Fact-checked by Grok 2 weeks ago

Thermal equator

The thermal equator, also known as the heat equator, is a dynamic belt encircling the that connects the locations experiencing the highest mean annual surface air temperature at each , typically ranging from 25.85°C to 34.75°C with an overall mean of 27.75°C ± 1.3°C. Unlike the geographic at 0° , it deviates significantly due to hemispheric asymmetries in land distribution, currents, and , averaging around 5°N latitude and extending from as far south as 20°S in to as far north as 29.3°N over the . This irregular, meandering path—smoother over oceans but jagged over continents—reflects complex interactions between solar insolation, land-sea thermal contrasts, and atmospheric circulation patterns, such as the (ITCZ), which the thermal equator closely tracks. Seasonally, it shifts southward in (reaching up to 30°S) and northward in , driven by the Sun's apparent migration and greater landmass heating. Historically, the thermal equator has migrated in response to interhemispheric temperature gradients; for instance, it shifted northward around 14,600 years ago during the Bølling-Allerød warming period and southward during the (1300–1850 CE), influencing global precipitation belts and hydrologic cycles. In the context of ongoing from fossil fuel CO₂ emissions, projections indicate a further northward shift—potentially by several degrees—due to faster warming in the (e.g., 4.8°C versus 2.4°C in the under 3.6°C ), which could intensify monsoonal rains in and while exacerbating droughts in regions like the American West and .

Overview

Definition

The thermal equator, also known as the heat equator, is the circumglobal belt of locations on that experience the highest mean annual surface air at each . It represents a dynamic climatic feature that follows the latitudinal band where solar heating is most effectively maximized on average, distinct from fixed geographical lines. This belt is identified by connecting points along each where the annual average peaks, forming an irregular path that encircles the . Another equivalent view describes it as the line delineating the points of maximum average along each , emphasizing its role as a boundary of thermal maxima rather than a uniform contour. Mathematically, it corresponds to the φ where the first of mean T with respect to is zero (∂T/∂φ = 0) and the second is negative (∂²T/∂φ² < 0), confirming a local maximum in the meridional profile. Measurements of the thermal equator are based on surface air at 2 meters above ground level, a standard height for meteorological observations that reflects near-surface conditions influencing and ecosystems. These temperatures are averaged over climatic normals spanning at least 30 years to account for long-term variability and ensure representativeness, as established by international standards for data. For instance, datasets like CHELSA derive such means from downscaled atmospheric reanalyses, providing high-resolution grids suitable for defining the thermal equator's path.

Distinction from Geographic Equator

The geographic is defined as the imaginary on Earth's surface that lies in the plane perpendicular to the planet's axis of rotation, equidistant from the North and Poles at 0° . In contrast, the thermal equator represents a dynamic belt where the highest mean annual temperatures occur at each , determined by the balance of incoming solar radiation and outgoing rather than geometric fixedness. This conceptual distinction underscores that the geographic serves as a static reference for dividing the planet into hemispheres, while the thermal equator reflects variable climatic conditions influenced by global heat distribution. Positionally, the thermal equator deviates from the fixed 0° of the geographic , averaging approximately 5°N with a standard deviation of 11° across longitudes. This offset arises from hemispheric asymmetries in surface characteristics, positioning the thermal belt generally northward, though it can range from 20°S near to over 29°N in the . Observational evidence from global temperature datasets confirms this northward shift, with mean annual near-surface air temperatures peaking north of the geographic in most regions due to uneven heating patterns. For instance, analyses of 1981–2010 climate data show thermal equator temperatures averaging 27.75°C, ranging from 25.85°C to 34.75°C, distinctly higher than symmetric expectations around 0° . These differences have practical implications: the geographic equator is fundamental for , , and the latitude-longitude used in global positioning. Conversely, the thermal equator informs zoning, , and model validation, helping delineate tropical boundaries and predict heat-related environmental patterns.

