Fact-checked by Grok 2 weeks ago

Atmospheric temperature

Atmospheric temperature refers to the measure of the average of the gas molecules comprising Earth's atmosphere, primarily and oxygen, which determines the thermal state of the air at any given location and altitude. This temperature varies widely, typically ranging from about 15°C (59°F) at to extremes exceeding 2,000°C (3,600°F) in the upper atmosphere, influenced by solar absorption, atmospheric , and dynamic processes like and . It is a fundamental parameter in and , essential for , climate modeling, and understanding phenomena such as and . The vertical profile of atmospheric temperature is characterized by distinct layers defined by gradients in thermal structure, as outlined in the (ISA) model. In the troposphere, the lowest layer extending from the Earth's surface to approximately 11-12 km (7-7.5 miles) at mid-latitudes, temperature decreases with altitude at an average environmental of 6.5°C per kilometer (3.57°F per 1,000 feet), dropping from a sea-level average of 15°C (59°F) to about -56.5°C (-70°F) at the ; this cooling results from decreasing air density and expansion of rising air parcels. Above this, the stratosphere (up to about 50 km or 31 miles) experiences a temperature inversion, warming from -60°C (-76°F) to around -15°C (5°F) due to the absorption of ultraviolet radiation by molecules. Further upward, the (50-85 km or 31-53 miles) sees temperatures plummet again to as low as -90°C (-130°F) at its upper boundary, owing to the thin air and lack of significant radiative heating. In the (above 85 km or 53 miles, extending to 600 km or 373 miles), temperatures rise sharply to over 2,000°C (3,600°F) from the absorption of high-energy solar ultraviolet and X-ray radiation by sparse atomic oxygen and other gases, though this heat is not felt as "hot" due to low density. Beyond this lies the , a transitional region where temperatures are not well-defined but molecules can escape into space. These variations are monitored using ground-based, airborne, and satellite instruments, with global datasets like those from the Microwave Sounding Units (MSU) and Advanced Microwave Sounding Units (AMSU) providing records since 1978 to track changes and anomalies. Atmospheric temperature influences nearly all atmospheric processes, from driving and storm formation in the to protecting life by filtering harmful in the . Human activities, such as , are altering surface and lower atmospheric temperatures, contributing to an observed global average increase of approximately 1.5°C (2.7°F) as of 2024 since the pre-industrial , with implications for ecosystems, sea levels, and . Understanding these dynamics is critical for , agricultural planning, and mitigating impacts.

Fundamentals

Definition and Importance

Atmospheric temperature refers to the measure of the average of air molecules, representing the thermal state of the atmosphere. It quantifies the associated with the random motion of gas particles in the air, where higher temperatures correspond to greater molecular speeds and energy. This temperature is typically expressed in degrees (°C) or (°F) for everyday use, or in (K) for scientific applications, with the Kelvin scale being absolute and starting at (0 K), the point where molecular motion theoretically ceases. The importance of atmospheric temperature lies in its fundamental role in shaping weather patterns, climate dynamics, and ecological systems. Temperature gradients drive by creating pressure differences that generate winds, as warmer air expands and rises while cooler air sinks and flows to replace it, powering global wind systems like and jet streams. In precipitation formation, temperature determines the condensation point of ; cooler air holds less moisture, leading to development and when saturated, while warmer conditions can enhance and intensify storms. For ecosystems, temperature influences , migration, and survival, with shifts altering and viability— for instance, warming trends have been linked to skewed sex ratios in offspring of temperature-sensitive species like sea turtles, potentially leading to population declines. Human activities are also profoundly affected: comfortable temperatures support outdoor work and daily life, but extremes strain energy demands for heating or cooling and pose health risks from stress. In , temperature impacts air density, which directly affects and performance, making accurate temperature data essential for safe takeoffs and flights, especially in hot conditions that reduce . Agriculture relies on optimal temperature ranges for crop growth; extremes, such as heatwaves or frosts, can reduce yields by stressing plants and disrupting , leading to significant productivity losses in staple crops like corn and wheat. The concept of atmospheric temperature has historical roots dating back to , where , in his work Meteorologica around 340 BCE, recognized temperature—embodied in the qualities of hot and cold—as one of the four elemental contraries influencing weather phenomena, alongside moist and dry, within a framework of the four classical elements (fire, air, water, earth). This early laid foundational ideas for understanding atmospheric processes, treating temperature variations as key drivers of meteorological events like winds and rains, though interpreted through a qualitative lens rather than modern quantitative .

Measurement Techniques

Atmospheric temperature measurements primarily rely on ground-based instruments that directly sense air temperature at specific locations. Traditional thermometers, such as mercury-in-glass and alcohol-in-glass types, have been used for over two centuries to record air temperature by thermal expansion of the liquid column, housed in protective shelters to minimize environmental influences. Modern ground-based systems employ digital sensors, including thermistors and resistance temperature detectors (RTDs), integrated into automated weather stations (AWS) for continuous, high-precision readings. To ensure accurate air temperature without bias from direct solar radiation or precipitation, these instruments are typically enclosed in standardized shelters like the Stevenson screen, a louvered wooden box that promotes natural ventilation while shielding the sensors. Psychrometers provide measurements of both dry-bulb (ambient air) and wet-bulb (evaporative cooling) temperatures, enabling the calculation of relative humidity and alongside direct data. These devices consist of two thermometers—one with a wet wick over its bulb—whose temperature difference reflects content in the air, often used in for comprehensive near-surface profiling. The evolution of these techniques traces back to the 17th century, when early thermoscopes—simple devices without scales—evolved into calibrated thermometers by inventors like and , marking the onset of systematic meteorological observations. By the , global networks like the Global Climate Observing System (GCOS) integrated AWS with electronic sensors, replacing manual readings with automated, real-time data collection across thousands of stations worldwide. Remote sensing techniques extend temperature observations to broader scales, particularly through satellite-based infrared radiometers that infer atmospheric and surface temperatures from emitted thermal radiation. Instruments like the Moderate Resolution Imaging Spectroradiometer (MODIS) aboard NASA's Aqua satellite measure brightness temperatures in specific infrared bands, applying the Stefan-Boltzmann law—where radiance is proportional to \sigma T^4 and \sigma is the Stefan-Boltzmann constant—to retrieve effective temperatures after accounting for emissivity and atmospheric effects. For vertical profiles, radiosondes—battery-powered instrument packages attached to weather balloons—ascend through the atmosphere up to approximately 30 km, transmitting real-time temperature, pressure, and humidity data via radio signals to ground stations. Accuracy in these measurements demands rigorous calibration and bias corrections to maintain reliability. Thermometers and sensors are calibrated against the International Temperature Scale of 1990 (ITS-90), which defines fixed points from the of equilibrium hydrogen (13.8033 K) to the freezing point of copper (1357.77 K) for thermodynamic consistency. effects, which can elevate recorded temperatures by up to 0.13°C in minimum readings due to impervious surfaces and anthropogenic heat, are corrected using homogenization algorithms that compare urban stations with rural references. For sea surface temperatures influencing lower atmospheric profiles, autonomous buoys like those in the array profile the upper ocean to 2,000 meters, providing near-surface data with accuracies of 0.002°C to complement air temperature observations.

