. Freshwater fluxes also contribute to sea level variability: direct addition from P and R raises levels by 1–2 mm/year regionally, while salinity changes induce halosteric effects, with freshening expanding low-salinity volumes and altering steric height. Enhanced fluxes under warming amplify ocean heat uptake by deepening the in some basins but risk tipping points in circulation, as evidenced by paleoclimate proxies linking meltwater events to AMOC slowdowns lasting millennia. These underscore freshwater flux as a key control on ocean stability, with ongoing freshening projected to intensify subpolar gyre shifts by mid-century.

Oceanic Variability

Planetary Waves and Rossby Waves

Planetary waves in the ocean are large-scale, low-frequency oscillations influenced by the and the variation of the Coriolis parameter with , known as the beta effect. These waves, which include Rossby waves and Kelvin waves, facilitate the adjustment of ocean circulation to changes in wind forcing over timescales of months to years. Unlike smaller-scale gravity waves, planetary waves propagate energy westward in the extratropics due to the restoring force provided by planetary gradients, enabling basin-wide redistribution of and heat. Rossby waves, a primary subclass of planetary waves, manifest as westward-propagating undulations in and fields, spanning hundreds to thousands of kilometers horizontally. They arise from the conservation of in a rotating fluid, where perturbations in relative vorticity are balanced by displacements that alter planetary vorticity. The theoretical foundation derives from the quasi-geostrophic approximation of the , yielding a for barotropic Rossby waves of \omega = -\frac{\beta k}{K^2}, where \omega is , \beta is the meridional of the Coriolis parameter, k and l are zonal and meridional s, and K^2 = k^2 + l^2. This results in westward phase speeds c_x = \omega / k = -\beta / K^2, which decrease with increasing wavenumber, making longer waves propagate faster and allowing energy to disperse over time. In the ocean, Rossby waves exhibit periods ranging from weeks to decades, with phase speeds typically on the order of 2-10 cm/s in mid-latitudes, slowing near the where the beta effect diminishes. Baroclinic modes, involving vertical structure via internal deformation radii, modify the dispersion to \omega \approx -\frac{\beta k}{K^2 + 1/L_d^2}, where L_d is the deformation radius, enabling deeper penetration and multiple vertical modes. Observations from satellite altimetry confirm these properties, revealing coherent westward-propagating signals in sea surface height anomalies across ocean basins, consistent with linear theory for long waves. Rossby waves play a pivotal role in oceanic variability by transporting information from wind forcing regions, such as the eastern boundaries, westward across basins, thereby modulating gyre-scale circulations and . They contribute to interannual fluctuations, including the adjustment phase of El Niño-Southern Oscillation events, where equatorial reflections influence remote anomalies. In extratropical regions, they drive variability in transport and eddy interactions, with wind-driven generation linking atmospheric patterns to ocean responses on decadal scales. Empirical studies indicate that dynamics explain a significant portion of low-frequency trends and circulation shifts, underscoring their causal importance in climate-ocean coupling.

Mesoscale Eddies and Submesoscale Processes

Mesoscale eddies constitute a dominant feature of variability, characterized by coherent, rotating parcels with scales of 10 to 100 kilometers and temporal durations spanning weeks to months. These structures emerge predominantly from in geostrophically balanced flows, including baroclinic instability driven by slanted isopycnals in the presence of vertical and barotropic instability arising from in fields. Eddies trap and advect properties such as heat, salt, and biogeochemical tracers, facilitating lateral and vertical transport that rivals or exceeds mean currents in magnitude; for instance, in the , eddy heat fluxes contribute substantially to poleward meridional transport, with estimates indicating up to 1 PW (petawatt) in eddy-driven components. The energetics of mesoscale eddies involve an inverse cascade where energy from smaller scales aggregates into these features, balanced by through interactions with or wind forcing. Observational data from altimetry reveal global eddy maxima in western extensions, such as the Kuroshio and , where eddy amplitudes exceed 50 cm in sea surface height anomalies. In regions like the , eddies modulate deep convection by restratifying the , with vertical velocities on the order of 10-100 m/day enhancing heat and salt fluxes. Submesoscale processes operate at finer resolutions, typically 0.1 to 10 kilometers horizontally and hours to days temporally, bridging the mesoscale and dissipative microscales through frontogenesis and symmetric instabilities. These generate strong vertical velocities, reaching 100-1000 m/day, that drive ageostrophic circulations and upscale energy transfer via a forward cascade, countering the inverse cascade at larger scales. In frontal zones, submesoscale coherent vortices and filaments intensify mixing and nutrient , with model simulations showing contributions to upper ocean flux exceeding 10^{-8} W/kg in wintertime conditions. The interplay between mesoscale and submesoscale features manifests in submesoscale instabilities deforming mesoscale eddies, enhancing vertical exchanges and modulating eddy lifecycles; high-resolution observations in the indicate submesoscale activity peaks during restratification phases, with Rossby numbers exceeding unity signaling departure from geostrophy. Quantitatively, submesoscale processes account for up to 50% of vertical tracer fluxes in the surface , underscoring their role in carbon export and , though direct measurements remain sparse due to sampling challenges. Advances in modeling, such as eddy-resolving simulations at 1/48° resolution, confirm that neglecting submesoscales underestimates ocean heat uptake by 20-30% in coupled projections.

Interannual Climate Modes

Interannual climate modes encompass coupled ocean-atmosphere oscillations operating on timescales of 1 to 7 years, primarily driven by air-sea interactions that alter s (SSTs), depths, and equatorial currents in the tropical oceans. These modes, including the El Niño-Southern Oscillation (ENSO) and the (IOD), modulate global heat redistribution and processes, with ENSO exerting the strongest influence through its impact on equatorial Pacific heat content and wave propagation. ENSO arises from instabilities in the tropical Pacific's mean state, where weakened during El Niño phases allow the warm pool to extend eastward, flattening the and reducing of cold water in the eastern Pacific, leading to anomalies exceeding 2°C. The recharge-discharge oscillator mechanism explains its periodicity: during La Niña, enhanced trades deepen the western and accumulate heat (recharge), which is released equatorially during transitions to El Niño via Kelvin waves, sustaining the cycle every 2-7 years. Observational records, such as the 1997-1998 event with Niño 3.4 index peaks above +2.5 standard deviations, highlight oceanic preconditioning by subsurface anomalies as predictors, with models confirming wave dynamics' role in phase transitions. The IOD, dominant in the tropical , features opposing anomalies between the western (warm) and eastern (cool) basins during positive phases, linked to anomalous equatorial easterlies that deepen the western and enhance off Sumatra-Java. Subsurface structures, evolving via equatorial dynamics, account for much of its interannual variance, with events peaking in boreal autumn and correlating with reduced Indian monsoon rainfall; the 1997 positive IOD coincided with ENSO but demonstrated independent subsurface control. Recent analyses indicate IOD predictability from preceding oceanic heat content, underscoring its role in regional salinity and circulation variability distinct from ENSO teleconnections.

Rapid Dynamic Phenomena

Tides and Tidal Currents

Tides result from the differential gravitational forces exerted by the and on 's oceans, as first explained by in 1687. The 's gravity, being stronger due to its proximity despite weaker mass compared to the Sun, pulls ocean water toward it, creating a bulge on the near side of ; a second bulge forms on the opposite side due to from the Earth- system's rotation. 's rotation relative to these bulges produces the observed periodic rise and fall, with the solar contribution modulating the lunar tides to produce spring tides (higher highs and lower lows) during full and new moons, and neap tides (reduced range) at quarter moons. In the open ocean, tidal amplitudes typically reach about 1 meter, but coastal amplification due to basin resonance and shallowing can exceed 10 meters, as in the where ranges up to 16 meters occur. Tidal cycles vary regionally: semidiurnal tides, predominant globally with two high and two low waters of similar height every (approximately 24 hours 50 minutes), characterize the Atlantic coasts; diurnal tides feature one high and one low per day, common in the ; mixed semidiurnal tides, with unequal highs and lows, prevail on the U.S. Pacific coast. These patterns arise from interference of tidal waves in ocean basins, described by dynamic theory accounting for via the Coriolis effect. Tidal currents are the horizontal water movements driven by the sloping sea surface during tidal rise (flood) and fall (ebb), reversing direction twice daily in semidiurnal regimes. Speeds vary from centimeters per second in deep ocean to several meters per second in straits like the , influencing navigation and . In semi-enclosed basins, currents often follow rotary patterns, rotating around amphidromic points—nodes of zero tidal elevation where water parcels trace ellipses or circles without net vertical motion. Co-tidal lines emanate from these points, with phase progressing counterclockwise in the , reflecting propagation. Tidal dissipation converts approximately 3.5 terawatts of gravitational potential energy into heat via bottom friction and internal wave breaking, with about 20-25% occurring in the deep ocean through topographic scattering rather than solely in shallow seas. This mixing sustains ocean stratification and nutrient upwelling, contributing to global thermohaline circulation. Empirical models from satellite altimetry confirm that barotropic tidal energy fluxes drive internal tides, enhancing vertical mixing rates by orders of magnitude in rough topography regions.

Surface Gravity Waves


Surface gravity waves are oscillatory disturbances at the ocean's surface where the primary restoring force is gravity, distinguishing them from capillary waves dominant at shorter wavelengths below approximately 1.7 centimeters. These waves constitute the bulk of ocean surface motions observed in open waters, with typical periods ranging from seconds to tens of seconds and wavelengths from meters to hundreds of meters. They arise predominantly from wind stress at the air-sea interface, transferring atmospheric momentum to the ocean surface through mechanisms such as pressure perturbations and tangential shear.
The for waves in deep water, where water depth exceeds half the , is given by \omega^2 = [g](/page/G) k, with \omega as , [g](/page/G) as (9.81 m/s²), and k as . This yields a c_p = \sqrt{[g](/page/G) / k} = \sqrt{[g](/page/G) \lambda / 2\pi}, where \lambda is , and a c_g = c_p / 2, indicating that wave energy propagates at half the phase speed. In contrast, shallow-water waves, where depth h < \lambda / 20, become non-dispersive with speed c = \sqrt{[g](/page/G) h}, independent of , facilitating propagation over long distances. Intermediate depths exhibit transitional behavior, with elliptical orbital motions flattening from circular in deep water. Wind generation depends on fetch length, wind duration, and speed; for instance, sustained of 10 m/s over 100 km fetch can produce significant wave heights exceeding 2 meters. Wave breaking, occurring when steepness exceeds about 1/7, dissipates energy into turbulence, enhancing vertical mixing and air-sea . Swells, waves detached from their generating , propagate thousands of kilometers with minimal in deep water due to low frictional losses. These dynamics underpin ocean-atmosphere coupling, with modulating momentum flux and influencing and .