Geographical Position

Mean Annual Location

The thermal equator is defined as the locus of points around the globe where the annual mean surface air reaches its maximum at each , representing the long-term of the zone of highest heat. Based on high-resolution global data, this belt is positioned on average at approximately 5°N , with a standard deviation of about 11° reflecting its latitudinal variability across longitudes. This northward shift from the geographic at 0° arises primarily from interhemispheric asymmetries driven by ocean circulation patterns. Analyses using reanalysis and datasets, such as CHELSA (derived from observations and models for the period 1981–2010) or similar products from ERA5 and NCEP/NCAR reanalyses, consistently depict the thermal equator as a wavy rather than a straight line, spanning from roughly 20°S to 29°N depending on the dataset and methodology. For instance, studies report mean positions ranging from 5°N to 8°N across different reanalysis products, with variations attributable to differences in and assimilated observations. These datasets typically compute the position by identifying the of peak annual mean (often 2-m air ) averaged over standard climatological normals, such as 1961–1990 or 1981–2010. Zonally, the thermal equator encircles the but exhibits notable deviations over continental landmasses, where higher heat capacities and lower push the maximum northward compared to regions. For example, it extends farther north over northern and , reaching up to 20°N in some sectors, while dipping southward over oceans like the Pacific. This undulating pattern underscores the influence of land-ocean contrasts on global temperature distribution, with the overall mean annual temperature along the belt averaging around 27.75°C.

Longitudinal Variations

The thermal equator displays pronounced longitudinal variations, with its position shifting northward over major continental landmasses compared to more consistent placements over oceans. Over , particularly around 40°E, the thermal equator extends to latitudes as high as 20°N, driven by regional surface heating patterns. In , it aligns closely with the continental margins, oscillating between 10°S and 10°N, including a southward excursion near 120°W influenced by coastal . Over the Maritime Continent (110°–145°E), the position varies markedly, forming a wide zone spanning approximately 10°N to 20°S, reaching northward over land areas and dipping to 20°S along the Australian coast. In oceanic regions, these shifts are subdued, resulting in a more stable thermal equator generally between 5° and 10°N. Across the Pacific and Atlantic basins, deviations remain minimal, maintaining the line near 5°–7.5°N in areas like the eastern Pacific and the Atlantic cold tongue vicinity. In the Indian Ocean, the thermal equator holds steady around 5°N. These patterns reflect the global annual mean position at 5°N, highlighting east-west irregularities. Mapping these longitudinal variations relies on geographic information systems (GIS) and satellite observations, such as temperature data from the MODIS instrument aboard NASA's and Aqua satellites, which enable precise delineation of maximum annual temperatures at fine spatial resolutions (e.g., 0.0083°). Recent analyses incorporate gridded datasets like CHELSA (1981–2010 ) processed via tools such as Python's RasterIO library to trace the thermal equator across all longitudes.

Formation and Causes

Solar Radiation Distribution

The distribution of incoming radiation, or insolation, exhibits a pronounced across Earth's latitudes, with the highest values occurring near the and progressively decreasing toward the poles. This pattern arises primarily because the experiences nearly perpendicular incidence of rays throughout the year, concentrating on a smaller surface area, whereas higher latitudes receive more oblique rays that spread the over larger areas. At the , the minimal impact of Earth's results in consistently high daily insolation, averaging around 400 W/m² annually, compared to less than W/m² at mid-latitudes and near zero during polar winters. The daily mean insolation at a given can be calculated using the following formula for top-of-atmosphere values on a surface: \overline{Q}^{\text{day}} = \frac{S_0}{\pi} \left( \frac{\overline{d}}{d} \right)^2 \left( h_0 \sin\phi \sin\delta + \cos\phi \cos\delta \sin h_0 \right) where S_0 is the (approximately 1366 W/m²), \overline{d}/d accounts for the variation in Earth-Sun distance due to , h_0 = \arccos(-\tan\phi \tan\delta) is the sunset in radians, \phi is the , and \delta is the solar declination angle (ranging from -23.45° to +23.45°). To obtain surface insolation, this value is multiplied by (1 - α), where α is the planetary albedo (typically 0.3 globally), though the latitudinal gradient in incoming radiation remains dominated by geometric factors rather than albedo variations. When integrated over an annual cycle, the total insolation peaks at the geographic (0° latitude), where the balance of day length and solar elevation yields the maximum input, approximately 35 /m² per day on . This equatorial maximum results from the symmetry of Earth's 23.5° axial obliquity, which causes the to migrate symmetrically between the and the , ensuring that low latitudes receive the most cumulative heating despite seasonal shifts. The obliquity introduces seasonal variations in insolation, with northern latitudes receiving more during the and southern latitudes during the , but the annual reinforces the poleward-decreasing gradient essential to the formation of the thermal equator. A minor exists due to Earth's elliptical , with perihelion in early slightly boosting southern hemispheric insolation by about 3-7% during its summer, though this does not shift the overall maximum from the .