Influencing Factors

Radiative and Thermodynamic Processes

The atmospheric temperature is fundamentally shaped by radiative processes involving the absorption and emission of , as well as thermodynamic exchanges that transfer heat within the system. Incoming solar , primarily in the shortwave spectrum, provides the primary energy input to the Earth-atmosphere system. A portion of this radiation is absorbed directly by atmospheric constituents: in the absorbs (UV) wavelengths below about 300 nm, heating that layer significantly, while in the absorbs near-infrared wavelengths, contributing to about 23% of total incoming solar energy absorption in the atmosphere. Additionally, clouds and aerosols scatter and absorb some shortwave radiation, further influencing the energy distribution. However, not all incoming reaches the surface or is absorbed; approximately 30% is reflected back to due to the planet's , which arises from high-reflectivity surfaces like , , and clouds, as well as atmospheric . This reflection, quantified as Earth's Bond of about 0.30, reduces the net energy available for heating and is a key factor in maintaining thermal balance. The remaining absorbed shortwave energy warms the surface and lower atmosphere, setting the stage for subsequent radiative and thermodynamic responses. The and its atmosphere emit longwave radiation as a counterbalance, approximating blackbody described by , which governs the spectral distribution of emitted energy peaking at wavelengths inversely proportional to ( as a derivative). For Earth's of around 255 K, this emission occurs mainly in the spectrum (5–50 μm). Greenhouse gases such as (CO₂) and (H₂O) absorb this outgoing longwave radiation and re-emit it in all directions, including downward, thereby trapping heat and elevating surface temperatures by about 33 K compared to a non-greenhouse scenario. This absorption-re-emission process, central to the natural , is particularly effective for H₂O in the and CO₂ across broader bands. Thermodynamic processes complement radiation by facilitating heat transfer near the surface and within the atmosphere. Sensible heat flux occurs through conduction from the warmer surface to adjacent air and subsequent convection, where heated air parcels rise, distributing thermal energy vertically; this flux is driven by temperature gradients and can reach tens of W/m² under clear-sky conditions. Latent heat, released during condensation of water vapor into cloud droplets, provides a significant warming mechanism, with the latent heat of vaporization for water at 0°C being approximately 2.5 × 10⁶ J/kg—equivalent to the energy needed to evaporate 1 kg of water, which is then liberated upon phase change. This release powers atmospheric dynamics and contributes to local temperature increases, especially in humid regions. The overall energy balance at the top of the atmosphere is expressed by the net radiative flux following the Stefan-Boltzmann law: Q = \sigma T^4, where \sigma = 5.67 \times 10^{-8} W/m²K⁴ is the Stefan-Boltzmann constant and T is the effective temperature in Kelvin; for Earth, incoming absorbed solar radiation (about 240 W/m²) balances outgoing longwave emission. These processes drive the basic diurnal cycle of atmospheric , with daytime net heating from dominant shortwave outpacing longwave emission, leading to warming, while nighttime exceeds any residual inputs, resulting in declines. This cycle establishes daily ranges of 5–15 K in the lower under typical conditions, modulated by local but fundamentally tied to the alternation of input.

Advection and Dynamic Influences

Advection refers to the horizontal transport of heat within the atmosphere by , which redistributes across regions and contributes to variations. This process occurs as air masses move from areas of differing s, such as when warmer air from lower latitudes displaces cooler air in higher latitudes, leading to localized warming, or vice versa for cooling effects. The advective can be expressed mathematically as -\rho c_p \mathbf{u} \cdot \nabla T, where \rho is air , c_p is the at constant pressure, \mathbf{u} is the wind , T is , and \nabla T is the ; this term represents the rate at which heat is transported horizontally per unit area. In practice, advection plays a key role in moderating mid-latitude temperatures by transporting warm tropical air masses northward, which can temper colder conditions, though cold air advection from polar regions often dominates winter cooling in these areas. For instance, persistent westerly winds in the mid-latitudes facilitate the poleward movement of heat, damping seasonal temperature extremes over land while amplifying them over oceans. Convection involves vertical mixing of air in unstable atmospheric layers, driven by forces that cause warmer, less dense air to rise and cooler air to sink, thereby transporting upward and influencing profiles. This is particularly active over heated surfaces, where solar radiation warms the ground and generates —rising parcels of buoyant air that initiate vertical motion and promote mixing in the . In moist environments, these thermals can develop into cumulonimbus clouds, where intense updrafts sustain deep , releasing that further destabilizes the atmosphere and leads to . Orographic effects arise when air flows over elevated , forcing ascent that results in adiabatic cooling as the air expands and loses without with its surroundings, often lowering temperatures on windward slopes and fostering formation. Similarly, frontal systems, such as fronts, advect cooler air masses into warmer regions, causing sharp temperature drops as the denser air undercuts and displaces the warmer air ahead, typically accompanied by gusty winds and . These dynamic influences highlight how air motion, rather than local heating alone, drives rapid temperature changes. The , a high-altitude fast-moving river of air, significantly modulates dynamics by influencing the containment of cold ; when the vortex weakens due to interactions with the meandering , frigid air can intrude southward, amplifying extremes. A notable example occurred during the January 2019 event in the United States, where a disrupted vortex allowed to plunge southward, resulting in record-low temperatures, including a national minimum of -48.9°C in the Midwest, driven by strong cold air advection.