Tsunamis and Internal Waves

Tsunamis are long-period gravity waves generated primarily by rapid vertical displacement of the seafloor due to earthquakes along tectonic faults, with secondary mechanisms including landslides, volcanic eruptions, and impacts. The initial seafloor deformation imparts energy to the overlying , creating a wave train that propagates as a shallow-water wave across ocean basins. Propagation speeds reach 700-800 km/h in deep water, governed by the formula c \approx \sqrt{gh}, where g is (9.81 m/s²) and h is water depth (typically 4000 m in open ocean), resulting in wavelengths of 100-1000 km but amplitudes under 1 m far from shore. Shoaling and nonlinear effects amplify near coasts, often exceeding 10 m and causing inundation, as modeled in hydrodynamic simulations validated against events like the 2004 . In physical oceanography, tsunamis differ from wind-driven surface waves by their non-dispersive nature and sensitivity to , with energy focusing along waveguides like mid-ocean ridges. Numerical models based on shallow-water equations simulate generation and , incorporating initial conditions from seismic , though uncertainties arise in complex regions with multiple faults. Probabilistic hazard assessments integrate these dynamics, estimating recurrence intervals from deposits and historical records, emphasizing empirical validation over purely theoretical . Internal waves arise in density-stratified oceans at interfaces like the pycnocline, where variations in temperature and salinity create stable layering, enabling oscillation under gravity restoration. Primary generation occurs via barotropic tidal currents interacting with seafloor topography, producing internal tides that radiate energy horizontally over thousands of kilometers, with global dissipation rates exceeding 1 TW. Wind forcing and mesoscale instabilities contribute secondarily, but tidal mechanisms dominate deep-ocean mixing, as evidenced by microstructure profiler measurements showing elevated turbulence at wave breaking sites. These facilitate diapycnal mixing essential for ventilating abyssal waters and nutrient , countering with eddy diffusivities of 10^{-5} to 10^{-4} m²/s, far above molecular levels. Nonlinear steepening can form solitary internal with amplitudes up to 100 m and speeds of 1-3 m/s, influencing sediment resuspension and acoustic propagation, as observed in of surface manifestations. strength modulates mode-1 versus higher-mode generation, with weaker profiles favoring energy transfer to smaller scales for dissipation, per laboratory and numerical experiments. In contrast to tsunamis' surface dominance, internal ' subsurface focus underscores their role in sustaining meridional overturning circulation against diffusive tendencies.

Observational Techniques and Recent Advances

Traditional Measurement Methods

Traditional measurement methods in physical oceanography relied heavily on shipboard hydrographic stations, where water samples were collected at discrete depths using Nansen bottles, devices invented by in 1894 and widely employed through the mid-20th century for capturing seawater without contamination from upper layers. These bottles, typically deployed in series on a wire, were triggered to close via messengers—metal weights dropped sequentially—to isolate samples at targeted depths, enabling analysis of vertical profiles of ocean properties. Temperature measurements at depth were obtained using protected reversing thermometers attached to the Nansen bottles, which were calibrated to withstand hydrostatic and designed to lock in readings upon inversion at the desired depth, a technique standard from the early until the 1970s./06:_Temperature_Salinity_and_Density/6.06:_Measurement_of_Temperature) Surface temperatures were often gauged by hauling seawater aboard in insulated buckets or canvas bags and immersing ordinary thermometers, a practice documented in merchant and naval logs persisting into the mid-20th century despite potential cooling biases from evaporation and exposure. Salinity was determined post-collection through chemical of chlorinity, primarily the Knudsen method involving to precipitate ions from the sample, with results converted to via empirical formulas relating chlorinity to total dissolved salts; this approach dominated oceanographic surveys until conductivity-based sensors emerged in the . Depth was inferred from the length of wire paid out, corrected for angle using inclinometers, or from the compression effects on protected thermometers. Ocean currents were measured using mechanical devices like the Ekman current meter, developed in 1903, which featured a that rotated with flow to register speed via revolution counts and a for direction, lowered from ships for short-term deployments at fixed depths. Lagrangian tracking supplemented Eulerian point measurements through drift bottles or surface floats released from ships, recovered ashore to infer mean flow paths, though recovery rates were low and biased toward coastal zones. These methods provided sparse, labor-intensive data, limited by ship time and weather, yielding profiles spaced tens to hundreds of kilometers apart during expeditions.

Satellite and Remote Sensing Technologies

Satellite altimetry measures sea surface height anomalies with centimeter-level precision, enabling the mapping of mesoscale circulation, geostrophic currents, and eddy across global scales. The TOPEX/Poseidon mission, launched on August 10, 1992, by and , pioneered dual-frequency radar altimetry, providing repeat observations every 10 days and revealing basin-scale circulation patterns previously unobserved from ships alone. Subsequent Jason-series missions—Jason-1 (2001), Jason-2 (2008), and (2016)—extended this record, achieving sub-centimeter accuracy in estimates, with the combined dataset spanning over three decades to quantify steric height variations linked to thermosteric expansion. The Surface Water and Topography (SWOT) mission, launched December 16, 2022, introduces wide-swath interferometric altimetry using Ka-band Radar Interferometer (KaRIn), resolving features as small as 10-25 km—ten times finer than nadir-only systems—thus capturing submesoscale dynamics and fine-scale eddies critical for air-sea flux and nutrient transport. Scatterometers employ active radar to derive surface vectors by measuring from capillary and short , offering all-weather, near-real-time data at 25-50 km independent of surface emissivity biases. NASA's QuikSCAT, operational from June 1999 to November 2009, utilized the SeaWinds Ku-band to speeds and directions at 10 m reference height, improving tracking and estimates with root-mean-square errors below 1 m/s against buoys. The Advanced Scatterometer (ASCAT) on EUMETSAT's satellites, deployed since 2007, continues this capability with C-band measurements less susceptible to , providing dual-swath coverage for enhanced ambiguity in high- regimes exceeding 20 m/s. These s reveal -driven phenomena like coastal and vorticity gradients, with validation studies confirming directional accuracy within 20 degrees over ice-free s. Infrared and microwave radiometers quantify () to within 0.5°C, tracing thermal fronts, heat content anomalies, and mode water formation that drive gyre-scale variability. The (AVHRR) on NOAA and platforms, operational since the , uses split-window algorithms at 3.7-12 μm channels to correct for atmospheric , yielding daily global composites at 1 km resolution despite cloud cover limitations. NASA's (MODIS) on Aqua and satellites, active since 2002, employs nonlinear regression of five infrared bands with climatological coefficients, achieving validation against drifters within 0.4 K and enabling detection of diurnal warming cycles up to 2°C in low-wind conditions. Multi-sensor reanalyses, such as the Multi-scale Ultra-high Resolution (MUR) SST product, fuse these with in situ data via optimal , resolving fronts at 0.01° grids and supporting El Niño-Southern Oscillation monitoring with lead times informed by subsurface heat convergence. Emerging () modes on missions like further complement these by imaging surface wave spectra and current slicks, deriving directional wave fields with 100 m resolution for validation and estimation.

In Situ Platforms: Argo Floats and Gliders

In situ platforms such as floats and gliders enable direct, subsurface measurements of ocean temperature, , and velocity, complementing observations by providing vertical profiles essential for resolving three-dimensional ocean . These autonomous systems have revolutionized physical by delivering sustained, high-resolution from remote regions, including the upper 2,000 where much of the ocean's and momentum exchange occurs. floats, in particular, form a global array that has amassed over two million temperature and profiles since deployments began in the late 1990s, supporting analyses of content variability and circulation patterns. Gliders, by contrast, offer targeted, adaptive sampling with greater maneuverability, facilitating process studies of mesoscale features and currents. The program deploys approximately 4,000 profiling floats worldwide, achieving near-global coverage with one float per 3° × 3° grid box in ice-free regions. Each float drifts passively at a parking depth of about 1,000 meters, periodically adjusting to ascend to , measure conductivity-temperature-depth (CTD) profiles from 2,000 meters to the sea , and transmit data via before descending again; cycles typically occur every 10 days. measurements achieve precision of ±0.002°C, while accuracy depends on regional but generally supports estimates of density-driven flows central to . This design yields time series exceeding 20 years, capturing decadal trends in ocean heat uptake—such as a global steric sea-level rise component of 0.8 mm/year from 2005–2019—and interannual signals like El Niño modulation of equatorial currents. Argo data underpin validations of general circulation models, revealing discrepancies in simulated eddy kinetic energy that arise from parameterized subgrid processes rather than observational gaps. Ocean gliders, including models like the Slocum, Seaglider, and Spray, propel themselves through buoyancy-driven gliding along winged hulls, achieving horizontal speeds of 0.2–0.5 m/s and dive cycles to 1,000 meters or more over missions lasting months. Unlike floats' Lagrangian drift, gliders follow commanded waypoints via adjustments in buoyancy and attitude, enabling repeated transects across dynamic features such as fronts or upwelling zones with along-track resolutions of 1–3 km. Slocum gliders, for instance, have been used to map velocity shears in meanders, resolving ageostrophic components that satellites miss due to barotropic aliasing. Seagliders provide near-real-time CTD into forecast models, enhancing predictions of mesoscale eddy propagation speeds, which average 10–20 cm/s in mid-latitudes. These platforms' endurance—up to 6 months on alkaline batteries—supports sustained observations of transient events, though biofouling limits depths below 1,500 meters in productive waters. Integration of glider fleets with enhances spatial heterogeneity, as gliders fill gaps in float sampling during high-variability periods, such as storm-induced mixing. Both platforms face challenges in polar and marginal ice zones, where floats risk entrapment and gliders contend with under-ice navigation, yet expansions like deep-cycle (to 6,000 meters since ) and bio-optical gliders extend their utility for full-depth circulation monitoring. Data from these systems reveal causal links in ocean physics, such as forcing driving meridional overturning rates of 15–20 Sverdrups in , grounded in empirical density gradients rather than model assumptions.

Modeling Challenges and Validation

Physical oceanography modeling grapples with the inherent multi-scale nature of oceanic flows, spanning global basin-wide circulation to submesoscale features under 10 km, where interactions between rotation, , and nonlinearity dominate energy cascades. Resolving these requires horizontal grid spacings finer than 1/10° (approximately 10 km), but even eddy-permitting models at 1/4° struggle with accurate representation of mesoscale eddies, which account for much of the ocean's and influence tracer transport. Computational constraints limit global simulations to resolutions around 1/10° in current high-end configurations, necessitating parameterizations for unresolved processes and risking suppression of intrinsic variability. Parameterization schemes for subgrid phenomena, such as eddy-induced via the Gent-McWilliams or diapycnal mixing tied to internal breaking, introduce uncertainties due to incomplete physical understanding and sensitivity to tuning. For instance, the Gent-McWilliams scheme, implemented since , often over-dissipates in coarser grids, leading to underestimated restratification and biased boundary currents like the . Air-sea turbulence and wave-driven mixing further complicate representations, as models inadequately capture episodic events without explicit nonhydrostatic dynamics, which demand even higher vertical resolution. These approximations propagate errors in long-term simulations, particularly for heat and freshwater budgets. Validation relies on hindcasting against diverse observations, including altimetry for sea surface height, float trajectories for subsurface velocities, and moored arrays for currents, employing metrics like direction of difference (DOD), percentage difference in velocity (PDOV), root-mean-square error, and Taylor skill scores. Ocean Model Intercomparison Project (OMIP) protocols, using standardized forcings like JRA55-do from 1948–2007, evaluate over 20 models and expose inter-model spreads, such as in strength varying by factors of two in CORE-II experiments spanning 1988–2007. High-resolution variants, like GFDL's CM4X at 1/8°, reduce biases by up to 40% compared to coarser counterparts after 150-year spin-ups, yet persist in diffusing western boundary currents. Quantitative assessments using 842,421 observations from 2001–2020 against models like ECCO2 and GLORYS12 reveal profound limitations: only 3.8% of mid-depth (∼1000 m) circulation is accurately captured, with velocities underestimated globally and errors exceeding 45° in 81.1% of the , reaching 70% deficits in western boundary currents like the Agulhas. Such disparities underscore parameterization failures in and high-frequency dynamics, impairing reliability for projections of and sea-level variability, and highlight the need for enhanced to mitigate systematic biases.