Influence of Land and Ocean

The thermal equator's position and intensity are significantly modified by the differing physical properties of and surfaces, which alter how incoming is absorbed, stored, and redistributed. surfaces, characterized by lower compared to oceans, respond more rapidly to solar heating. The of is approximately 4.18 J/g°C, while that of typical soils is around 0.8–1.0 J/g°C, meaning requires about one-fourth the to raise its by the same amount as an equivalent mass of . This results in higher surface temperatures over continental regions near the equator, such as in and , where the thermal equator aligns more closely with landmasses rather than remaining fixed over open . Albedo, or the reflectivity of surfaces, further influences the net absorbed in equatorial regions, though its effects are secondary to differences. Open ocean surfaces have a low albedo of about 0.06, absorbing roughly 94% of incoming shortwave , while equatorial land covers like tropical rainforests exhibit albedos of 0.12–0.15, reflecting slightly more and absorbing 85–88%. This modest increase in reflectivity over vegetated land reduces absorption compared to adjacent oceans, but sparse or sandy equatorial soils can have albedos up to 0.20–0.35, further moderating heating in those areas. Overall, these albedo variations contribute to localized temperature maxima over land by creating differential heating patterns that the thermal equator tracks, particularly in regions like the . Ocean currents play a crucial role in redistributing heat along and across the equator, thereby shifting the thermal equator's latitudinal position. Warm currents, such as the North Equatorial Countercurrent flowing eastward between 5°N and 10°N in the Pacific and Atlantic, transport excess heat from western equatorial upwelling zones toward the east, effectively shifting the thermal maximum northward in those basins. Conversely, cold currents like the Humboldt Current along South America's western coast cool surface waters to as low as 25.85°C near 120°W, displacing the thermal equator from about 10°S over land to 10°N over ocean. These meridional heat transports by ocean currents thus prevent a uniform thermal peak directly at the geographic equator, with the thermal equator's average temperature varying from 25.85°C to 34.75°C longitudinally. Monsoon systems, driven by seasonal land-ocean thermal contrasts, amplify local heating in equatorial continental interiors by drawing in warm, moist air masses. Intense solar heating over land in spring and early summer establishes a strong temperature gradient with cooler oceans, promoting low-level convergence that pulls humid air from surrounding seas, as seen in the Asian and African monsoons. This influx enhances near-surface temperatures through adiabatic warming and reduced evaporative cooling in the boundary layer, sustaining elevated thermal conditions that anchor the thermal equator over land during peak seasons. For instance, the land-sea contrast exceeding 10–15°C in monsoon-prone regions like South Asia reinforces the thermal equator's northward migration in boreal summer.

Seasonal Dynamics

Migration with ITCZ

The thermal equator closely follows the seasonal migration of the (ITCZ), a band of low pressure where surface air converges and rises, driven by the sun's apparent movement across the . As the ITCZ shifts northward during the summer and southward during the summer, the locus of maximum surface temperatures—the thermal equator—tracks this path, reflecting enhanced heating in the region of active . This migration results in the thermal equator reaching its northernmost extent in , approximately 20–30°N over continental regions, while over oceans it remains closer to 10–15°N due to the moderating influence of . In contrast, it attains its southernmost position in January around 20–30°S, particularly over oceans and landmasses like and , before returning toward the . Annually, the thermal equator averages about 5°N, underscoring a slight northern influenced by greater land coverage in that hemisphere. The underlying mechanism involves the of northeast and southeast at the ITCZ, which funnels moist air equatorward and promotes intense vertical motion. This rising air enhances local surface heating through reduced and increased that traps outgoing radiation, while convective processes release , further amplifying temperature peaks along the thermal equator. The shift is primarily driven by changes in solar declination, with the ITCZ lagging slightly behind the due to thermal . Observational evidence from satellite missions, such as the Tropical Rainfall Measuring Mission (TRMM), demonstrates strong correlations between ITCZ precipitation maxima and thermal equator positions, with peak rainfall bands aligning closely with zones of elevated sea surface and land temperatures. For instance, TRMM data reveal that in , intense convective activity over the coincides with thermal peaks at 20–30°N, confirming the coupled dynamics of convergence and heating.