Vertical Distribution

Temperature Versus Altitude

In the , the lowest layer of Earth's atmosphere extending from the surface to the at approximately 10-15 km altitude, generally decreases with increasing height at an average rate of 6.5°C per kilometer. This profile is exemplified by the U.S. Standard Atmosphere 1976 model, which defines a sea-level of 288.15 decreasing linearly to about 216.65 at the near 11 km. The decline results primarily from reduced pressure and adiabatic expansion of rising air parcels, with most phenomena confined to this layer due to its thermal gradient. Above the tropopause lies the stratosphere, spanning roughly 15-50 km, where temperature increases with altitude due to absorption of ultraviolet radiation by ozone molecules concentrated around 20-30 km. This warming reaches a maximum of approximately 270 K near the stratopause at 50 km, creating a stable inversion that limits vertical mixing. In the overlying (50-85 km), temperatures decrease sharply again, dropping to as low as -90°C at the due to minimal solar heating and radiative cooling by carbon dioxide. The , above 85 km and extending to about 600 km, experiences extreme heating from solar (EUV) radiation absorbed by atomic oxygen and , with temperatures rising to 500-2000 K or higher, though the low means this heat has little sensible impact. These vertical temperature profiles exhibit variations influenced by and season. The height is typically higher in equatorial regions (up to 17-18 ) compared to polar areas (around 8-10 ), reflecting stronger convective heating near the . Seasonally, the rises in the summer hemisphere due to enhanced solar insolation and , reaching maxima of about 16-18 in tropical summer conditions, while dipping lower in winter. Global average profiles are derived from composite datasets, including observations from the Integrated Global Radiosonde Archive (IGRA), which provide high-resolution vertical soundings from thousands of stations worldwide, and satellite microwave measurements from instruments like the Microwave Sounding Unit (MSU) for upper-air layers. These sources confirm the standard model's applicability, with observed zonal mean temperatures aligning closely to modeled values within 1-2 in the and .
Atmospheric LayerApproximate Altitude (km)Temperature Trend with AltitudeKey Temperature Range (K)
Troposphere0-15Decreasing288 to 217
Stratosphere15-50Increasing217 to 270
Mesosphere50-85Decreasing270 to 180
Thermosphere85-600Increasing180 to >2000
This table summarizes the U.S. Standard Atmosphere 1976 profile for mid-latitudes, illustrating the characteristic inversions and gradients.

Lapse Rates and Stability

The environmental (ELR) represents the observed rate of decrease in with increasing altitude in the atmosphere, typically expressed as -dT/dz, where T is and z is . Globally, the ELR in the is approximately 6.5°C per kilometer. In contrast, adiabatic lapse rates describe the theoretical cooling rates of a displaced air parcel under adiabatic conditions, where no heat is exchanged with the surroundings. For unsaturated (dry) air, the dry adiabatic lapse rate (DALR) is 9.8°C per kilometer, derived from the first law of for a hydrostatic atmosphere via . This value arises from the relation \Gamma_d = \frac{g}{c_p}, where g is the (approximately 9.8 m/s²) and c_p is the of dry air at constant pressure (about 1004 J/·). For saturated (moist) air, the moist adiabatic (MALR) is lower, typically ranging from 4°C to 6°C per kilometer, due to the release of during , which offsets some of the cooling from . A representative average value is about 5.5°C per kilometer. Atmospheric is assessed by comparing the ELR to these adiabatic lapse rates. If the ELR is less than the DALR (i.e., the environment cools more slowly than a dry rising parcel), the atmosphere is , as displaced air parcels tend to return to their . Conversely, if the ELR exceeds the DALR, the atmosphere is unstable, promoting buoyant ascent and vertical motion. For moist air, the situation is more nuanced: when the ELR lies between the DALR and MALR, the atmosphere exhibits conditional instability, where dry parcels are stable but saturated parcels can become buoyant upon reaching saturation. These comparisons are crucial for forecasting convective weather phenomena, such as , where steep ELR values exceeding the DALR allow air parcels to rise buoyantly, releasing and sustaining intense updrafts. Steep lapse rates in the lower , often exceeding 7°C per kilometer, signal high potential for severe development.

Temperature Inversions and Lifted Minimum

Temperature inversions represent deviations from the standard environmental , where temperature increases with altitude rather than decreasing, leading to stable atmospheric layers that inhibit vertical mixing. These anomalies can form through various mechanisms and have significant implications for , air quality, and forecasting. Inversions typically occur in the lower and can persist for hours to days, altering local climates and influencing phenomena such as and pollution dispersion. Several distinct types of temperature inversions exist, each driven by specific atmospheric processes. Surface inversions, also known as inversions, develop nocturnally when the Earth's surface loses through longwave , cooling the adjacent air layer more rapidly than the air aloft; these are common in clear, calm conditions and often dissipate with sunrise. inversions arise in anticyclonic high-pressure systems, where large-scale descending motion compresses and warms the air, creating a persistent warm cap typically 1–2 above the surface. Frontal inversions occur along weather fronts, particularly when warm air advances over a denser air mass, forming a sloped layer of increasing with . Inversions profoundly impact air quality by trapping aerosols, , and pollutants in the stable layer below, preventing their dispersion. A historical example is the incident in , where a prolonged surface inversion combined with industrial emissions from zinc and steel plants led to severe ; this event caused at least 20 deaths and affected over half the town's 14,000 residents with respiratory issues, highlighting the health risks of such conditions. Marine inversions provide another illustrative case, particularly along the coast, where subsidence from the Pacific system overlies cool, moist air advected from the cold ; this inversion, often within 300–500 m of the surface, caps vertical development of the marine layer, restricting formation to low-level stratus or stratocumulus decks that frequently advect onshore as , influencing coastal patterns. The lifted minimum temperature is a key diagnostic in upper-air soundings, such as skew-T log-P diagrams, for predicting low-level moisture-related phenomena like . It refers to the lowest temperature attained by a surface air parcel when traced upward: first dry adiabatically (at approximately 9.8 °C km⁻¹ cooling rate) until reaching the (LCL), where saturation occurs, and then moist adiabatically (at a reduced rate of about 6 °C km⁻¹ due to release) beyond. This parcel trajectory reveals potential cooling to at low altitudes, indicating fog risk if the LCL is near the surface and limits mixing; forecasters use it to assess whether or slight lifting could initiate without significant development. The LCL height, marking the transition in this process, is approximated by the z_{\text{LCL}} \approx 125 (T - T_d) where z_{\text{LCL}} is in meters, T is the surface air temperature in °C, and T_d is the dew point temperature in °C; this rule-of-thumb assumes standard pressure and provides a quick estimate for operational use.