Controversies and Debates

Attribution of Circulation Changes

Attribution of changes in ocean circulation involves distinguishing between natural internal variability and responses to external forcings, primarily gases and aerosols, using detection-attribution methods that compare observations to simulations with and without forcing. These methods, such as optimal fingerprinting, seek signals in sea surface height, temperature profiles, and transport estimates that exceed expected variability from unforced models. Challenges arise from the long timescales of circulation modes, which can span centuries, against observational records often limited to decades. In the Atlantic Meridional Overturning Circulation (AMOC), proxy and direct measurements indicate a slowdown of approximately 15% since the mid-20th century, with transport at 26.5°N decreasing from about 20 in the to around 17 Sv in recent years. Attribution studies link this trend to anthropogenic forcing through increased freshwater input from ice melt and altered density gradients, as simulated in CMIP6 models where AMOC weakens under rising CO2 levels. However, a 2025 analysis using century-long North Atlantic air-sea data concludes no significant AMOC decline since the mid-20th century, attributing apparent weakenings to measurement artifacts or decadal variability like the Atlantic Multidecadal (AMO). This discrepancy highlights reliance on sparse in-situ data, such as the array operational only since , versus longer but indirect proxies. For wind-driven gyres, observed strengthening of subtropical cells from 1979 to 2014 has been attributed to changes in surface winds, amplifying by redistributing heat equatorward and enhancing upper-ocean heat uptake. In the , westerly wind intensification is often ascribed to and greenhouse gases, yet multi-decadal reanalyses show internal variability dominating circulation shifts, with externally forced signals emerging only in the late . Attribution here is complicated by coupled air-sea interactions and model biases in simulating compensation, where observed circulation responses to winds are weaker than predicted. Debates center on the , where high internal variability—such as interdecadal Pacific and Atlantic oscillations—masks trends in short records, leading to overconfidence in some model-based attributions. Peer-reviewed critiques note that models often underestimate historical variability while projecting forced changes, potentially inflating contributions; for instance, studies emphasize that decadal fluctuations can mimic long-term trends without requiring external forcing. Conversely, detection studies detect fingerprints in interior ocean properties earlier than surface, supporting causal links to via heat and carbon uptake alterations. Overall, while some circulation elements show emerging forced signals, definitive attribution remains contested due to observational limitations and modeling uncertainties, with natural variability likely explaining much of the multidecadal changes observed to date.

Reliability of Climate Models for Ocean Projections

Climate models, particularly those within the Coupled Model Intercomparison Project Phase 6 (CMIP6), exhibit significant challenges in reliably projecting ocean dynamics due to coarse spatial resolutions typically ranging from 0.5° to 1° latitude-longitude, which inadequately resolve mesoscale eddies and submesoscale processes critical for vertical mixing and heat transport. These parametrizations introduce uncertainties, as evidenced by inter-model spreads in simulated ocean heat uptake, where CMIP6 ensembles show biases in regional heat content, such as excessive warming in the subpolar North Atlantic compared to observations from Argo floats and ship-based measurements spanning 2004–2020. Globally, while CMIP6 models reproduce the observed increase in upper ocean heat content of approximately 0.4–0.6 W/m² imbalance from 1971–2018, they diverge in deep ocean penetration, underestimating abyssal warming detected by Deep Argo profiles below 2000 meters. Projections of ocean circulation, including the Atlantic Meridional Overturning Circulation (AMOC), reveal substantial model-dependent variability, with CMIP6 simulations predicting AMOC weakening of 20–50% by 2100 under high-emission scenarios, yet failing to hindcast observed AMOC stability over the instrumental record since the 1950s or rapid past fluctuations inferred from proxy data like sediment cores. This discrepancy raises doubts about forward projections, as models that poorly simulate historical AMOC variability—such as multidecadal oscillations linked to North Atlantic salinity anomalies—may overestimate tipping risks, with recent analyses across 34 CMIP6 models indicating resilience to extreme forcings without collapse this century. Uncertainties stem partly from uncertain freshwater forcing and air-sea flux parametrizations, contributing to divergent European climate impacts in projections. Validation against empirical data underscores persistent limitations; for instance, CMIP6 projections of sea surface salinity trends align broadly with satellite observations from 2000–2020 but exhibit errors in subtropical gyre intensities due to misrepresented wind-driven circulation. Emerging data-driven approaches, leveraging on observational datasets like World Ocean Atlas, aim to mitigate dynamical biases but remain unproven for long-term projections beyond decadal scales. Overall, while models capture large-scale trends in rising by 436 ± 11 ZJ from 2005–2023, their reliability for fine-scale ocean projections is compromised by structural uncertainties, necessitating cautious interpretation in policy contexts.

Natural Variability vs. Anthropogenic Forcing

In physical oceanography, distinguishing natural variability from anthropogenic forcing is essential for attributing changes in ocean temperature, circulation, and heat content. Natural variability encompasses internal climate modes such as the El Niño-Southern Oscillation (ENSO) on interannual scales, the Pacific Decadal Oscillation (PDO) and Atlantic Multidecadal Oscillation (AMO) on decadal to multidecadal scales, which redistribute heat and momentum through ocean-atmosphere interactions, causing fluctuations in sea surface temperatures (SSTs) and currents without external drivers. Anthropogenic forcing, primarily from greenhouse gas emissions, imposes a radiative imbalance leading to net ocean heat uptake, with estimates indicating that oceans have absorbed over 90% of excess heat since the mid-20th century. However, detection and attribution studies reveal challenges, as natural modes can amplify or mask forced trends, particularly on timescales shorter than a century. Observations of upper-ocean heat content (OHC) show a robust increase since the , with multi-model analyses attributing the majority of this warming to forcings using optimal fingerprinting techniques that compare observed patterns to simulated responses. For instance, between 1866 and the present, factors explain most of the multidecadal rise in global and ocean surface , though interannual and decadal fluctuations align with PDO and AMO phases. Positive AMO phases, occurring roughly every 60-80 years, correlate with warmer North Atlantic SSTs, contributing up to 0.3°C to global anomalies, potentially confounding attribution during recent decades. Critics note that climate models often underestimate the amplitude of these natural oscillations, leading to overconfidence in forced signals, as evidenced by discrepancies in simulating the 1998-2013 warming partly linked to negative PDO . Regarding ocean circulation, the Atlantic Meridional Overturning Circulation (AMOC) exhibits pronounced natural variability of about 5 Sverdrups (Sv) on decadal timescales, dominating observed changes through the and into recent observations up to 2022, with no reliable detection of an slowdown signal yet. Projections from coupled models anticipate a 10-50% AMOC weakening by 2100 under high-emission scenarios due to ice melt-induced freshwater stratification, but empirical proxies like subpolar gyre indices show fluctuations consistent with internal variability rather than a secular decline. Similarly, heat redistribution during PDO shifts has driven apparent multidecadal trends in equatorial warming, complicating separation from effects. These findings underscore that while forcing is detectable in long-term OHC trends, short-term attributions risk conflating internal dynamics, particularly given observational uncertainties in deep ocean coverage before the era (post-2000). The interplay influences and biogeochemical cycles, where natural variability like ENSO modulates steric height by 10-20 cm regionally, while anthropogenic contributes a global mean of about 1.5 mm/year since 1993. Debates persist on emergence timescales, with some analyses indicating that forced signals in AMOC or properties may not surpass natural noise until mid-century or later, urging caution in policy-driven interpretations that downplay variability. Peer-reviewed assessments emphasize empirical over model consensus alone, highlighting that academia's institutional incentives may favor narratives emphasizing human causation, yet data-driven separations reveal natural processes as co-dominant on observational horizons.