Hemispheric Asymmetries

The thermal equator displays a pronounced hemispheric , primarily manifesting as a consistent northward displacement from the geographic . This bias stems from the unequal distribution of land and ocean surfaces between the hemispheres, with the featuring approximately 39% land coverage compared to just 19% in the . Land's lower allows for more rapid and intense seasonal heating in the north during summer, elevating mean annual temperatures and pulling the locus of maximum warmth northward overall. Earth's orbital configuration introduces a countervailing , as perihelion occurs in early —aligning with summer and delivering about 7% more intense solar radiation to the south than during summer at aphelion. However, this eccentricity-driven effect proves insufficient to overcome the dominant land heating asymmetry, yielding negligible differences in annual mean insolation between hemispheres (less than 0.1 W m⁻²). Cloud cover disparities further exacerbate the northward shift, with the Southern Hemisphere's vast oceanic expanses fostering more persistent low- and mid-level clouds that enhance and suppress surface temperatures through increased reflection of incoming solar radiation. In contrast, the experiences relatively less cloud persistence over its land-dominated , permitting greater net absorption of heat. These combined factors result in the being warmer on average by about 1.24°C, reinforcing the thermal equator's bias. Quantitatively, the annual mean position of the thermal equator lies at approximately 5°N , with a variability of ±11° across longitudes; northern extensions can reach up to 29°N over regions like the , while southern excursions remain limited to a maximum of about 20°S, notably over . This asymmetry underscores the thermal equator's sensitivity to hemispheric contrasts in geography and atmosphere, rather than symmetric solar forcing alone.

Climatic and Environmental Impacts

Role in Atmospheric Circulation

The thermal equator functions as the primary heat source driving the meridional component of global , particularly through its role in the Hadley cells. Intense solar heating along this zone, which lies slightly north of the geographic due to hemispheric asymmetries in land-ocean , causes rapid ascent of air masses, forming a persistent low-pressure belt. This ascent creates surface convergence, drawing in moist air from surrounding regions and establishing the upward branch of the Hadley cells, which extend to about 30° latitude in both hemispheres. The resulting thermally direct circulation transports tropical heat poleward, mitigating the equator-to-pole essential for Earth's stability. Surface convergence at the thermal equator directly contributes to the formation of the , which constitute the lower branch of the Hadley cells. Air flowing equatorward from the subtropical high-pressure zones is deflected by the Coriolis effect, producing the northeast trades in the and southeast trades in the . These steady, low-level winds, typically blowing at 5–10 m/s, converge near the thermal equator, enhancing moisture transport and upward motion while helping to ventilate the . Zonal variations in the thermal equator's position, arising from longitudinal differences in sea surface temperatures and land influences, modulate the Walker circulation, an east-west overturning cell prominent over the equatorial Pacific. For instance, warmer waters in the western Pacific shift the thermal equator westward, strengthening easterly surface winds and ascent over the Maritime Continent while inducing descent over the eastern Pacific. This zonal asymmetry drives the Walker cell's intensity, with variations linked to phenomena like El Niño-Southern Oscillation, influencing global teleconnections. As a net energy surplus region, the thermal equator plays a pivotal role in the global balance of the atmosphere by exporting excess poleward. This export occurs primarily through the poleward of moist static , mainly via release from in the and midlatitudes, along with dry static , which are the dominant components of atmospheric to extratropical regions. Such mechanisms ensure by countering the imbalance where the absorb more solar radiation than they emit, sustaining the overall three-cell circulation model.