Horizontal and Surface Variations

Near-Surface Air Temperature

The near-surface air temperature refers to the temperature of the air in the planetary boundary layer (PBL), the lowest portion of the atmosphere directly influenced by Earth's surface through friction, turbulence, and heat exchange. This layer typically extends from the surface to about 1-2 km in height, where turbulent mixing driven by surface friction and buoyancy forces homogenizes temperature and moisture profiles. During daytime conditions, the PBL often becomes superadiabatic, characterized by a strong due to intense solar heating of the surface, which promotes convective instability and vigorous vertical mixing of heat. At night, the layer tends to stabilize as at the surface creates a temperature inversion, reducing and limiting vertical . These dynamics result in significant heat redistribution near the surface, with the PBL acting as a between the rigid surface and the freer flow aloft. Local surface characteristics profoundly influence near-surface air temperatures within the PBL. heat capacity, which varies with moisture content and composition, determines how effectively the ground stores and releases heat; for instance, dry s with lower heat up and cool down more rapidly than wet soils, leading to amplified diurnal fluctuations in overlying air temperatures. contributes to cooling through , where plants release absorbed from the , reducing local air temperatures by 2-5°C in vegetated areas compared to bare surfaces, particularly in or arid settings. In coastal regions, sea breezes—circulations driven by land-sea temperature contrasts—moderate near-surface temperatures by advecting cooler marine air inland, often lowering daytime highs by several degrees. Standard meteorological observations of near-surface air are conducted at a height of approximately 2 meters above the ground, as recommended by the to represent human-scale conditions and minimize direct surface influences. Globally, the long-term average near-surface air is about 14°C, based on 20th-century baselines from extensive surface networks. Urban environments exhibit elevated near-surface temperatures compared to rural areas due to the effect, where heat-absorbing materials like store solar radiation and release it slowly, resulting in air temperatures 1-3°C warmer in city centers.

Diurnal, Seasonal, and Regional Ranges

The diurnal temperature range, representing the variation between daily maximum and minimum near-surface air temperatures, is influenced primarily by the daily cycle of heating and the region's capacity to retain or lose heat overnight. In arid desert environments like the , these ranges commonly span 10–20°C or more, with daytime highs often surpassing 50°C under intense insolation and nighttime lows approaching 0°C due to clear skies and low atmospheric moisture that facilitate rapid . In contrast, humid tropical regions exhibit much narrower ranges of around 5°C, as , high , and dampen the effects of heating and nocturnal cooling. Seasonal temperature ranges, the difference between winter and summer averages, are markedly larger in continental interiors than near the , owing to the thermal inertia of landmasses versus oceans. In regions like , annual ranges exceed 40°C, with winter lows reaching -50°C and summer highs around +30°C, reflecting the rapid seasonal response of large land areas to changing angles and limited moderation by surrounding bodies. At the , seasonal variations are minimal, often less than 5°C, due to consistently high elevation throughout the year; however, coastal locations experience a of 1–2 months in peak temperatures, attributable to the high of oceans that delays warming and cooling. Regional patterns in near-surface temperatures reveal a pronounced latitudinal , with annual means decreasing from approximately 25°C at the to -20°C at the poles, primarily driven by the variation in angle and incident across latitudes. Additionally, in topographically varied areas such as valleys and mountain bases, temperatures decline with at a near-surface of about 0.6°C per 100 m, resulting from the expansion and cooling of air ascending from lower altitudes. Notable extremes illustrate the breadth of these ranges: the highest reliably recorded near-surface air is 56.7°C, measured in Death Valley, California, on July 10, 1913. Conversely, the lowest is -89.2°C, observed at in on July 21, 1983.