References

  1. [1]
    Physical Oceanography
    Physical Oceanography is the field of study that deals with mechanisms of energy transfer through the sea and across its boundaries.
  2. [2]
    What is Oceanography? - Texas A&M College of Arts and Sciences
    Physical oceanography involves the study of the properties (temperature, density, etc.) and movement (waves, currents, and tides) of seawater and the ...
  3. [3]
    Physical Oceanography - Athens, Georgia - UGA Marine Sciences
    Physical Oceanography applies the laws of physics to investigate the structure and dynamics of ocean circulation and water properties.Missing: definition | Show results with:definition
  4. [4]
    Physical Oceanography - Rutgers-Marine Sciences
    Physical oceanography studies ocean currents driven by temperature and salinity differences, using instruments like thermometers and salinometers.
  5. [5]
    [PDF] Physical Oceanography Division
    The Physical Oceanography Division carries out interdisciplinary scientific investigations of the physics of ocean currents and water properties to assess and ...
  6. [6]
    [PDF] What is Physical Oceanography? | WHOI GFD
    Oceanography is a huge topic, that definitely includes the study of all life in the oceans, but also topics such as waves, currents, rising sea level, ...
  7. [7]
    [PDF] Brief History of Physical Oceanography
    Early charting of the ocean's surface currents came hand in hand with exploration of coastlines and ocean basins and was performed by the earliest seafaring ...
  8. [8]
    Charting the Gulf Stream | Worlds Revealed
    Jan 7, 2016 · Folger and Franklin jointly produced a chart of the Gulf Stream in 1768, first published in London by the English firm Mount and Page.
  9. [9]
    [PDF] 1768 chart of the Gulf Stream by Benjamin
    Franklin also compiled many meticulous temperature measurements that helped delineate the Gulf Stream. (Map image courtesy Library of Congress,. Geography and ...
  10. [10]
    The tidal measurements of James Cook during the voyage ... - HGSS
    We conclude that Cook's tidal measurements were accurate in general to about 0.5 ft (15 cm) in height and 0.5 h in time.
  11. [11]
    View of Captain Cook's Influence on Hydrographic Surveying
    ... eighteenth century, when James Cook firstwent to sea in Captain John Walder's Whitby Colliers.In the year 1693, two important atlases of sea charts were ...
  12. [12]
    Biography: Matthew Fontaine Maury
    Oct 12, 2022 · He published the Wind and Current Chart of the North Atlantic, which showed sailors how to use the ocean's currents and winds to their advantage ...
  13. [13]
    Impacts of Maury's Navigation Charts on Sailing Time, 1850s
    The navigator Matthew Fontaine Maury collected between 1842 and 1861 an extensive array of ship logs, enabling him to chart prevailing winds and sea currents.
  14. [14]
    History: Timeline: NOAA Office of Ocean Exploration and Research
    January 3, 1840, Sir James Clark Ross takes the first modern sounding in the deep sea at Latitude 27°S Longitude 17°W. 1842. Darwin publishes The Structure and ...
  15. [15]
    [PDF] PHYSICAL OCEANOGRAPHY - University of Colorado Boulder
    Nov 9, 2014 · 2.3 Milestones in the Understanding of the Ocean . ... More information on the history of physical oceanography can be found in Ap-.
  16. [16]
    Global Oceanographic Data Archeology and Rescue Project (GODAR)
    The Global Oceanographic Data Archaeology and Rescue (GODAR) Project was created to consolidate and digitize physical oceanographic data.
  17. [17]
    100 Years of the Ocean General Circulation in - AMS Journals
    Jan 1, 2018 · Observations and explanations before 1945. Any coherent history of physical oceanography must begin not in 1919 but in the nineteenth century, ...
  18. [18]
    Operational Oceanography | American Scientist
    During the Cold War, military funding shaped deep-sea research priorities, which resulted in a new understanding of forces below the water's surface.
  19. [19]
    HENRY STOMMEL | Biographical Memoirs: Volume 72
    Stommel, probably the most original and important physical oceanographer of all time, was in large measure the creator of the modern field of dynamical ...
  20. [20]
    CTD 101 - Ocean Observatories Initiative
    Jun 15, 2011 · Advancements in these individual sensors culminated in 1969 with the invention of the microprofiler, or CTD, by Neil Brown at Woods Hole ...
  21. [21]
    A numerical method for the study of the circulation of the world ocean
    A model is presented for studying ocean circulation problems taking into account the complicated outline and bottom topography of the World Ocean.
  22. [22]
    Climate Calculations with a Combined Ocean-Atmosphere Model in
    For the fulldetails the reader is referred to Manabe (1969) andBryan (1969). The numerical model of the atmosphere is verysimilar to that described by Manabe -t ...
  23. [23]
    Is Computational Oceanography Coming of Age? in - AMS Journals
    Computational oceanography is the study of ocean phenomena by numerical simulation, especially dynamical and physical phenomena.Maturation of computational... · Opportunities for... · Prospects for computational...
  24. [24]
    How deep is the ocean? - NOAA's National Ocean Service
    Jun 16, 2024 · The average depth of the ocean is about 3,682 meters (12,080 feet). The deepest part of the ocean is called the Challenger Deep and is located ...
  25. [25]
    1.1 Overview of the Oceans – Introduction to Oceanography
    The average depth of the world ocean is about 3800m (12,500 ft), which is about four times deeper than the average land elevation is high (840m or 2800 ft). In ...Missing: topography | Show results with:topography
  26. [26]
    Ocean floor features - NOAA
    Ocean floor features · Underwater landforms · Plate tectonics and the ocean floor · Hydrography · Marine life and exploration on the ocean floor · Hydrothermal vents.
  27. [27]
    Pacific Ocean - 2022 World Factbook Archive - CIA
    Dec 27, 2022 · ... Pacific Ocean is the deepest ocean basin averaging 4,000 m in depth ... Pacific Northeast region (Region 67) is the fourth largest Pacific Ocean ...
  28. [28]
    Layers of the Ocean - NOAA
    Mar 28, 2023 · The deepest zone of the ocean, the hadalpelagic zone extends from 6,000 meters (19,700 feet) to the very bottom, 10,994 meters (36,070 feet) in ...
  29. [29]
    Our Blue Planet - My Framer Site - UNESCO Ocean Literacy
    Aug 19, 2022 · The Pacific is the largest ocean basin, covering 32% of Earth's surface, with the deepest point, Challenger Deep, at 11,034m. It has twice the ...
  30. [30]
    1.2 Continental Margins – Introduction to Oceanography
    After the rise comes the abyssal plain, or the deep ocean floor, lying between 4500 – 6000 m. The abyssal plain includes most of the ocean floor, and is the ...
  31. [31]
    Ocean Temperature Profiles - University of Hawaii at Manoa
    Ocean surface temperatures range from -2°C to 36°C, with the average being 17°C. The average of all ocean water is 3.5°C. Water layers form based on  ...Missing: oceanography | Show results with:oceanography
  32. [32]
    6.2 Temperature – Introduction to Oceanography
    Ocean temperatures range from -2°C to 30°C, with surface water warmest, and the average ocean temperature is about 4°C. Deep ocean temperature is around 2°C.
  33. [33]
    Key Physical Variables in the Ocean: Temperature, Salinity, and ...
    a) Vertical profiles of in-situ temperature t, potential temperature θ, and Conservative Temperature Θ. Inset shows an expanded view of deep ocean values.
  34. [34]
    Seasonal Variation in Ocean Temperature Vertical Profiles
    Ocean temperature profiles vary seasonally depending on latitude. The equator and poles show little seasonal change, while mid-latitudes have seasonal ...Missing: oceanography | Show results with:oceanography
  35. [35]
    Overview - NASA Salinity
    On average, sea surface salinity is about 35 PSS. Over the globe, sea surface salinity varies from 32 to 37 PSS. Salinity variations are caused by precipitation ...
  36. [36]
    Sea Water | National Oceanic and Atmospheric Administration
    Mar 28, 2023 · In sea water, there is typically close to 35 grams of dissolved salts in each liter (35ppt), but ranges between 33-37 grams per liter (33ppt - ...Missing: PSU | Show results with:PSU
  37. [37]
    Science Overview - NASA Salinity
    In general, ocean regions dominated by evaporation have higher salinities and areas with high precipitation have lower salinities. Global view of sea surface ...Missing: controls | Show results with:controls
  38. [38]
    [PDF] Salinity Patterns in the Ocean
    Salinity is changed near the sea surface by precipitation and evaporation of fresh water and by salty water produced sea ice forms and excludes salt.
  39. [39]
    Ocean salinity - Science Learning Hub
    Salinity is controlled by a balance between water removed by evaporation and freshwater added by rivers and rain. The Mediterranean Sea in Europe has very high ...Missing: precipitation input
  40. [40]
    Spatial-temporal structure of ocean salinity seasonal variation
    Nov 18, 2021 · Salinity is always regarded as a natural water gauge. It plays a vital role in regulating ocean density, stratification and circulation.
  41. [41]
    Spatiotemporal Variability of Rainfall and Surface Salinity in the ...
    Dec 13, 2023 · Sea surface salinity (SSS) is strongly affected by local rainfall. The Eastern Pacific Fresh Pool (EPFP) is one of the lowest salinity regions ...
  42. [42]
    Sea Water Equation of State Calculator - JHU/APL
    The underlying equations are valid for temperatures from -2 to 35 deg C, pressures from 0 to 10,000 dbar, and practical salinity from 2 to 42. Numerics: The ...
  43. [43]
    A seasonal climatology of the upper ocean pycnocline - Frontiers
    The upper ocean vertical structure can be decomposed into several layers based on density and stratification (Sprintall and Cronin, 2009). Near the surface ...
  44. [44]
    Subtropical Mode Water and Permanent Pycnocline Properties in ...
    Feb 4, 2019 · The upper part of the permanent pycnocline is half thinner (80 m) than the lower part (150 m), which leads to an average form factor of 50% ( ...
  45. [45]
    Eddy-induced pycnocline depth displacement over the global ocean
    The pycnocline is almost in the depth layer where the eddy core is located (about 227 m for the global average, see Fig. S2 for the geographical distribution), ...
  46. [46]
    Summertime increases in upper ocean stratification and mixed layer ...
    For example, increased pycnocline stratification will expectedly weaken surface-to-depth exchanges as enhanced density gradients decouple surface and subsurface ...
  47. [47]
    Generation of the Southern Ocean pycnocline by sea ice-ocean ...
    The ocean's permanent pycnocline is a layer of elevated stratification ... This is a preprint and has not been peer reviewed. Data may be preliminary ...
  48. [48]
    Pycnocline - an overview | ScienceDirect Topics
    Mixing across the pycnocline induces vertical exchange of nutrients, mainly by IWs, sustaining a substantial proportion of the total oceanic primary production.
  49. [49]
    Navier-Stokes Equations - NASA
    These equations describe how the velocity, pressure, temperature, and density of a moving fluid are related. The equations were derived independently by G.G. ...
  50. [50]
    [PDF] Atmospheric and Oceanic Fluid Dynamics
    These books are on the fundamentals of geophysical fluid dynamics (gfd); waves, instabilities and turbulence; atmospheric circulation; and ocean circulation.
  51. [51]
    Navier-Stokes and tracer conservation equations · Oceananigans.jl
    It is an accurate approximation for many flows, and especially so for oceanic flows where density differences are very small. See Vallis (2017, section 2.4) for ...
  52. [52]
    [PDF] Lecture 18 Ocean General Circulation Modeling
    Instead we make a series of approximations to simplify the Navier-Stokes equations to yield the “primitive equations” which are the basis of most general circu.
  53. [53]
    [PDF] Lecture 3: 5 Oct Momentum equation
    much of the atmospheric and oceanic circulation the vertical momentum balance is nearly the same as if the fluid were at rest…hydrostatic. These fluids are ...
  54. [54]
    [PDF] Formulating the Equations of Ocean Models
    For fluid motion, we are led to notions of generalized vertical coordinates, which are a critical element in theoretical and numerical models of the ocean.
  55. [55]
    Momentum Balance of the Wind-Driven and Meridional Overturning ...