Effects on Precipitation and Weather Patterns

The thermal equator acts as a primary zone of atmospheric , where intense solar heating at the surface promotes deep and the ascent of moist air, resulting in a narrow band of heavy known as the . This convective activity generates frequent thunderstorms and showers, with annual rainfall totals often exceeding 2000–3000 mm in equatorial regions such as parts of the and . The rain belt's position closely tracks the thermal equator's mean annual location, typically within a few degrees, ensuring persistent wetness that supports lush tropical ecosystems and influences global moisture transport. The thermal equator's seasonal migration plays a pivotal role in driving monsoon systems across and , where its northward progression during boreal summer enhances land-sea thermal contrasts and pulls moist air inland. In , this shift triggers the monsoon, delivering critical rainfall for across the subcontinent, with the thermal equator's position serving as a predictor for monsoon onset and intensity. Similarly, in , the thermal equator's advance fuels the West African monsoon, shifting the rain belt northward and alleviating seasonal dryness in the . These dynamics underscore the thermal equator's influence on regional water cycles, as evidenced by paleoclimate records showing amplified monsoon precipitation during periods of northward thermal shifts. Tropical cyclone formation is preferentially supported in the 5°–20° latitude bands flanking the thermal equator, where convergence in the associated intertropical convergence zone (ITCZ) supplies low-level vorticity, warm sea surface temperatures, and abundant moisture essential for storm development. This positioning avoids the weak Coriolis force directly at the equator while leveraging the thermal equator's convective environment to initiate disturbances that evolve into hurricanes, typhoons, or cyclones. For instance, in the Atlantic and western North Pacific basins, a significant portion of tropical cyclone genesis occurs within this off-equatorial zone influenced by thermal equator dynamics. In contrast, areas poleward of the thermal equator, particularly the , endure drier conditions due to in the descending branches of the Hadley circulation cells, which suppress and promote clear skies and . This creates expansive arid zones, such as the deserts of the , , and , where annual can drop below 250 mm. The thermal equator's role as the Hadley cells' ascending boundary thus delineates a sharp gradient between tropical wetness and subtropical , amplifying weather pattern contrasts globally.

Historical and Scientific Context

Early Observations

Early observations of the thermal equator emerged from 19th-century scientific expeditions focused on global temperature patterns. Alexander von Humboldt's travels through from 1799 to 1804 yielded extensive temperature measurements, which he analyzed in his 1817 memoir "Des lignes isothermes et de la distribution de la chaleur sur le globe." These data formed the basis for the first global isotherm maps, revealing that the highest temperatures often occurred north of the geographic , influenced by continental landmasses absorbing and retaining heat more effectively than oceans. Humboldt's work demonstrated a latitudinal band of elevated warmth deviating from strict equatorial alignment, laying foundational insights into uneven heat distribution. Building on such measurements, mid-19th-century efforts by integrated temperature data into navigational charts. In his publication "The Physical Geography of the Sea," Maury compiled ship logs to produce thermal sheets mapping sea surface temperatures alongside winds and currents. These charts implied a dynamic band of maximum heat encircling the globe, shifted northward in many regions due to hemispheric asymmetries in land-ocean coverage, and varying seasonally with insolation. Maury's visualizations highlighted how this heat maximum drove atmospheric and patterns, advancing recognition of a non-geographic equatorial thermal zone. By the early 20th century, incorporated profiles into his system, first outlined in 1884 and refined through publications like "Die Klimate der Erde" in 1923. Köppen's framework identified "A" climates—tropical zones with all months averaging above 18°C—as peaking in near the , yet with northern extensions in continental interiors where annual maxima exceeded oceanic equatorial values. His classifications emphasized these thermal peaks as drivers of vegetation and weather regimes, underscoring the thermal equator's role in delineating global climatic boundaries without coinciding precisely with 0° . The explicit term "thermal equator" gained currency in meteorological literature during the , reflecting accumulated evidence on planetary heat budgets. It refers to the belt of maximum temperatures north of the geographic in the doldrums region, synthesizing prior isotherm and classification studies into a cohesive concept for . This nomenclature formalized observations of the thermal band's migratory nature and its offset due to land-ocean thermal contrasts.