Global and Climatic Perspectives

Global Mean Temperature

The global mean , often referred to as the global mean surface (GMST), represents the area-weighted average of near-surface air temperatures over land and sea surface temperatures (SSTs) over the oceans, computed from gridded observational datasets spanning the Earth's surface. This calculation involves interpolating measurements from weather stations, buoys, ships, and satellites onto a uniform grid (typically 5° × 5° or finer), then applying cosine weighting by latitude to account for varying grid cell areas, ensuring larger equatorial regions contribute proportionally more than polar ones. Datasets such as NOAA's NOAAGlobalTemp and the Met Office's HadCRUT5 perform this aggregation monthly and annually, providing estimates of absolute temperatures or anomalies relative to a reference period. A key reference baseline for many analyses is the 1951–1980 period, during which the global mean temperature averaged approximately 14°C, though estimates vary slightly across datasets due to differences in measurement homogenization and grid resolution. For context, the pre-industrial baseline (1850–1900) is estimated at about 13.7°C, establishing a for assessing long-term changes. These baselines play a central role in defining (WMO) normals, which are 30-year averages of meteorological variables used to characterize typical climatic conditions and detect deviations. The global is heavily influenced by its components, with covering about 70% of Earth's surface and thus dominating the ; SSTs are generally 1–2°C cooler than contemporaneous land air temperatures due to the ocean's higher and evaporative cooling, pulling the overall downward compared to a land-only . Hemispheric further shapes this value, as the —containing roughly 40% land versus less than 20% in the —is warmer by about 1–2°C on , driven by the greater thermal inertia of expanses and land's faster response to solar heating. Uncertainties in global mean temperature estimates arise primarily from sparse observational coverage in remote regions like the poles, where data gaps can results; for instance, amplification—faster warming at high latitudes—exacerbates under-sampling effects, leading to ranges of ±0.05–0.1°C in recent annual means. Modern datasets mitigate this through statistical infilling and reanalysis, but polar regions remain a key source of variability across products like HadCRUT5 and NOAAGlobalTemp. The historical record of atmospheric temperature reveals significant variations over millennia, reconstructed primarily through paleoclimate proxies such as tree rings, ice cores, and sediment records that capture signals from —orbital variations influencing solar insolation and glacial-interglacial transitions. During the Holocene warm period around 8000 BCE, global temperatures were approximately 1°C warmer than at the end of the last , marking a peak in warmth driven by these orbital forcings and retreating ice sheets. This period transitioned into cooler phases, with proxy data indicating a gradual decline until the pre-industrial era. Key historical events include the (approximately 900–1300 CE), during which regional temperatures in the North Atlantic and parts of were about 0.5°C above the subsequent centuries' averages, as evidenced by tree-ring and historical records showing expanded and reduced . This was followed by the (roughly 1300–1850 CE), a cooler interval with global temperatures roughly 0.5°C below pre-industrial levels, reconstructed from ice cores and proxies that document increased volcanic activity and solar minima contributing to alpine glacier advances and harsher winters. Instrumental measurements, beginning reliably around 1850 with surface station networks, provide direct evidence of these trends and the onset of modern warming. Since 1880, global surface have risen by more than 1.2°C as of 2025, with the rate accelerating to about 0.2°C per after the 1970s, as documented by comprehensive datasets from weather stations, ships, and satellites. As of 2025, the global mean surface for 2024 was approximately 1.55°C above the 1850–1900 pre-industrial average, confirming it as the warmest year in the instrumental record. This warming builds on the global mean baseline, intensifying from mid-century onward due to observed shifts in atmospheric composition. Future projections from climate models in the (AR6) outline scenarios based on (SSPs), estimating additional warming relative to pre-industrial levels. Under SSP1-2.6, a low-emissions pathway with rapid mitigation, global temperatures are projected to reach about 1.5°C by 2050, stabilizing near 1.8°C by 2100. In contrast, the high-emissions SSP5-8.5 scenario forecasts up to 4.4°C warming by 2100, reflecting continued reliance on fossil fuels and limited adaptation. These projections incorporate from greenhouse gases, including the well-established formula for CO₂: \Delta F = 5.35 \ln\left(\frac{C}{C_0}\right) \, \mathrm{W/m^2} where C is the current CO₂ concentration and C_0 is the pre-industrial value, derived from spectroscopic calculations and validated in atmospheric models.

Temperature Anomalies and Climate Forcing

Temperature anomalies represent deviations from a long-term average , typically calculated over a 30-year to account for natural variability and provide a stable reference for assessing changes. These anomalies are expressed as the between observed and the mean, with positive values indicating warmer-than-average conditions and negative values cooler conditions. For global surface air , the 1951–1980 is a commonly used by organizations like , allowing for consistent tracking of trends. In 2024, the global reached approximately +1.28°C above this , marking it as the warmest year on record. The first quarter of 2025 was the second-warmest on record, continuing the trend of exceptional warmth. Such anomalies are often amplified by internal climate modes like the (ENSO), which can drive short-term global spikes of +0.1 to +0.2°C during strong El Niño phases by altering ocean-atmosphere heat exchanges in the tropical Pacific. On longer timescales, anomalies are primarily shaped by , the imbalance in induced by changes in atmospheric composition or solar input. Anthropogenic greenhouse gases (GHGs), including CO₂, , and , have exerted a positive effective radiative forcing (ERF) of about +3.3 W/m² since 1750, with CO₂ alone contributing roughly +2.2 W/m², driving sustained warming. In contrast, natural forcings are smaller: solar variability contributes ±0.1 W/m² over an 11-year cycle, while volcanic eruptions impose temporary cooling of -0.5 to -2 W/m² through stratospheric aerosols that reflect for 1–3 years. Anthropogenic aerosols, such as sulfates from combustion, provide a counteracting negative ERF of approximately -1.3 W/m² by scattering and enhancing cloud reflectivity, partially masking GHG-induced warming. Internal variability, independent of external forcings, also modulates anomalies; for instance, the Atlantic Multidecadal Oscillation (AMO), a 60–80-year cycle in North Atlantic sea surface temperatures, has warmed the region by about 0.4°C during its positive phase over recent decades, influencing hemispheric patterns without altering the global energy balance. Detection and attribution studies, which compare observed changes to model simulations with and without human influences, conclude that virtually all observed since 1950—exceeding 1°C—is attributable to forcings, with natural factors contributing negligibly or even slightly cooling.