    When a force is applied to the ocean, fluid parcels are accelerated both locally, by the applied force, and nonlocally, by the pressure gradient forces ...
  56. [56]
    Ocean Barotropic Vorticity Balances: Theory and Application to ...
    Apr 17, 2023 · It involves the balance between the beta effect caused by meridional flows and the curl (vertical vorticity) of the turbulent momentum stress ...2 Vorticity Balances Of The... · 3 Discretization In An Ocean... · 4.2 Barotropic Vorticity...
  57. [57]
    [PDF] coriolis Effects on Wind-Driven ocean currents
    Here we study the mechanism by which the wind drives ocean circulation. We induce cyclonic and anticyclonic circulation by blowing air over the surface of a ...
  58. [58]
    [PDF] Effects of rotation on a sphere: f- and β-planes
    the Coriolis parameter, f , is the vertical component of the rotation vector. This approximation will shaky in high latitude regions where vertical.
  59. [59]
    [PDF] 2.011 Geostrophic Currents
    Apr 25, 2006 · The geostrophic balance does not apply within about 2◦ of the equator where the Coriolis force goes to zero because sinϕ → 0. 3. The geostrophic ...
  60. [60]
    Coriolis Parameter - an overview | ScienceDirect Topics
    The Coriolis parameter f = 2Ω sin θ varies with latitude θ. However, this variation is important only for phenomena having very long time scales.<|control11|><|separator|>
  61. [61]
    Geostrophic balance – Physics Across Oceanography
    The steady-state force balance between the Coriolis force and the pressure gradient force is called geostrophic balance and is the approximate state of most ...
  62. [62]
    Geostrophy Assessment and Momentum Balance of the Global ...
    Sep 17, 2021 · The geostrophic balance, which is a balance between the Coriolis force and the pressure gradient force, is a fundamental assumption that enables ...
  63. [63]
    Geostrophic Velocity - Upper Ocean Dynamics Laboratory
    Changes in sea-surface height determine general flow characteristics called geostrophic flow, the balance between the Coriolis force and the horizontal pressure ...
  64. [64]
    [PDF] a Coriolis tutorial, Part 4: - Wind-driven ocean circulation
    The Coriolis parameter f is f(y) via the β-plane approximation, Eqn. (8), which is crucially important in what follows. There are two steady forces ...
  65. [65]
    Theory of the Turbulent Drift Friction Layer of the Ocean in
    The paper proposes a theory of the stationary turbulent drift friction layer that, under the conditions of stable stratification, can appear as the upper mixed ...
  66. [66]
    An empirical formula of bottom friction coefficient with a dependence ...
    Oct 17, 2023 · Bottom friction is a major factor in tidal dissipation and is usually parameterized by the bottom friction coefficient (BFC). BFC is a critical ...Abstract · Introduction · Empirical formula obtained... · Discussions
  67. [67]
    The Role of Bottom Friction in Mediating the ... - AMS Journals
    Bottom friction, especially from dense water overflows, weakens the Weddell Gyre's circulation by acting as a frictional sink of cyclonic vorticity.
  68. [68]
    Wave-Breaking Turbulence in the Ocean Surface Layer in
    Jun 1, 2016 · Wave breaking at the ocean surface limits wave growth (Melville 1994), enhances gas exchange (Zappa et al. 2007), and generates turbulence that mixes the ocean ...
  69. [69]
    Estimating Rain-Generated Turbulence at the Ocean Surface Using ...
    Jun 14, 2019 · Turbulence is critical in destratifying the surface layer, and although wind-related mechanisms (shear stress, wave breaking) dominate the ...
  70. [70]
    Global Patterns of Diapycnal Mixing from Measurements of the ...
    Jul 1, 2014 · Over rough or abrupt ridge topography, the turbulent diffusivity is bottom enhanced compared to smooth topography (Fig. 7). Given that internal ...
  71. [71]
    Breaking Internal Waves and Ocean Diapycnal Diffusivity in a High ...
    Jun 5, 2024 · However, the high intermittency of turbulent mixing makes it extremely difficult to measure effective diapycnal mixing observationally. In fact, ...Abstract · Introduction · Internal Waves in the High... · Ocean Diapycnal Diffusivity...
  72. [72]
    [PDF] 3.3: Diapycnal mixing processes in the ocean interior
    Mar 11, 2013 · Diapycnal mixing in the ocean interior is driven by various processes, playing a key role in the ocean's thermodynamic balance and upwelling.
  73. [73]
    Topographic enhancement of vertical turbulent mixing in the ...
    Mar 6, 2017 · The diapycnal and isopycnal mixing experiment in the Southern Ocean found the turbulent diffusivity inferred from the vertical spreading of a ...
  74. [74]
    Wind Driven Surface Currents: Gyres Background - Ocean Motion
    In this section, we examine the characteristics of wind-driven surface-ocean currents, components of the huge gyres that dominate the central regions of the ...
  75. [75]
    9.1 Surface Gyres – Introduction to Oceanography
    Furthermore, wind-driven surface currents only affect the top 100-200m of water, meaning surface currents only involve about 10% of the world's ocean water.
  76. [76]
    Sea Surface Currents and Temperature (vegetation on land)
    May 1, 2012 · Sea surface currents are driven by winds and solar energy, forming subtropical gyres. They move at 3% wind speed, with clockwise and ...<|control11|><|separator|>
  77. [77]
    [PDF] Lecture 10: Ocean Circulation - UCI ESS
    The Sverdrup balance, or Sverdrup relation, is a theoretical relationship between the wind stress exerted on the surface of the open ocean and the vertically.
  78. [78]
    [PDF] Chapter 10 The wind-driven circulation
    This simple formula is remarkably successful – it predicts the sense of circulation and volume transport of all the major ocean gyres, rationalizing the ...
  79. [79]
    What is a gyre? - NOAA's National Ocean Service
    Jun 16, 2024 · There are five major gyres, which are large systems of rotating ocean currents. The ocean churns up various types of currents.
  80. [80]
    Pacific Ocean upper layer circulation
    Nov 11, 2019 · The tropical gyre is also very narrow due to the wind pattern which causes it, and its vertical penetration is also not clear. Current ...<|separator|>
  81. [81]
    Currents, Gyres, & Eddies - Woods Hole Oceanographic Institution
    They are formed primarily by wind blowing across the surface of the ocean and by differences in the temperature, density and pressure of water and are steered ...
  82. [82]
    On Wind‐Driven Energetics of Subtropical Gyres - AGU Journals
    Jan 20, 2021 · In this paper, we argue that the ocean receives energy over the broader interior from the wind and recirculates that energy to the open ocean Gulf Stream.
  83. [83]
    2. Wind driven surface currents - Ocean Currents
    In the subtropics, these currents form huge gyres with warm, fast boundary currents on the western side of the oceans and slow, cold boundary currents on the ...
  84. [84]
    The Ekman Spiral - Currents - NOAA's National Ocean Service
    The Ekman spiral, named after Swedish scientist Vagn Walfrid Ekman (1874-1954) who first theorized it in 1902, is a consequence of the Coriolis effect.<|separator|>
  85. [85]
    SIO 210: Introduction to Physical Oceanography
    Oct 29, 2019 · This direction of the "Ekman transport" is independent of the exact details of the spiral, hence exact details of the vertical eddy viscosity. ...
  86. [86]
    9.3 The Ekman Spiral and Geostrophic Flow
    A spiraling pattern of water motion called the Ekman spiral, which usually penetrates to about 100 m deep before the motion ceases.
  87. [87]
    [PDF] ATOC 5051 INTRODUCTION TO PHYSICAL OCEANOGRAPHY
    c) As we shall see below: Ekman transport is the vertical integral of Ekman spiral – 90deg to the right (left) of wind in Northern Hemisphere (Southern ...
  88. [88]
    Wind Energy Input to the Ekman Layer* - AMS Journals
    Although the Ekman theory has been the backbone of modern dynamical oceanography, the Ekman spiral predicted by classical theory has not been exactly ver- ified ...
  89. [89]
    Thermohaline Circulation - Fact Sheet by Stefan Rahmstorf
    The thermohaline circulation (Fig. 1) is that part of the ocean circulation which is driven by density differences.
  90. [90]
    The World Ocean Thermohaline Circulation in - AMS Journals
    Sep 1, 2012 · The World Ocean thermohaline circulation is frequently idealized as a conveyor belt transporting heat and freshwater from the Indo–Pacific to ...Missing: mechanism | Show results with:mechanism
  91. [91]
    Thermohaline Circulation - an overview | ScienceDirect Topics
    The thermohaline circulation is that part of the ocean circulation which is driven by fluxes of heat and freshwater across the sea surface and subsequent ...
  92. [92]
    Thermohaline Circulation - Currents - NOAA's National Ocean Service
    Currents Tutorial​​ Thermohaline circulation begins in the Earth's polar regions. When ocean water in these areas gets very cold, sea ice forms. The surrounding ...
  93. [93]
    Deep Water Formation - an overview | ScienceDirect Topics
    Deep water formation takes place in a few localized areas (Figure 1): the Greenland–Norwegian Sea, the Labrador Sea, the Mediterranean Sea, the Weddell Sea, and ...
  94. [94]
    Earth scientists reveal how Atlantic Ocean circulation has changed ...
    Aug 14, 2025 · This is approximately 18 Sverdrups, with one Sverdrup corresponding to a volumetric flow rate of 1 billion liters per second. "Our findings ...Missing: facts | Show results with:facts
  95. [95]
    Atlantic Meridional Overturning Circulation and Its Impact on Global ...
    The Atlantic Meridional Overturning Circulation (AMOC) is characterized by northward flowing warm and salty water in the upper layer of the Atlantic Ocean, ...
  96. [96]
    Ocean Conveyor Belt - National Geographic Education
    May 27, 2025 · Thermohaline circulation moves a massive current of water around the globe, from northern oceans to southern oceans, and back again.
  97. [97]
    Decades of Data on a Changing Atlantic Circulation | News
    Apr 24, 2024 · The NCEI and UMD scientists found that AMOC remained stable and consistent from 1955 to 1994 but declined in strength and speed in the last two decades.
  98. [98]
    Abrupt climate change and thermohaline circulation - PNAS
    The global thermohaline circulation (THC) consists of cooling-induced deep convection and sinking at high latitudes, upwelling at lower latitudes, and the ...
  99. [99]
    Advancing our understanding of the Atlantic Meridional Overturning ...
    Jan 22, 2025 · Therefore, the amount of anthropogenic carbon sequestered by the ocean is closely tied to the strength of the AMOC. Researchers at AOML made ...
  100. [100]
    CMIP6 Models Predict Significant 21st Century Decline of the ...
    We explore the representation of the Atlantic Meridional Overturning Circulation (AMOC) in 27 models from the CMIP6 multimodel ensemble.
  101. [101]
    [PDF] Chapter 13 Western Boundary Currents
    Strong, persistent currents along the western boundaries of the world's major ocean basins are called "western boundary currents" (WBCs).
  102. [102]
    Gulf Stream - an overview | ScienceDirect Topics
    The Gulf Stream is a semicontinuous, narrow (∼ 100 km wide) current jet with fastest speeds of 200–250 cm s−1 located at the surface decreasing to around 20 cm ...
  103. [103]
    Kuroshio | Japan, Current, Location, & Map | Britannica
    Only about 1,300 feet (400 metres) deep, the Kuroshio travels at rates ranging between 20 and 120 inches (50 and 300 cm) per second. Flowing past Taiwan ( ...
  104. [104]
    9.4 Western Intensification – Introduction to Oceanography
    To move the same volume of water through each side, western boundary currents are faster, deeper, and narrower than eastern boundary currents. The center of ...
  105. [105]
    11.2: Western Boundary Currents - Geosciences LibreTexts
    Nov 11, 2024 · We now know that the variation of Coriolis force with latitude is required for the existence of the western boundary current, and that other ...<|separator|>
  106. [106]
    Western boundary current theory - UHSLC
    The main idea is that potential vorticity is conserved following a water column, so a western boundary return flow to the north (that is, f is increasing) ...
  107. [107]
    [PDF] the westward intensification of wind-driven ocean currents
    Abstract--A study is made of the wind-driven circulation in a homogeneous rec- tangular ocean under the influence of surface wind stress, linearised bottom ...
  108. [108]
    [PDF] Boundary Layers: Homogeneous Ocean Circulation - WHOI GFD
