Fact-checked by Grok 2 weeks ago

Cloud physics

Cloud physics is the study of the physical processes that lead to the formation, growth, and precipitation of atmospheric clouds, encompassing microscale interactions among water molecules, aerosols, and atmospheric dynamics. It focuses on the microphysical aspects of clouds, including the , , and coalescence of cloud droplets and ice crystals, which occur at microscales from molecular to centimeter levels. Cloud formation begins when moist air cools to its , typically through processes such as surface heating, orographic lifting, frontal lifting, or convergence, leading to where condenses onto (CCN) like or particles. These nuclei, present in concentrations of 10 to 10³ per cubic centimeter (or 10⁶ to 10⁹ per cubic meter), lower the supersaturation threshold required for droplet formation—around 0.1–1% for typical cloud droplets (about 10–20 μm in diameter)—compared to over 300% for homogeneous of pure water droplets due to the curvature effect, where smaller droplets have higher and tend to evaporate more readily. Growth of these droplets occurs primarily through diffusional growth via and subsequent collision-coalescence, enabling the development of precipitation-sized particles in clouds that reach sufficient vertical extent or contain phases. Clouds play a pivotal role in Earth's climate system by influencing the radiation budget, reflecting approximately 76 W m⁻² of incoming solar radiation while trapping outgoing infrared radiation, thereby modulating surface temperatures and energy exchanges. They also drive the hydrological cycle through , which delivers freshwater globally, and facilitate wet deposition of atmospheric pollutants while enabling chemical reactions in aqueous phases. Variations in cloud types—such as low-level stratus (0–2 km), mid-level altostratus (2–7 km), high-level (5–13 km), and vertically developed cumulonimbus—reflect diverse microphysical and dynamic regimes, impacting patterns, severe storms, and long-term feedbacks. Understanding these processes is essential for advancing models and simulations, where accurate representation of cloud microphysics remains a key challenge.

History

Early observations and theories

The earliest conceptual frameworks for understanding clouds emerged in , where philosophers sought to explain atmospheric phenomena through natural processes rather than . , in his work (circa 340 BCE), proposed that clouds formed from the condensation of "exhalations" rising from the Earth's surface, distinguishing between moist from water bodies that cooled into clouds and dry exhalations that contributed to other weather events like wind. This theory dominated Western thought for over two millennia, viewing clouds as a mixture of air and water particles suspended by atmospheric motion. In the 17th and 18th centuries, empirical observations began to refine these ideas, incorporating measurements of atmospheric variables. Evangelista Torricelli's invention of the mercury barometer in 1643 provided the first quantitative tool to measure air pressure, revealing variations that correlated with weather changes, including development in regions of lower pressure where air rises and cools. Building on this, conducted key experiments in the mid-18th century, demonstrating the roles of and in formation; in his 1784 essay "Meteorological Imaginations and Conjectures," he described how warm, moist air ascends, cools, and condenses into visible , releasing heat that sustains upward motion. These observations shifted focus from qualitative exhalations to dynamic processes involving temperature, pressure, and . The marked significant advancements in systematic classification and quantitative theory. In 1803, Luke Howard, an English and , introduced the first standardized cloud nomenclature in his "Essay on the Modification of Clouds," categorizing clouds into genera such as (wispy, high-altitude), cumulus (puffy, convective), and stratus (layered), based on their form, altitude, and associated weather— a system that remains foundational today. Concurrently, applied his to in works like "Meteorological Observations and Essays" (1793; 2nd ed. 1834), calculating vapor pressures and introducing the concept of the as the temperature at which air becomes saturated, enabling predictions of when —and thus cloud formation—would occur. Early laboratory experiments further illuminated the mechanisms of cloud formation, particularly the necessity of nucleation sites. In 1875, French physicist Pierre Jean Coulier conducted pioneering tests using a simple expansion chamber, showing that fog droplets formed readily in dusty air but not in filtered, dust-free air, thereby demonstrating that tiny particles act as nuclei for water vapor condensation. These findings laid the groundwork for understanding heterogeneous nucleation in natural clouds. By the late 19th century, such experiments had transitioned into broader investigations of aerosols, paving the way for 20th-century research on atmospheric particles.