References

  1. [1]
    Chapter 1: Atmospheric Basics - UH Pressbooks
    Use mathematical formulas to define atmospheric temperature, pressure, and density; Compute pressure and density changes with altitude; Describe the vertical ...
  2. [2]
    Layers of the Atmosphere - NOAA
    Aug 20, 2024 · Heat is produced in the process of the formation of ozone, and this heat is responsible for temperature increases, from an average -60°F (-51°C) ...
  3. [3]
    Atmospheric Temperature - NASA Earthdata
    Sep 30, 2025 · Measuring atmospheric temperatures is essential for many forms of scientific work including weather forecasting and studying climate processes, ...
  4. [4]
    Atmospheric Properties & the ISA – Introduction to Aerospace Flight ...
    General Properties of the Atmosphere. The figure below illustrates the variations in atmospheric temperature, pressure, and density with altitude. Both pressure ...
  5. [5]
    Temperature - Dry Bulb/Web Bulb/Dew Point
    Temperature - the measure of average kinetic energy (KE) of a gas, liquid, or solid. KE is energy of motion. For air and other gases (in the troposphere only), ...<|control11|><|separator|>
  6. [6]
    Glossary - NOAA's National Weather Service
    TEMP)- The temperature is a measure of the internal energy that a substance contains. This is the most measured quantity in the atmosphere. Temperature ...
  7. [7]
    How Clouds Form | National Oceanic and Atmospheric Administration
    Jul 1, 2025 · Temperature's role​​ The presence of water-attracting nuclei alone is not enough for a cloud to form; the air temperature must also be below the ...
  8. [8]
    A Degree of Concern: Why Global Temperatures Matter
    Jun 19, 2019 · Our warming climate may be driving sea turtles into extinction by creating a shortage of males, according to several studies.
  9. [9]
    Climate Impacts on Society - US EPA
    Dec 22, 2016 · Higher temperatures and more extreme events will likely affect the cost of energy air and water quality, and human comfort and health in cities.Missing: aviation | Show results with:aviation
  10. [10]
    [PDF] Climate Change and US Agriculture: An Assessment of Effects - USDA
    Prolonged exposure to extreme temperatures will also further increase production costs and productivity losses associated with all animal products, e.g., meat, ...
  11. [11]
    Ancient and pre-Renaissance Contributors to Meteorology
    In Meteorologica, Aristotle considered four "contraries" (hot, cold, moist and dry) and four "elements" (fire, air, water and earth) and used them to explain ...
  12. [12]
    [PDF] On the reliability of the U.S. surface temperature record
    Jun 8, 2010 · LiG thermometers were generally housed in wooden Cotton Region Shelters (CRS; also known as Stevenson Screens) that were more easily located ...
  13. [13]
    The Elusive Absolute Surface Air Temperature (SAT) - NASA
    Mar 18, 2022 · This is the air temperature that would be measured 2 meters (around six feet) above the ground in a weather monitoring station with a Stevenson ...
  14. [14]
    [PDF] Guide to Instruments and Methods of Observation - WMO Library
    For example, the training needed to read air temperature in a Stevenson screen is at the lower end of the range of necessary skills, while theoretical and ...
  15. [15]
    [PDF] Evolution of temperature measurement – beginnings, progress and ...
    This paper charts the emergence of reliable temperature measurement beginning in the late 16th century with the thermoscope, leading to the emergence of true ...Missing: AWS GCOS
  16. [16]
    [PDF] MODIS Infrared Sea Surface Temperature Algorithm ... - NASA
    Apr 30, 1999 · This Algorithm Theoretical Basis Document (ATBD) describes our current working model of the algorithm for estimating bulk sea surface ...<|separator|>
  17. [17]
    Radiosonde Atmospheric Temperature Products for Assessing ...
    This product is recommended for assessing long-term changes in tropospheric and lower stratospheric temperatures on large spatial scales. For other uses, please ...Missing: profile | Show results with:profile
  18. [18]
    [PDF] International Temperature Scale of 1990 (ITS-90)
    The ITS-90 defines International Kelvin and Celsius temperatures, adopted in 1989, superseding older scales, and extending to the highest measurable  ...
  19. [19]
    [PDF] Quantifying the Effect of Urbanization on U.S. Historical Climatology ...
    Nov 15, 2012 · Essentially all of the bias was associated minimum temperatures in urban areas, which were about 0.13°C higher on average than rural areas; ...
  20. [20]
    Argo program
    The Argo program is an international program that uses temperature/salinity profiling float to measure the upper 2000 meters of the ocean.
  21. [21]
    Climate and Earth's Energy Budget - NASA Earth Observatory
    Jan 14, 2009 · About 23 percent of incoming solar energy is absorbed in the atmosphere by water vapor, dust, and ozone, and 48 percent passes through the ...Missing: UV | Show results with:UV
  22. [22]
    The Atmosphere: Earth's Security Blanket - NASA Science
    Oct 2, 2019 · Not only does it contain the oxygen we need to live, but it also protects us from harmful ultraviolet solar radiation. It creates the pressure ...
  23. [23]
    Measuring Earth's Albedo - NASA Earth Observatory
    Oct 20, 2014 · Using satellite measurements accumulated since the late 1970s, scientists estimate Earth's average albedo is about about 0.30.
  24. [24]
    Blackbody Radiation | EARTH 103: Earth in the Future - Penn State
    A blackbody(link is external) is an idealized material that absorbs perfectly all EM radiation that it receives (nothing is reflected), and it also releases or ...
  25. [25]
    The Earth's Radiation Budget - NASA Science
    Aug 4, 2023 · Greenhouse gases in the atmosphere (such as water vapor and carbon dioxide) absorb most of the Earth's emitted longwave infrared radiation, ...Missing: Planck's law blackbody
  26. [26]
    The Earth-Atmosphere Energy Balance - NOAA
    Jun 6, 2023 · The absorption of infrared radiation trying to escape from the Earth back to space is particularly important to the global energy balance.Missing: terrestrial Planck's blackbody CO2 re-<|control11|><|separator|>
  27. [27]
    3.3 Phase Diagram for Water Vapor: Clausius Clapeyron Equation
    where lv is the enthalpy of vaporization (often called the latent heat of vaporization, about 2.5 x 106 J kg–1), Rv is the gas constant for water vapor (461.5 J ...