    The first explanation for the western intensification of the wind-driven ocean circulation was provided by Henry Stommel (1948). The following chapter considers ...
  109. [109]
    The Stommel-Munk Correspondence, 1947–1953 | Oceanography
    Jan 12, 2021 · Two of the dominating figures of twentieth century physical oceanography/geophysics were Henry (Hank) Stommel and Walter Munk, whose early correspondence is ...
  110. [110]
    Advancing Observations of Western Boundary Currents
    Feb 27, 2025 · Western boundary currents (WBCs) are crucial components of the global ocean circulation, responsible for transporting water (momentum), heat, ...Missing: explanation | Show results with:explanation
  111. [111]
    Wind Stress Drag Coefficient over the Global Ocean in - AMS Journals
    The wind stress drag coefficient (C_D) is a dimensionless value used to calculate wind stress. It has a global mean of ≈ 1.25 × 10^-3, but can change ...
  112. [112]
    [PDF] Coupled effects of ocean current on wind stress in the Bay of Bengal
    ... oceans. 46 (Wunsch, 1998). According to the bulk aerodynamic formula, the wind stress (τ) is calculated as. 47 τ = ρ C a D ( u a − u o ) u u a − o. (1),. 48 ...
  113. [113]
    4.6: Wind Stress - Geosciences LibreTexts
    Nov 11, 2024 · where ρ a = 1.3 kg/m 3 is the density of air, U 10 is wind speed at 10 meters, and C D is the drag coefficient.
  114. [114]
    Revised Estimates of Ocean Surface Drag in Strong Winds - Curcic
    May 7, 2020 · In high winds, the drag coefficient is saturated between 2.3×10−3 and 2.9×10−3, with an average value of 2.49×10−3. The drag saturation signal ...
  115. [115]
    On the Exchange of Momentum over the Open Ocean in
    Aug 1, 2013 · This study investigates the exchange of momentum between the atmosphere and ocean using data collected from four oceanic field experiments.<|separator|>
  116. [116]
    Improving Wave‐Based Air‐Sea Momentum Flux Parameterization in Mixed Seas
    ### Summary of Air-Sea Momentum Flux Role in Ocean-Atmosphere Coupling
  117. [117]
    Surface Currents and Relative-Wind Stress in Coupled Ocean ...
    Ocean currents and atmospheric winds are coupled by the exchange of momentum across the air–sea interface, the strength of which depends on the relative wind, ...
  118. [118]
    Mesoscale atmosphere ocean coupling enhances the transfer of ...
    Jun 13, 2016 · This pathway emerges, as the wind stress exerted by the atmosphere on the ocean depends not only on the motion of the air but also on the ...
  119. [119]
    Air-Sea Fluxes With a Focus on Heat and Momentum - Frontiers
    The net radiative component of the air-sea heat flux comprises a shortwave component that is emitted by the sun (SW↓) and reflected from the ocean surface (SW↑ ...
  120. [120]
    Air-Sea Fluxes | Ocean Climate Stations - NOAA/PMEL
    Solar radiation (Qsw), net longwave radiation (Qlw), latent heat flux (Qlat), and sensible heat flux (Qsen) combine to form the net surface heat flux (Q0).
  121. [121]
    Wave-Coherent Air–Sea Heat Flux in - AMS Journals
    The coupled air–sea boundary layers play an important role in the fluxes of momentum, heat, and mass between the atmosphere and the ocean. These exchanges are ...
  122. [122]
    The role of ocean fluxes and radiative forcings in determining ...
    Aug 16, 2017 · The net heat flux into a column of atmosphere was calculated by summing the net radiative (TOA and surface), latent, and sensible heat fluxes ( ...
  123. [123]
    SP - The role of air–sea heat flux for marine heatwaves in ... - Reports
    Sep 30, 2024 · This study investigates the relative role of air–sea heat exchange and oceanic processes during the onset and decline phases of surface MHWs in the ...
  124. [124]
    Climate Change: Ocean Heat Content
    Adding up the rates of all layers, the full-depth ocean heat gain rate ranges from 0.66 to 0.74 Watts per square meter applied to Earth's entire surface. Less ...
  125. [125]
    Ocean Warming - Earth Indicator - NASA Science
    Sep 25, 2025 · Water has a high heat capacity, which means it can store a lot of heat. The atmosphere has warmed from increased greenhouse gases.
  126. [126]
    Ocean heat content in 2024 - Nature
    Apr 11, 2025 · Global full-depth ocean heat content (OHC) gain since 1960 reached a record 452 ± 77 ZJ in. 2024. OHC was 15 ± 9 ZJ higher than in 2023,.
  127. [127]
    Ocean Heat Content | CMEMS - Copernicus Marine Service
    Ocean Heat Content (OHC) is the surplus heat in Earth's system, mostly stored in the ocean, from the surface to 2000 meters, and is an indicator of Earth's ...
  128. [128]
    Real Freshwater Flux as a Natural Boundary Condition for the ...
    Freshwater flux used as a natural boundary condition for the salinity balance is applied to a primitive equation model of the oceanic general circulation.
  129. [129]
    A conceptual model of ocean freshwater flux derived from sea ...
    Sep 3, 2014 · A conceptual model is proposed to express freshwater flux (evaporation minus precipitation) as a function of sea surface salinity (and vice ...
  130. [130]
    NASA Earth Science: Water Cycle | Precipitation Education
    Evaporation ("E") controls the loss of fresh water and precipitation ("P") governs most of the gain of fresh water. Scientists monitor the relationship between ...
  131. [131]
    A global relationship between the ocean water cycle and near ...
    Oct 18, 2011 · [1] Ocean evaporation (E) and precipitation (P) are the fundamental components of the global water cycle. They are also the freshwater flux ...
  132. [132]
    Global Freshwater Fluxes into the World's Oceans (GRDC, 2021) - BfG
    The total freshwater input to the World's oceans between 90°N and 60°S is computed to 40181 km³/a. Excluding Greenland, the total is 39693 km³/a. Excluding ...
  133. [133]
    [PDF] Global air-sea heat and freshwater fluxes constrained by ocean ...
    Jan 13, 2025 · Mean freshwater transport and trends in both heat and freshwater transport are less consistent. Either a narrowing of the range of estimates ...
  134. [134]
    Arctic Ocean stratification set by sea level and freshwater inputs ...
    Aug 16, 2021 · This sequence suggests that precipitation and riverine freshwater fluxes control the stratification of the central Arctic Ocean. Based on these ...
  135. [135]
    [PDF] The influence of transient surface fluxes on North Atlantic ...
    Changes in net surface freshwater fluxes (precipitation, evaporation, and runoff from land) are found to be the dominant cause for the model's THC weakening.Missing: implications | Show results with:implications
  136. [136]
    Vertical redistribution of salt and layered changes in global ocean ...
    Aug 1, 2019 · Ocean salinity could, therefore, be useful in estimating the global net freshwater input to the ocean and understanding sea level changes.
  137. [137]
    Is There Robust Evidence for Freshwater-Driven AMOC Changes? A ...
    Jun 23, 2025 · In this paper we review (1) the paleoceanographic proxy data that have led to the widespread view that the AMOC sharply decreased for periods of several ...
  138. [138]
    Transport of Surface Freshwater from the Equatorial to the ...
    Apr 4, 2015 · Changes in surface freshwater con- tent in the tropical North Atlantic may also affect the ocean's thermohaline circulation through their ...
  139. [139]
    What is a Rossby wave? - NOAA's National Ocean Service
    Jun 16, 2024 · Rossby waves are huge, undulating movements of the ocean that stretch horizontally across the planet for hundreds of kilometers in a westward direction.
  140. [140]
    The Speed of Observed and Theoretical Long Extratropical ...
    Planetary or Rossby waves are the predominant way in which the ocean adjusts on long (year to decade) timescales. The motion of long planetary waves is westward ...
  141. [141]
    The Observed Dispersion Relationship for North Pacific Rossby ...
    We find a significant fraction of the energy present in the North Pacific Ocean is consistent with the linear Rossby wave dispersion relation Eq. (1). The ...Abstract · Introduction · A bit of theory and preliminary... · Results
  142. [142]
    [PDF] ROSSBY WAVES IN THE OCEAN (Part 1) - University of Washington
    Apr 5, 2010 · approximation, allowing useful solutions to be found. • The properties of Rossby waves (their dispersion relation) make them inherently ...
  143. [143]
    Global Observations of Oceanic Rossby Waves - Science
    Abstract. Rossby waves play a critical role in the transient adjustment of ocean circulation to changes in large-scale atmospheric forcing.
  144. [144]
    A Multi‐Layer Linear Rossby Wave Dispersion Relation for Vertical ...
    Nov 23, 2022 · The study aims at deriving a depth-dependent long Rossby wave dispersion relation to explain the different translation speed of mesoscale ...
  145. [145]
    Role of Ocean and Atmosphere Variability in Scale‐Dependent ...
    Jun 28, 2022 · This study investigates the importance of atmospheric processes (weather) and ocean currents in driving variations in sea surface temperature ( ...
  146. [146]
    Wind-Driven Mechanisms for the Variations of the Pacific Equatorial ...
    Jun 8, 2022 · However, a recent study found that a large-scale circulation associated with Rossby waves can also lead to EUC variation, implying that the ...<|control11|><|separator|>
  147. [147]
    Oceanic Mesoscale Eddies | Ocean-Land-Atmosphere Research
    Mar 28, 2025 · Oceanic mesoscale eddies typically span tens to hundreds of kilometers and last several weeks to months. They are a key dynamic process influencing the ...
  148. [148]
    Review of oceanic mesoscale processes in the North Pacific
    This review synthesizes current knowledge of mesoscale eddies and their impacts on the marine ecosystem across most of the North Pacific and its marginal Seas.
  149. [149]
    Global contributions of mesoscale dynamics to meridional heat ... - OS
    Aug 5, 2021 · Mesoscale ocean processes are prevalent in many parts of the global oceans and may contribute substantially to the meridional movement of heat.
  150. [150]
    Submesoscale Dynamics in the Upper Ocean - Annual Reviews
    Jan 19, 2023 · Submesoscales modify the density stratification in the upper ocean and redistribute energy between scales. These energy transfers are complex, ...
  151. [151]
    Trends and innovations in ocean mesoscale eddy studies ... - Frontiers
    Jul 23, 2025 · This bibliometric review investigates the trends and innovations in mesoscale eddy research using satellite altimetry data over the past decades ...
  152. [152]
    Large Mesoscale Eddy Properties and Dynamics in the Labrador ...
    Oceanic mesoscale eddies are crucial and active players in the lateral and vertical transport and mixing of heat, salt, and momentum in the Labrador Sea.
  153. [153]
    Submesoscale Processes and Dynamics - AGU Journals - Wiley
    Jan 1, 2008 · This chapter contains sections titled: Introduction Phenomenology Submesoscale Dynamics Implications Discussion.
  154. [154]
    [PDF] Submesoscale processes and dynamics
    Submesoscale pro- cesses make an important contribution to the vertical flux of mass, buoyancy, and trac- ers in the upper ocean. They flux potential vorticity ...
  155. [155]
    Observed regimes of submesoscale dynamics in the Southern ...
    Sep 25, 2025 · Submesoscale processes evolve rapidly, which necessitates high-resolution sampling in both space and time, making them difficult to observe.
  156. [156]
    S-MODE: The Sub-Mesoscale Ocean Dynamics Experiment in
    1)Quantitatively measure the three-dimensional structure of the submesoscale features responsible for vertical exchange. 2)Examine vertical transport processes ...
  157. [157]
    Mesoscale Meridional Heat Transport Inferred From Sea Surface ...
    Mar 7, 2024 · Rotating water masses in the ocean, termed as mesoscale eddies, are dynamically important features because they can transport nutrients and ...Abstract · Introduction · Mechanisms and... · Results: Meridional Heat...
  158. [158]
    review of ENSO theories | National Science Review - Oxford Academic
    Oct 10, 2018 · These four ENSO mechanisms are respectively called the delayed oscillator, the recharge–discharge oscillator, the western-Pacific oscillator and ...Abstract · SELF-SUSTAINED ENSO... · SUMMARY AND FUTURE...
  159. [159]
    The El Niño Southern Oscillation (ENSO) Recharge ... - AGU Journals
    Mar 20, 2025 · ENSO arises from air-sea interactions in the tropical Pacific, but influences climate and societies globally. In recent decades, progress in the ...
  160. [160]