Modern Mapping and Studies

Modern mapping of the thermal equator relies on high-resolution satellite observations and reanalysis datasets to track its position with greater precision than historical methods. Data from instruments like the Advanced Very High Resolution Radiometer (AVHRR), operational since the 1970s, provide long-term records of land surface temperatures and cloud cover that contribute to delineating thermal patterns. Similarly, the Clouds and the Earth's Radiant Energy System (CERES) on NASA's Terra satellite measures Earth's energy balance, including incoming solar and outgoing thermal radiation, enabling the identification of latitudinal zones of maximum net energy absorption since the late 1990s. These datasets are integrated into products like the CHELSA V2.1 climate dataset (1981–2010), which combines satellite-derived temperatures with reanalysis to map the thermal equator at 0.0083° resolution, revealing its mean position at approximately 5°N with deviations up to ±20° influenced by land-ocean contrasts. Reanalysis products such as ERA5 from the European Centre for Medium-Range Weather Forecasts further refine this tracking by assimilating global observations into a consistent spanning 1940 to present, allowing for the analysis of temporal variations in the thermal equator's location. ERA5 data indicate a subtle northward trend in energy transport anomalies across the , consistent with observed hemispheric asymmetries in recent decades. Building on early observations as a baseline, these tools enable quantitative assessment of the thermal equator's jagged path, such as northward bulges over continents like and . In climate modeling, the thermal equator serves as a key metric for validating general circulation models (GCMs) in IPCC assessments, including the Sixth Assessment Report (AR6, ), where simulations of equator-to-pole temperature gradients and energy fluxes are evaluated against observations to assess model performance in projecting global circulation. AR6 highlights how GCMs incorporate thermal equator dynamics to simulate hemispheric warming differences, with northern latitudes warming faster due to reduced and land feedbacks. Recent studies have advanced understanding through comparative analyses and geospatial applications. A 2024 study in the Bulletin of the examined the thermal equator on and Mars, using CHELSA data for to show its complex, continent-driven path (mean 27.75°C ± 1.3°C) versus Mars' simpler, southward-displaced track (5°–10°S), informed by thermal models; this underscores the of surface heterogeneity in thermal patterns. In 2025, researchers mapped the thermal equator using and CHELSA maximum data, employing zonal statistics to connect high- points at one-degree intervals, highlighting its utility in delineating ecosystems and validating models against observed jagged features like deviations over . Future projections under warming suggest a continued northward shift and potential widening of the thermal equator, driven by amplified northern hemispheric warming and altered energy transport, which could reshape global atmospheric patterns by 2100. GCM ensembles in AR6 project enhanced equator-to-pole gradients weakening under high-emission scenarios (SSP5-8.5), leading to expanded tropical zones and implications for circulation stability.