  28. [28]
    [PDF] Lecture 3: Atmospheric Radiative Transfer and Climate - UCI ESS
    This relationship is called the Stefan-Boltzmann Law. E = σT. 4. E = radiation emitted in W/m2 σ= 5.67 x 10-8 W/m2 * K *sec. T = temperate (K). Page 7. ESS200.
  29. [29]
    [PDF] Diurnal cycle of upper-air temperature estimated from radiosondes
    May 3, 2005 · [2] Diurnal variations in upper-air temperature result from atmospheric absorption of solar and long-wave radi- ation, latent and sensible heat ...
  30. [30]
    [PDF] The seasonal cycle of atmospheric heating and temperature
    Effect of neglecting the kinetic energy on the energy fluxes in equation A1. The seasonal amplitude (the amplitude of the annual Fourier harmonic in phase with ...
  31. [31]
    [PDF] Lecture 2: Atmospheric Thermodynamics - UCI ESS
    atmosphere is related to the vertical structure of atmospheric temperature. ❑ To determine the static stability, we need to compare the lapse rate of the.
  32. [32]
    [PDF] Atmospheric Stability
    Around the cloud, the air is sinking. Cloud Development - Convection. • Convection usually occurs when the surface is heated and a surface parcel becomes warmer ...
  33. [33]
    [PDF] Marine weather of the inland waters of western Washington
    Thus, orographic lifting and cooling of these air masses moving inland readily result in widespread cloudiness and precipitation.
  34. [34]
    Basic Discussion on Pressure - National Weather Service
    The wind is blowing from north to south moving cooler air toward the warmer air. This is called "cold air advection", and is what usually occurs behind cold ...
  35. [35]
    Understanding the Arctic polar vortex
    ### Summary: Jet Stream Influence on Polar Vortex and Temperature Effects
  36. [36]
    2019 starts with extreme, high-impact weather
    The national low temperature record was measured at minus 56 °F (-48.9°C). The bitterly cold temperatures are caused by the influence of the Polar Vortex.
  37. [37]
    Basic Ozone Layer Science | US EPA
    Mar 5, 2025 · The stratosphere gets warmer at higher altitudes. In fact, this warming is caused by ozone absorbing ultraviolet radiation. Warm air remains in ...
  38. [38]
    The Mesosphere - UCAR Center for Science Education
    Temperature decreases with height throughout the mesosphere. The coldest temperatures in Earth's atmosphere, about -90° C (-130° F), are found near the top of ...Missing: profile | Show results with:profile
  39. [39]
    Variability and trends in the global tropopause estimated from ...
    Nov 3, 2006 · This study examines global tropopause variability on synoptic, monthly, seasonal, and multidecadal timescales using 1980–2004 radiosonde data.
  40. [40]
    Integrated Global Radiosonde Archive (IGRA)
    RATPAC. The Radiosonde Atmospheric Temperature Products for Assessing Climate (RATPAC) are a collection of radiosonde-based temperature anomaly time series for ...
  41. [41]
    Upper Air Temperature - Remote Sensing Systems
    RSS upper air temperature products are based on measurements made by microwave sounders. Microwave sounders are capable of retrieving vertical temperature ...
  42. [42]
    Comparing radiosonde and COSMIC atmospheric profile data to ...
    Dec 3, 2010 · Although on global average, raob and COSMIC-derived temperature profiles agree within 0.15 K, there are differences among sonde types, which ...
  43. [43]
    Chapter 5: Atmospheric Stability - UH Pressbooks
    An atmosphere where the environmental lapse rate is the same as the dry adiabatic lapse rate, meaning that the temperature in the environment also drops by 9.8 ...
  44. [44]
    [PDF] Dry Adiabatic Temperature Lapse Rate - atmo.arizona.edu
    The dry adiabatic lapse rate (defined as –. dT/dz) is about +9.8 K/km. The dry adiabatic temperature lapse rate is the temperature change with altitude when the.
  45. [45]
    [PDF] BASELINE CLIMATOLOGY OF SOUNDING DERIVED ...
    For reference, a typical moist adiabatic lapse rate is ~ 5.5°C km-1, the standard atmosphere lapse rate is ~ 6.5°C km-1 from 0-6 km AGL, while a dry adiabatic ...
  46. [46]
  47. [47]
    Temperatures - The Inversion
    An inversion is an increase of temperature with height. It can be created by high pressure, WAA, radiational cooling, warm air over cold water, cold fronts, or ...
  48. [48]
    The Deadly Donora Smog of 1948 Spurred Environmental ...
    Oct 26, 2018 · The 1948 Donora smog was the worst air pollution disaster in US history. It jumpstarted the fields of environmental and public health.
  49. [49]
    What are Marine Layer Clouds and How Do They Form?
    Another type of inversion often seen in California is called a radiation inversion. Radiation inversions form when the ground cools during the night resulting ...The Pacific High and Sinking Air · The Inversion Layer · Position of the LCL and...
  50. [50]
    Skew-T Parameters and Indices
    The Skew-T Log-P offers an almost instantaneous snapshot of the atmosphere from the surface to about the 100 millibar level.
  51. [51]
    [PDF] Planetary Boundary Layer Turbulence
    The PBL in the ocean and atmosphere is typically tens or hundreds of meters thick, respectively; this is thin compared to the 3-10 km depth scale of free ...
  52. [52]
    [PDF] The parametrization of the planetary boundary layer May 1992
    The planetary boundary layer (PBL) is the region of the atmosphere near the surface where the influence of the surface is felt through turbulent exchange of ...
  53. [53]
    Soil versus air temperatures: Understanding the relationship
    Feb 25, 2021 · Soil particles have a low heat capacity and water's heat capacity is relatively high. This mean that dry soils are more easily heated or cooled ...
  54. [54]
    Impact of Subsurface Temperature Variability on Surface Air ...
    The results show that allowing an interactive subsurface soil temperature does, indeed, significantly increase surface air temperature variability and memory ...
  55. [55]
    Tree effects on urban microclimate: Diurnal, seasonal, and climatic ...
    Results show that evapotranspiration of well-watered trees alone can decrease local 2 m air temperature at maximum by 3.1– 5.8 °C in the four climates during ...
  56. [56]
    Sea Breezes - VORTEX
    This circulation can lead to temperature drops of 8-11°C, clearer skies, increased humidity, and changes in wind speed and direction. While most commonly ...
  57. [57]
    Weather Station Siting: How to locate your weather station