    [PDF] El Niño and Southern Oscillation (ENSO): A Review - Staff
    Section 4.3 reviews the understand- ing of ENSO mechanisms. Section 4.4 summarizes the recent development of the central versus eastern Pacific ENSO events ...
  161. [161]
    Recharge Oscillator Mechanisms in Two Types of ENSO in: Journal ...
    The above results suggest that the recharge oscillator mechanism operates for the WP ENSO type. To further contrast the WP and CT ENSO life cycles, we sample ...
  162. [162]
    Interannual subsurface variability in the tropical Indian Ocean with a ...
    The dominant mode of the interannual variability in the subsurface TIO is characterized by a dipole. Its evolution is controlled by equatorial ocean dynamics ...
  163. [163]
    Oceanic Heat Content as a Predictor of the Indian Ocean Dipole
    Nov 21, 2022 · The Indian Ocean Dipole (IOD) is a critical mode of interannual variability in the tropical Indian Ocean, featuring a zonal dipole structure ...
  164. [164]
    Seasonality and Predictability of the Indian Ocean Dipole Mode
    May 25, 2021 · The Indian Ocean dipole (IOD) is an east–west dipole mode of interannual variability over the tropical Indian Ocean (Saji et al. 1999 ...
  165. [165]
    What Causes Tides - Tides and Water Levels
    In 1687, Sir Isaac Newton explained that ocean tides result from the gravitational attraction of the sun and moon on the oceans of the earth.
  166. [166]
    Tides - NASA Science
    The Moon's gravitational pull plays a huge role in the formation of tides. Tides are a cycle of small changes in the distribution of Earth's oceans.
  167. [167]
    Cause and Effect: Tides - National Geographic Education
    May 27, 2025 · For example, when an area covered by the ocean faces the moon, the moon's gravitational force on the water causes a high high tide. As the ...
  168. [168]
    Tides and Water Levels: NOAA's National Ocean Service Education
    Mixed Semidiurnal tide cycle (lower middle). An area has a mixed semidiurnal tidal cycle if it experiences two high and two low tides of different size every ...
  169. [169]
    11.3 Tide Classification – Introduction to Oceanography
    Mixed semidiurnal tides (or mixed tides), have two high tides and two low tides per day, but the heights of each tide differs; the two high tides are of ...
  170. [170]
    Tidal Currents - Currents: NOAA's National Ocean Service Education
    Tidal currents occur in conjunction with the rise and fall of the tide. The vertical motion of the tides near the shore causes the water to move horizontally, ...
  171. [171]
    Tide Patterns and Currents - University of Hawaii at Manoa
    A semidiurnal tidal cycle is characterized by two high tides daily of about equal heights occurring about 12 hours and 25 minutes apart (Fig. 6.16 A). The east ...
  172. [172]
    Activity: Modeling Amphidromic Points - University of Hawaii at Manoa
    Amphidromic points are places where the tidal range (difference in height between high and low tide) is zero compared with other areas in the basin.
  173. [173]
    11.7: Amphidromic Points and Co-tidal Lines - Geosciences LibreTexts
    Feb 14, 2021 · Amphidromic points are locations where there are little or no tide in the ocean. (This is also related to influence of continental land masses.<|separator|>
  174. [174]
    [PDF] Tidally driven mixing in a numerical model of the ocean general ...
    Astronomical data reveals that approximately 3.5 terawatts (TW) of tidal energy is dissipated in the ocean. Tidal models and satellite altimetry suggest that 1 ...
  175. [175]
    Tidal Energy Available for Deep Ocean Mixing
    Jan 1, 1999 · Our results show that 20-25% of tidal energy dissipation occurs outside of the shallow seas, the traditional sink for tidal energy. This ...
  176. [176]
    Dissipation of Tidal Energy - NASA Earth Observatory
    Jun 13, 2000 · “The observed pattern of deep-ocean dissipation is consistent with topographic scattering of tidal energy into internal motions within the water ...
  177. [177]
    Significant Dissipation of Tidal Energy in the Deep Ocean Inferred ...
    There has long been suggestive however, that tidal dissipation also occurs in the open ocean through the scattering by ocean-bottom topography of surface tides ...
  178. [178]
    [PDF] Surface gravity waves
    This limit is applicable to wind-waves in the open ocean. The dispersion relationship simplifies to. 2 = gk. Since the relationship between and k is not linear, ...
  179. [179]
    [PDF] 12.802 :DYHV LQ WKH 2FHDQ DQG $WPRVSKHUH Class Notes ...
    A course on wave motions for oceanographers and meteorologists has (at least) two purposes. The first is to discuss the important types of waves that occur ...
  180. [180]
    Surface Gravity Waves and Their Role in Ocean‐Atmosphere ...
    Jun 10, 2020 · They are comprised of locally generated wind seas and remotely generated swells. As indicated by its name, the wind supplies energy for the wind ...Abstract · Introduction · Effects of the Loop Current and... · Wave Effect on the...
  181. [181]
    Surface gravity waves (1/3) - OCN620
    Dispersion relations are fundamental wave concept! • Every type/class of wave has a unique dispersion relation that defines the relationship between ...
  182. [182]
    [PDF] Deep vs. Shallow Water Waves - UCLA
    Waves behave differently in “deep” water than they do in “shallow” water. Deep Water Waves: depth > 1/2 length: ocean bottom does not affect wave.
  183. [183]
    [PDF] Chapter 2 - Deep water gravity waves
    Figure 2.10: Scaled phase and group velocities for surface waves. In fact, both group and phase speeds are greatest and equal in the long wave (shallow water) ...
  184. [184]
    [PDF] PROCESSES - Benoit Cushman-Roisin
    Nov 30, 2012 · Gravity waves on the surface of water are one of the most visible manifestations of fluid motions and one with which we all have a certain ...
  185. [185]
    [PDF] Waves in the Ocean: Linear Treatment - Falk Feddersen
    Mar 4, 2019 · These lecture notes are designed for the Linear Ocean Waves course taught at Scripps. Institution of Oceanography, UCSD as part of the core ...
  186. [186]
    Surface Gravity - an overview | ScienceDirect Topics
    Surface gravity waves are most familiar form of wave motions occurring at the interface between the atmosphere and water, which we see on ponds, lakes, and ...Missing: fundamentals | Show results with:fundamentals
  187. [187]
    Probabilistic Tsunami Hazard Analysis: Multiple Sources and Global ...
    Nov 14, 2017 · This discussion presents a broad overview of PTHA, including (i) sources and mechanisms of tsunami generation, emphasizing the variety and ...
  188. [188]
    Tsunami Propagation Models Based on First Principles - IntechOpen
    Nov 21, 2012 · This chapter takes an alternative approach to the study of tsunami propagation in the open ocean. It commences with the theory of water wave propagation in ...
  189. [189]
    [PDF] Lesson 9: Waves - the NOAA Institutional Repository
    Wave Speed Formula​​ Speed of the tsunami (meters/second) is equal to the square root of g (the acceleration due to gravity, which a constant 9.81 meters/second) ...Missing: evidence | Show results with:evidence
  190. [190]
    The shallow water wave equation and tsunami propagation
    Mar 13, 2011 · The speed of propagation {v} of a tsunami can be approximated by the formula \displaystyle v \approx \sqrt{gb} \ \ \ \ \ (1) where {b} is the depth of the ...Missing: evidence | Show results with:evidence
  191. [191]
    Exploring the Tsunami Generation Potential of Major Faults in the ...
    Sep 1, 2025 · This study employs 3D hydrodynamic modelling using Delft3D to simulate tsunami generation, propagation, and inundation resulting from three ...
  192. [192]
    (PDF) Tsunami wave propagation along waveguides - ResearchGate
    Aug 9, 2025 · This is a study of tsunami wave propagation along the waveguide on a bottom ridge with flat sloping sides, using the wave rays method.
  193. [193]
    Tsunami modelling over global oceans | Royal Society Open Science
    Jan 15, 2025 · The major steps involved in the numerical simulation of the tsunami are generation and propagation. Generation includes generating initial ...
  194. [194]
    What Are Internal Waves in the Ocean, and Why Are They Important?
    May 25, 2022 · They have a vital role in transferring heat, energy, and momentum in the ocean. The turbulence they create when they break helps move nutrient- ...
  195. [195]
    Deep-ocean mixing driven by small-scale internal tides - Nature
    May 8, 2019 · In the deep ocean, tides supply much of the mechanical energy required to sustain mixing via the generation of internal waves, known as internal ...
  196. [196]
    Sensitivity of Internal-Tide Generation to Stratification and Its ...
    Being located on the irregular sea floor, the internal tide generation site is a function of both horizontal location and depth. So far, the internal tide ...
  197. [197]
    Interacting internal waves explain global patterns of interior ocean ...
    Aug 29, 2024 · Across the stable density stratification of the abyssal ocean, deep dense water is slowly propelled upward by sustained, though irregular, ...
  198. [198]
    Density stratification influences on generation of different modes ...
    Sep 22, 2014 · Density influence is characterized for internal tide generation and propagation The formation of ISWs is by nonlinear disintegration of ...
  199. [199]
    GLOBAL INTERNAL WAVES AND MIXING
    Major sources of internal waves are the tides flowing over topography (see Flow around Topography) and the wind blowing on the surface layer.
  200. [200]
    Nansen-bottle stations at the Woods Hole Oceanographic Institution
    Nansen-bottle stations were occupied by ships and personnel of the Woods Hole Oceanographic Institution from 1931 to about 1981.
  201. [201]
    [PDF] Instrumentation for Physical Oceanography: the last two decades ...
    The practice of making traditional wire cast hydrographic stations using Nansen bottles and reversing thermometers had largely disappeared by 1980 in favor ...
  202. [202]
    Planet Postcard: A Bucket Full of Data | News
    May 9, 2017 · Scientists and sailors were using actual wooden buckets to measure sea surface temperatures well into the mid-20th century.Missing: instruments | Show results with:instruments
  203. [203]
    Ocean Current Measurements: a Review | Hydro International
    Current measurement techniques have progressed from mechanical sensors to electromagnetic, acoustic and optical sensors. A broad distinction can be made ...Point Sensors · Em Sensors - Electromagnetic... · Doppler Profilers<|control11|><|separator|>
  204. [204]
    Topex-Poseidon - AVISO.altimetry.fr
    The Topex/Poseidon satellite was launched on 10 August 1992 with the objective of "observing and understanding the ocean circulation".
  205. [205]
    TOPEX, Jason series & SWOT | JPL Earth Science
    Aug 6, 2020 · TOPEX/Poseidon, Jason 1, 2, and 3 are altimetry missions. SWOT is a mission launching in 2021 to continue spaceborne altimetry.
  206. [206]
    SWOT (Surface Water Ocean Topography) - eoPortal
    As a swath-based Synthetic Aperture Radar (SAR) altimetry mission, following from the Jason-1, -2 and -3 missions with innovative technology, SWOT will perform ...
  207. [207]
    Two years after launch, exceptionally precise SWOT data ... - CNES
    Dec 17, 2024 · In the field of oceanography, SWOT is able to detect eddies ten times smaller than anything seen by previous satellite altimetry missions.
  208. [208]
    QuikSCAT Mission | PO.DAAC / JPL / NASA
    It's primary purpose is to provide all-weather ocean surface wind vector measurements over the ice-free global oceans. ... The reference height for the QuikSCAT ...
  209. [209]
    QuikSCAT: near sea-surface wind speed and direction
    The NASA scatterometer QuikSCAT operated from June 1999 until Nov 2009 and provided accurate satellite measurements of ocean surface wind speed and direction.<|separator|>
  210. [210]
    Description of ASCAT Data Products
    Aug 21, 2012 · ASCAT provides ocean surface wind vectors (50km and 25km resolution), storm data, global ambiguity, and ice data. The wind data is 10 meter ...
  211. [211]
    QuikScat / SeaWinds - Remote Sensing Systems
    QuikScat and SeaWinds are scatterometers that measure winds near the ocean surface using radar, also useful for some land and sea ice applications.
  212. [212]
    Sea Surface Temperature | NOAA CoastWatch
    The AVHRR FRAC SST data are produced from AVHRR/3s onboard Metop-A, -B and -C satellites using the NOAA Advanced Clear-Sky Processor for Ocean (ACSPO) v2.80 ...
  213. [213]