References

  1. [1]
  2. [2]
    Temperature Distribution on Earth & Heat Budget - PMF IAS
    The thermal equator lies to the south of geographical equator (because the Intertropical Convergence Zone or ITCZ has shifted southwards with the apparent ...Missing: definition | Show results with:definition<|control11|><|separator|>
  3. [3]
    Earth's Hottest Line: Mapping the Thermal Equator - Esri
    Feb 21, 2025 · The thermal equator is far from a smooth, continuous line; instead, it exhibits a jagged, meandering path that reflects the complexity of ...Missing: definition | Show results with:definition
  4. [4]
  5. [5]
    The Thermal Equator on Earth and Mars in - AMS Journals
    Jun 17, 2024 · The thermal equator (also known as the heat equator) is the circumplanetary set of points that represent the highest mean annual temperature at each longitude.
  6. [6]
    The Ascending Branch of the Hadley Cell | METEO 3 - Dutton Institute
    The thermal equator connects all the points that have the highest annual mean temperatures compared to other locations at their longitude.
  7. [7]
    THERMAL EQUATOR Definition & Meaning - Dictionary.com
    Thermal equator definition: an imaginary line round the earth running through the point on each meridian with the highest average temperature.Missing: characteristics sources
  8. [8]
    Equator - National Geographic Education
    Oct 19, 2023 · An equator is an imaginary line around the middle of a planet or other celestial body. It is halfway between the north pole and the south pole, ...
  9. [9]
  10. [10]
    Precipitation–Radiation–Circulation Feedback Processes ...
    Sep 10, 2020 · ... thermal equator (~5°–8°N). In the subtropical middle to lower troposphere, a pronounced RH deficit appears at the poleward flank of ...
  11. [11]
  12. [12]
    Climate and Earth's Energy Budget - NASA Earth Observatory
    Jan 14, 2009 · The climate's heat engine must not only redistribute solar heat from the equator toward the poles, but also from the Earth's surface and ...Missing: integration | Show results with:integration
  13. [13]
    Insolation | EARTH 103: Earth in the Future - Penn State
    On a yearly average, the equatorial region receives the most insolation, so we expect it to be the warmest, and indeed it is. Earlier, we mentioned the Solar ...
  14. [14]
    Insolation - The Climate Laboratory
    The seasonal dependence can be expressed in terms of the declination angle of the sun: the lat itude of the point on the surface of Earth directly under the sun ...
  15. [15]
    Heating Imbalances - Climate and Earth's Energy Budget
    Jan 14, 2009 · Graph of annual solar insolation versus latitude. The total energy received each day at the top of the atmosphere depends on latitude. The ...Missing: maximum | Show results with:maximum
  16. [16]
    Timing and significance of maximum and minimum equatorial ...
    Jan 31, 2008 · The timing of the maximum (and minimum) annual equatorial insolation may change around the equinoxes (solstices), alternating between the vernal and autumnal ...
  17. [17]
    Orbital eccentricity and Earth's seasonal cycle | PLOS Climate
    Jul 2, 2024 · The decrease in insolation absorbed by the Earth at aphelion (relative to the annual mean) is ~8 W/m2. This can be compared to the peak ...
  18. [18]
    Guest post: Why does land warm up faster than the oceans?
    Sep 1, 2020 · This is because land has a smaller “heat capacity” than water, which means it needs less heat to raise its temperature.
  19. [19]
    Albedo Values | MyNASAData
    Surfaces with low albedos include forests, the ocean, and some urban surfaces, such as asphalt. Albedo generally applies to visible light, although it may ...
  20. [20]
    Ocean Currents and Climate - National Geographic Education
    Jul 31, 2025 · Ocean currents, including the ocean conveyor belt, play a key role in determining how the ocean distributes heat energy throughout the planet.
  21. [21]
    [PDF] THE GLOBAL MONSOON SYSTEMS
    The fundamental driver of all the monsoon systems is solar heating of the land during the spring season that helps to establish a land-sea temperature ...
  22. [22]
    Inter-Tropical Convergence Zone - NOAA
    Jul 18, 2023 · The position of the ITCZ varies seasonally because it follows the Sun; it moves north in the Northern Hemisphere summer and south in the ...
  23. [23]
    Seasonal migration of ITCZ precipitation across the equator: Why ...
    Aug 15, 2003 · As the ITCZ migrates toward the equator during boreal winter, the CMT effect tends to weaken the Northern Hemisphere precipitation belt, produce ...Missing: thermal | Show results with:thermal
  24. [24]
    Observing the ITCZ with IMERG - NASA GPM
    Jul 17, 2019 · The ITCZ is a band near the Equator where trade winds converge, forming a rainfall band. IMERG uses satellite data to observe it, showing a ...
  25. [25]
    Long‐term characterization of the Pacific ITCZ using TRMM, GPCP ...
    Mar 21, 2016 · The climatological location of the ITCZ was found near 8°N, consistent with previous studies, with a preferred southern boundary location of 4°N ...Missing: position thermal
  26. [26]
    Interhemispheric effect of global geography on Earth's climate ... - CP
    Feb 26, 2019 · The climate response of the Earth to orbital forcing shows a distinct hemispheric asymmetry due to the unequal distribution of land in the ...
  27. [27]
    [PDF] Croll Revisited: Why is the Northern Hemisphere Warmer than the ...
    As land and ocean temperatures are vastly different due to contrasting thermal inertia. 139 and heat storage, it is useful to breakdown the inter-hemispheric ...
  28. [28]
    Milankovitch (Orbital) Cycles and Their Role in Earth's Climate
    Feb 27, 2020 · Currently perihelion occurs during winter in the Northern Hemisphere and in summer in the Southern Hemisphere. This makes Southern Hemisphere ...
  29. [29]
    Balancing act: why Earth's hemispheres reflect sunlight equally ...