    Suggested Measurement Heights and Exposure · 1.5 m ± 1 m (AASC) · 1.25 to 2.0 m (WMO) · 2.0 m for temperature only (EPA) · 2 m & 10 m for temperature difference ( ...
  58. [58]
    Monthly Climate Reports | Global Climate Report | October 2024
    The October global surface temperature was 1.32°C (2.38°F) above the 20th-century average of 14.0°C (57.2°F). This is 0.05°C (0.09°F) less than the record warm ...
  59. [59]
    Vegetation Essential for Limiting City Warming Effects
    Aug 25, 2015 · The urban heat island effect has long been observed to raise the temperature of big cities by 1 to 3°C (1.8 to 5.4°F), a rise that is due to ...
  60. [60]
    Desert: Mission: Biomes
    During the day, desert temperatures rise to an average of 38°C (a little over 100°F). At night, desert temperatures fall to an average of -3.9°C (about 25°F).
  61. [61]
  62. [62]
    Siberia climate: average weather, temperature, rain, when to go
    The average temperature ranges from -37 °C (-34.5 °F) in January to 19.5 °C (67 °F) in July. The weather station was already in operation at the end of the ...Chelyabinsk · Oymyakon · Yakutsk · Vladivostok
  63. [63]
    Introduction to the Basic Drivers of Climate | Learn Science at Scitable
    Coastal areas experience less seasonal variation in their climates than do continental interiors. Winters are milder, summers cooler, and the temperature ...
  64. [64]
    [PDF] University of Colorado and Black Swift Technologies RPAS-based ...
    The observations from LAPSE-RATE were aimed at improving our understanding of boundary layer structure, cloud and aerosol properties, and surface–atmosphere ...Missing: per | Show results with:per
  65. [65]
    World Meteorological Organization Assessment of the Purported ...
    The WMO assessment is that the highest recorded surface temperature of 56.7°C (134°F) was measured on 10 July 1913 at Greenland Ranch (Death Valley), ...
  66. [66]
    Eastern Hemisphere; Antarctica: Lowest Temperature
    The world record for low temperature was set at Vostok Station; Antarctica; on 21 July 1983. Cerveny et al. (2007) give this temperature as -89.4°C.
  67. [67]
    Climate at a Glance - National Centers for Environmental Information
    This tool provides near real-time analysis of monthly temperature and precipitation for the globe and is intended for the study of climate variability.Missing: 14° | Show results with:14°
  68. [68]
    HadCRUT5 - Met Office Hadley Centre observations datasets
    Sep 17, 2025 · HadCRUT5 is a gridded dataset of global historical surface temperature anomalies relative to a 1961-1990 reference period.Missing: NOAA | Show results with:NOAA
  69. [69]
    Climate change: global temperature
    May 29, 2025 · The global average surface temperature is an indicator of the state of Earth's energy balance: how much sunlight it absorbs minus how much heat ...
  70. [70]
    Updated 30-year reference period reflects changing climate
    Until the end of 2020, the most current and widely used standard reference period for calculating climate normals was the 30-year period 1981-2010.
  71. [71]
    Global Temperature Report for 2024 - Berkeley Earth
    Jan 10, 2025 · In Berkeley Earth's analysis the global mean temperature in 2023 is estimated to have been 1.62 ± 0.06 °C (2.91 ± 0.11 °F) above the average ...Missing: absolute | Show results with:absolute
  72. [72]
    Global land-ocean surface temperature data: HadCRUT5
    Dec 1, 2020 · HadCRUT5 is one of the main datasets used to monitor global and regional surface temperature variability and trends.Missing: calculation | Show results with:calculation
  73. [73]
    Paleoclimatology: Explaining the Evidence - NASA Earth Observatory
    May 9, 2006 · Coral, tree rings, and cave rocks record cycles of drought and rainfall. ... Evidence supporting Milankovitch's theory of the precise timing of ...
  74. [74]
    Climate Change and the Ice Ages - University of Michigan
    Figure 13 shows the correlation of the Milankovitch cycles with the paleoclimate records for the past million years. There is a good correlation, between ...
  75. [75]
    The Holocene temperature conundrum - PNAS
    Aug 11, 2014 · Here, we show that climate models simulate a robust global annual mean warming in the Holocene, mainly in response to rising CO 2 and the retreat of ice sheets.
  76. [76]
    Medieval Warm Period (MWP) | Research Starters - EBSCO
    The initial estimates by Lamb suggest MWP temperatures were about 0.5° Celsius warmer than those of the late twentieth century. This peak subsequently ...
  77. [77]
    The Little Ice Age & Global Warming - The Pigeon Roost
    Jan 23, 2018 · Global temperatures dipped about 0.5 C over a period of several centuries during the Little Ice Age. Since 1800, global temperatures have ...
  78. [78]
    Climate change widespread, rapid, and intensifying – IPCC
    Aug 9, 2021 · The report shows that emissions of greenhouse gases from human activities are responsible for approximately 1.1°C of warming since 1850-1900, ...
  79. [79]
    [PDF] Summary for Policymakers
    Only SSP1-2.6 and SSP5-8.5 are projected at 2300, as simulations that extend ... Coloured areas show the assessed very likely range of global surface temperature ...
  80. [80]
    IPCC AR6 Outlines Five Critical Future Scenarios | Anthesis Global
    Scenario 5: Avoid at All Costs (4.4°C by 2100) – SSP5-8.5. This worst-case ... The IPCC projects average global temperature to soar by 4.4°C by 2100.
  81. [81]
    [PDF] Radiative Forcing of Climate Change
    Radiative forcing, ∆F (Wm. −2). Constants. CO2. ∆F= α ln(C/C0). ∆F= α ln(C/C0) + β(√C − √C0). ∆F= α(g(C)–g(C0)) where g(C)= ln(1+1.2C+0.005C2+1.4 × 10. −6. C3).Missing: C_0) | Show results with:C_0)
  82. [82]
    Global Temperature Anomalies from 1880 to 2023 - NASA SVS
    Jan 12, 2024 · Global temperatures in 2023 were around 2 degrees Fahrenheit (1.1 degrees Celsius) above the average for NASA's baseline period (1951-1980), ...
  83. [83]
    The 2023 global warming spike was driven by the El Niño–Southern ...
    Oct 10, 2024 · Our results underscore the importance of the El Niño–Southern Oscillation in driving the occurrence of global warming spikes such as the one in 2023.
  84. [84]
    3.6.6 Atlantic Multi-decadal Oscillation - AR4 WGI Chapter 3
    North Atlantic SSTs show a 65 to 75 year variation (0.4°C range), with a warm phase during 1930 to 1960 and cool phases during 1905 to 1925 and 1970 to 1990.