    avhrr - Advanced Very High Resolution Radiometer - NASA Earthdata
    Sep 30, 2025 · The Advanced Very High Resolution Radiometer (AVHRR) acquires measurements of land and sea surface temperature, cloud cover, snow and ice cover, soil moisture, ...
  214. [214]
    [PDF] Sea Surface Temperatures from MODIS
    MODIS derives SST using infrared measurements, five spectral bands, and a non-linear algorithm with brightness temperatures and a climatological estimate.
  215. [215]
    Multi-scale Ultra-high Resolution Sea Surface Temperature (MUR)
    MUR uses satellite and surface data to create high-resolution sea surface temperature maps, blending data from different sensors to fill data gaps.
  216. [216]
    Satellite Remote Sensing and the Marine Biodiversity Observation ...
    Nov 9, 2021 · Satellite remote sensing, in particular ocean color, can contribute to monitoring biological patterns and processes by providing information ...
  217. [217]
    The Argo Program: Two Decades of Ocean Observations
    Oct 14, 2020 · The Argo Program has collected, processed, and distributed over two million vertical profiles of temperature and salinity from the upper ocean in the past two ...Missing: capabilities | Show results with:capabilities
  218. [218]
    [PDF] The Argo Program
    The float program and its data management system began with regional arrays ... data coverage, quality, and timeliness of delivery. This achievement ...
  219. [219]
    [PDF] Observing the ocean with gliders - The Oceanography Society
    Ocean glider missions can provide a catalyst for student partici- pation and education in oceanography, as recently demonstrated by the trans-Atlantic Slocum ...
  220. [220]
    About - Argo floats
    Argo is an international program that measures water properties across the world's ocean using a fleet of robotic instruments that drift with the ocean currents ...Missing: coverage | Show results with:coverage
  221. [221]
    Data FAQ - Argo floats
    The temperatures in the Argo profiles are accurate to ± 0.002°C and pressures are accurate to ± 2.4dbar. For salinity, there are two answers.Missing: capabilities | Show results with:capabilities
  222. [222]
    Global phenomenon - Argo floats
    There is now a 15+ year time series of good global coverage, capturing decadal trends in ocean heat content and year to year variability. At the moment, the ...
  223. [223]
    Slocum Glider - Woods Hole Oceanographic Institution
    The Slocum Glider is a uniquely mobile network component capable of moving to specific locations and depths and occupying controlled spatial and temporal grids.
  224. [224]
    Seaglider - APL-UW
    These small, free-swimming vehicles can gather conductivity-temperature-depth (CTD) data from the ocean for months at a time and transmit it to shore in near- ...
  225. [225]
    [PDF] Underwater Gliders - UW School of Oceanography
    There are now three operational under- water gliders: Seaglider (Eriksen et al., 2001) built at the University of Washington,. Slocum Battery manufactured by ...
  226. [226]
    Argo and climate change
    Data from Argo allows for unprecedented spatial coverage of the world's oceans, with approximately one float per 3-degree box across the global ocean, where it ...Missing: capabilities | Show results with:capabilities
  227. [227]
    A revised ocean glider concept to realize Stommel's vision and ... - OS
    Mar 6, 2020 · This paper revisits Stommel's vision for a global glider network and the Argo design specification. A concept of floats with wings, ...
  228. [228]
    Challenges and Prospects in Ocean Circulation Models - Frontiers
    We revisit the challenges and prospects for ocean circulation models following Griffies et al. (2010). Over the past decade, ocean circulation models ...
  229. [229]
    [PDF] Opportunities and Challenges of High-Resolution Ocean Climate ...
    Jan 13, 2025 · In CM5, should we directly simulate the ocean mesoscale (i.e., eddy remove parameterizations), or should we parameterize it? Stephen Griffies, ...
  230. [230]
    Widespread global disparities between modelled and observed mid ...
    Apr 12, 2023 · Using comprehensive and quantitative methods, we present here the detailed assessment and validation of global ocean circulation model skills ...Results · Spatial Patterns Of... · Methods
  231. [231]
    Attribution of ocean temperature change to anthropogenic and ...
    Jul 31, 2019 · We examine whether significant changes in ocean temperatures can be detected in recent decades and if so whether they can be attributed to anthropogenic or ...
  232. [232]
    Multi-model attribution of upper-ocean temperature changes using ...
    Jun 1, 2016 · Can oceanic reanalyses be used to assess recent anthropogenic changes and low-frequency internal variability of upper ocean temperature?
  233. [233]
    Is the Atlantic Overturning Circulation Approaching a Tipping Point?
    Apr 10, 2024 · The Atlantic Meridional Overturning Circulation has a major impact on climate, not just in the northern Atlantic but globally.
  234. [234]
    Warning of a forthcoming collapse of the Atlantic meridional ... - Nature
    Jul 25, 2023 · Based on the CMIP5 generation of models, the AR6 IPCC report quotes a collapse in the 21st century to be very unlikely (medium confidence).
  235. [235]
    New study finds that critical ocean current has not declined in the ...
    Jan 15, 2025 · The Atlantic Meridional Overturning Circulation (AMOC) has not slowed down since the mid-20th century based on the North Atlantic air-sea heat ...Missing: attribution controversy
  236. [236]
    Stagnant North Atlantic Deep Water Heat Uptake With Reduced ...
    Jun 13, 2023 · This is of concern as recent studies suggest an ongoing AMOC slowdown during the 19th and 20th century as a response to increasing atmospheric ...<|separator|>
  237. [237]
    Historical Changes in Wind‐Driven Ocean Circulation Can ...
    Feb 22, 2023 · We find that changes to the wind-driven ocean circulation amplify global warming by 17% from 1979 to 2014.
  238. [238]
    [PDF] Internal Variability Dominates Over Externally Forced Ocean ...
    present evidence that recent Southern Ocean circulation changes may be dominated by internal variability ... trate that anthropogenic changes in ocean circulation ...
  239. [239]
    Satellite Observations of Ocean Circulation Changes Associated ...
    Oct 2, 2015 · Decades of satellite observations have greatly improved our understanding of large-scale ocean circulation changes associated with climate ...
  240. [240]
    Early detection of anthropogenic climate change signals in ... - Nature
    Feb 21, 2023 · For most variables, anthropogenic changes emerge earlier in the interior ocean than at the surface, due to the lower background variability at ...
  241. [241]
    What Causes Anthropogenic Ocean Warming to Emerge from ...
    Anthropogenic warming in the upper ocean water masses is generally driven by the uptake of excess heat passively stored by the ocean circulation.
  242. [242]
    Ocean Heat Content Responses to Changing Anthropogenic ...
    Jun 8, 2023 · We find that OHC is sensitive to aerosol forcing magnitude across much of the tropics and sub-tropics, and stronger negative forcing induces more ocean cooling.<|separator|>
  243. [243]
    An ocean perspective on CMIP6 climate model evaluations
    This paper serves as an introduction to the Deep-Sea Research II special issue entitled an ocean perspective on CMIP6 model evaluations.
  244. [244]
    Substantial Warming of the Atlantic Ocean in CMIP6 Models in
    CMIP6 models tend to produce stronger warming in the subpolar oceans of the Northern Hemisphere and weaker warming in the subtropical Atlantic and the Southern ...<|separator|>
  245. [245]
    Historical Ocean Heat Uptake in Two Pairs of CMIP6 Models
    Approximately 89% of the Earth energy imbalance between 1971 and 2018 is stored in the global ocean (von Schuckmann et al. 2020), raising its heat content.Models and data · Global ocean heat uptake · Regional ocean heat content...
  246. [246]
    Ocean Warming: From the Surface to the Deep in Observations and ...
    Dec 9, 2018 · Observations show that the ocean is warming from the near-surface through to the abyss, a conclusion that is strengthened with each new analysis.
  247. [247]
    AMOC Variability in Climate Models and Its Dependence on the ...
    Feb 9, 2025 · We find two main types of AMOC variations: one that occurs over decades to multiple decades, and another that happens over centuries.Missing: reliability | Show results with:reliability
  248. [248]
    Can we trust projections of AMOC weakening based on climate ...
    Oct 23, 2023 · In 2018, two independent studies were published in Nature that concluded that the AMOC had weakened by about 15% over the course of the ...
  249. [249]
    Continued Atlantic overturning circulation even under climate ...
    Feb 26, 2025 · Here we show that the AMOC is resilient to extreme greenhouse gas and North Atlantic freshwater forcings across 34 climate models.Missing: facts | Show results with:facts
  250. [250]
    Ocean circulation is key to understanding uncertainties in climate ...
    Jun 16, 2021 · Thirty state-of-the-art IPCC-climate models predict dramatically different climates for the Northern Hemisphere, especially Europe.
  251. [251]
    Data-driven global ocean modeling for seasonal to decadal prediction
    Aug 13, 2025 · In addition, the reliability of climate projections is also often hampered by dynamical model deficiencies (4). Encouragingly, a body of ...
  252. [252]
    Fifty Year Trends in Global Ocean Heat Content Traced to Surface ...
    Apr 11, 2021 · Using climate models (collectively called CMIP6) and ocean observations, we calculate the heat input into layers of water at the ocean ...
  253. [253]
    Quantification of The Performance of CMIP6 Models for Dynamic ...
    Mar 30, 2023 · Overall, the CMIP6 models reasonably reproduce the physical quantities of the driving variables and the warming ocean heat content and ...
  254. [254]
    Contributions of internal climate variability in driving global and ...
    Most of the warming in GST and OST since 1866 is found to be attributable to anthropogenic forcing, while the multidecadal and interannual GST and OST ...Contributions Of Internal... · 2. Data And Methodology · 4. Causes Of Global And...
  255. [255]
    Multi-model attribution of upper-ocean temperature changes using ...
    Jun 1, 2016 · The present study builds on previous work using two different indicators of upper-ocean temperature changes for the detection of both ...
  256. [256]
    Arctic amplification modulated by Atlantic Multidecadal Oscillation ...
    Apr 6, 2022 · This work determines the important influences of the AMO and anthropogenic greenhouse gas forcing on AA variations in the last millennium.
  257. [257]
    Simulated and observed variability in ocean temperature ... - PNAS
    Such detection and attribution studies have identified a large anthropogenic component in the observed changes and find that the “noise” of natural climate ...Abstract · Sign Up For Pnas Alerts · Variability On Different...
  258. [258]
    Natural variability has dominated Atlantic Meridional Overturning ...
    Apr 25, 2022 · ... natural variability has been dominant in AMOC changes; anthropogenic forcing is not yet reliably detectable by this method.
  259. [259]
    Heat Storage Pattern Linked to the Atlantic Meridional Overturning ...
    Jul 2, 2025 · Considering the pronounced natural variability, this weak anthropogenic AMOC slowdown probably has not emerged yet. Plain Language Summary.
  260. [260]
    On the association of Pacific Decadal Oscillation with the ...
    This study reports a salient contribution of the Pacific Decadal Oscillation (PDO) to the interdecadal variability of the APO during summer.
  261. [261]
    [PDF] Attribution of upper-ocean heat content changes using isothermal ...
    Changes of ocean heat content (OHC) play a critical role in understanding Earth's energy imbalance caused by human activities due to the large heat capacity ...
  262. [262]
    Climate impacts of a weakened Atlantic Meridional Overturning ...
    Jun 26, 2020 · While the Atlantic Meridional Overturning Circulation (AMOC) is projected to slow down under anthropogenic warming, the exact role of the ...
  263. [263]
    [PDF] Distinguishing the roles of natural and anthropogenically forced ...
    Capsule: In decadal forecasts, the magnitude of natural decadal variations may rival that of anthropogenically forced climate change on regional scales.Missing: peer- | Show results with:peer-