    The scientists revealed that the Northern Hemisphere's aerosols and land reflect more sunlight, balanced by the Southern Hemisphere's low- and mid-level clouds.
  30. [30]
    Global Atmospheric Circulations - NOAA
    Oct 3, 2023 · Global Atmospheric Circulation is the movement of air around the planet. It explains how thermal energy and storm systems move over the Earth's surface.
  31. [31]
    Hadley Cells
    In this arrangement, heat from the equator generally sinks around 30° latitude where the Hadley Cells end. As a result, the warmest air does not reach the poles ...
  32. [32]
    Wind Systems | manoa.hawaii.edu/ExploringOurFluidEarth
    The Hadley cell distributes heat away from the equator, and the polar cell absorbs this heat. Hadley cell and polar cell circulation is straightforward as they ...Missing: thermal | Show results with:thermal
  33. [33]
    What Are Trade Winds? | NESDIS - NOAA
    Trade winds are winds that reliably blow east to west near the equator, about 30 degrees north and south of it, closer to the surface.
  34. [34]
    Global scale circulation - The Physical Environment
    Equatorward of the subtropical high, the pressure gradient between the high at 30o N and the low over the equator creates the northeast trade winds. In the ...<|control11|><|separator|>
  35. [35]
    [PDF] Changes in Zonal Surface Temperature Gradients and Walker ...
    Sep 1, 2011 · Variations in zonal surface temperature gradients and zonally asymmetric tropical overturning circulations (Walker circulations) are examined ...
  36. [36]
    The Walker Circulation: ENSO's atmospheric buddy - Climate
    Aug 1, 2014 · ENSO can affect this atmospheric circulation, the Walker Circulation, over the entire tropics, impacting rainfall near the equator across multiple continents.Missing: zonal | Show results with:zonal
  37. [37]
  38. [38]
    The Earth's Radiation Energy Balance
    During the summer the OLR is greater over land than the oceans, because the temperatures are warmer, while the albedo is greater over the oceanic regions where ...
  39. [39]
    Annual Migration of Tropical Rain Belt | NOAA Climate.gov
    May 4, 2011 · This daily cycle of heating, evaporation, and convection creates a persistent band of showers and storms around Earth's middle. This ...
  40. [40]
    Hydrologic impacts of past shifts of Earth's thermal equator ... - PNAS
    Sep 27, 2013 · The paleo-hydrologic record bears witness to past shifts in the position of the thermal equator that led to significant geographic alterations ...
  41. [41]
    Human-induced changes in the distribution of rainfall - Science
    May 31, 2017 · A second possibility is that Earth's thermal equator, around which the planet's rain belts and dry zones are organized, will migrate northward.
  42. [42]
    African Monsoon - NASA Earth Observatory
    Jun 16, 2004 · The first monsoon season occurs in late spring or early summer, centered about 5 degrees north of the equator, while the second monsoon arrives ...
  43. [43]
    Tropical Cyclone Introduction - NOAA
    Feb 24, 2025 · However, with only the rarest exceptions, these storms do not form within 5° latitude of the equator.
  44. [44]
    How do Hurricanes Form? | Precipitation Education - NASA GPM
    Tropical cyclones are like giant engines that use warm, moist air as fuel. That is why they form only over warm ocean waters near the equator.
  45. [45]
    The Sensitivity of Tropical Cyclone Activity to Off-Equatorial Thermal ...
    Changes in the latitude of the genesis region, and thus the absolute vorticity available to developing TCs, also influence the genesis frequency. Although ...
  46. [46]
    [PDF] Subtropical Drying: a robust response to Global Warming
    Stippled areas are where more than 90% of the models agree in the sign of the change. Precipitation increases very likely in high latitudes.
  47. [47]
    Transition to a more arid Southwest
    Descending air suppresses precipitation by drying the lower atmosphere so this process expands the subtropical dry zones.
  48. [48]
    ALEXANDER VON HUMBOLDT'S CLIMATOLOGICAL WRITINGS
    Jun 11, 2021 · In South America he was the first to describe the cold-water current of the west coast ('Humboldt current'). He also described and analysed ...
  49. [49]
  50. [50]
    Handbook of Meteorology/Local Winds - Wikisource
    Jul 9, 2021 · Handbook of Meteorology/Local Winds. Page · Source ... This calm belt lies north of the equator and practically covers the thermal equator.
  51. [51]
    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, ...
  52. [52]
    CERES – Clouds and the Earth's Radiant Energy System
    The CERES instruments provide direct measurements of reflected solar radiation and emission of thermal infrared radiation to space across all wavelengths ...CERES Data Products · CERES Operations · CERES Acronyms · Satellite MissionsMissing: equator AVHRR
  53. [53]
    Anomalous Northward Energy Transport due to Anthropogenic ...
    Sep 6, 2023 · We find a significant northward anomaly in the total atmospheric plus oceanic transport from about 10°S to 40°N that peaks in approximately 1975 ...
  54. [54]
    Chapter 4 | Climate Change 2021: The Physical Science Basis
    This chapter assesses simulations of future global climate change, spanning time horizons from the near term (2021–2040), mid-term (2041–2060), and long term ( ...Missing: validation | Show results with:validation
  55. [55]
    [PDF] Future global climate: scenario-based projections and near-term ...
    Jun 5, 2025 · IPCC AR6 WGI. Do Not Cite, Quote or Distribute. 4-2. Total pages: 163 ... equator-to-pole temperature gradient (ΔT2). Top. 38 and middle ...
  56. [56]
    Hydrologic impacts of past shifts of Earth's thermal equator offer ...
    As Earth warms in response to the continuing buildup of fossil fuel CO2, there will be a northward shift in the location of its thermal equator (1). This shift ...Missing: definition characteristics sources