Modern developments and key experiments

In the 1930s, significant theoretical advancements in cloud physics emerged with Tor Bergeron's ice crystal theory, which posited that precipitation in mixed-phase clouds primarily forms through the growth of ice crystals rather than liquid droplets alone. This idea built on Alfred Wegener's 1911 thermodynamic insights and was formalized in Bergeron's 1935 work, emphasizing the Bergeron-Findeisen process—also known as the Wegener-Bergeron-Findeisen mechanism—whereby ice crystals grow rapidly by vapor deposition in environments supersaturated with respect to ice but subsaturated with respect to liquid water. Walter Findeisen's 1938 experimental validations further refined this process, demonstrating how supercooled droplets evaporate to sustain ice growth, a mechanism central to understanding rain and snow formation in mid-latitude clouds. Post-World War II observations leveraging wartime radar and aircraft technologies propelled experimental cloud physics forward, culminating in Project Cirrus (1947–1952), a pioneering U.S. effort to modify clouds through seeding. Initiated by General Electric researchers Irving Langmuir and Vincent Schaefer, the project originated from 1940s studies on aircraft icing and precipitation static during Aleutian operations, using to track cloud dynamics and for in-situ sampling. Key experiments included the first seeding in 1946, which produced artificial snow, and 1947–1949 trials with in over and , yielding rainfall estimates up to 1.3 billion tons in single events and confirming efficacy for . These efforts established foundational techniques for and enhanced models of ice particle initiation. From the 1960s to , laboratory facilities like chambers and wind tunnels enabled controlled studies of cloud microphysics, simulating droplet collisions, , and formation under varied conditions. Vertical wind tunnels, such as those used by Pruppacher and colleagues, quantified raindrop terminal velocities and effects on coalescence, with seminal 1971 experiments revealing drag coefficients for realistic cloud environments. simulation chambers facilitated aerosol-cloud interaction tests, including nucleation thresholds, though many U.S. facilities were decommissioned by the in favor of field campaigns. NASA's Atmospheric Physics Laboratory (ACPL), flown on missions starting in 1980, provided microgravity insights into nucleation and crystal growth free from biases, with experiments demonstrating enhanced vapor deposition rates in zero-g conditions. The 2000s introduced space-based observations that revolutionized global cloud analysis, with the CloudSat and satellites, launched on April 28, 2006, as part of NASA's constellation. CloudSat's millimeter-wavelength profiled cloud vertical structure and , while 's measured layers and thin , revealing that multilayered clouds occur 60% of the time globally and correcting underestimations in passive data. Over a of joint data (2006–2016 and beyond), these missions quantified indirect effects, such as suppressed in polluted marine stratocumulus, and highlighted supercooled liquid layers in clouds critical for climate modeling. By 2025, ongoing analyses from these datasets informed studies on -cloud interactions, including a report showing that reduced East Asian decreased cloud droplet concentrations and reflectivity, amplifying regional warming by up to 0.2 W/m². Another 2025 investigation in Geoscience demonstrated high susceptibility of stratiform clouds to perturbations, with droplet number changes altering by 20–50% in high-latitude regions. These findings, aligned with IPCC AR7 scoping on , underscore aerosols' role in modulating cloud lifetime and efficiency amid declining emissions.

Fundamental principles

Atmospheric thermodynamics

Atmospheric thermodynamics underpins the behavior of air parcels in the atmosphere, dictating the conditions necessary for cloud initiation through and stability assessments. of thermodynamics, dU = \delta Q - p \, dV, describes the change in U of a moist air parcel as the sum of heat added \delta Q and work done by pressure p on volume dV. In cloud physics, this law is crucial because condensation releases , serving as a positive \delta Q term that warms the parcel, enhances , and promotes ascent leading to further cloud development. For unsaturated air undergoing adiabatic expansion or compression without phase changes, the dry adiabatic lapse rate governs the temperature decrease with height, derived from the first law combined with the hydrostatic equation and Poisson's relation for potential temperature. This rate is \Gamma_d = \frac{g}{c_{pd}} \approx 9.8 K/km, where g is gravitational acceleration and c_{pd} is the specific heat capacity of dry air at constant pressure. When saturation occurs, the moist adiabatic lapse rate applies, typically around 6 K/km, as latent heat release during condensation reduces the cooling rate compared to the dry case; this is obtained by modifying the first law to include moist enthalpy and follows from Poisson's equation adapted for reversible moist processes. Potential temperature \theta, defined as \theta = T \left( \frac{p_0}{p} \right)^{R/c_p} where T is , p_0 is reference pressure (usually 1000 hPa), p is actual , R is the for air, and c_p is specific heat at constant , remains conserved during adiabatic motions and indicates static when increasing with height. For moist air, the \theta_e extends this concept by incorporating the effects of , approximating \theta_e \approx \theta \exp\left( \frac{L_v r_s}{c_p T} \right) where L_v is of vaporization and r_s is saturation mixing ratio; it is conserved during pseudo-adiabatic ascent with removal and quantifies the total content available for in cloud-forming processes. The Clausius-Clapeyron equation relates saturation vapor pressure e_s to temperature via \frac{d \ln e_s}{dT} = \frac{L_v}{R_v T^2}, where R_v is the for , explaining the exponential increase in e_s with warming and thus the potential for when air cools adiabatically. This relation is essential for initiation, as it determines the vapor threshold beyond which becomes thermodynamically favorable in rising parcels.

Phase changes of water

Water exists in three primary phases—solid (), , and vapor (gas)—with transitions between these phases governed by and , as depicted in its . The illustrates the boundaries where each phase is stable, including the solid-liquid equilibrium curve (melting/freezing line), the liquid-vapor equilibrium curve (boiling/condensation line), and the solid-vapor equilibrium curve (/deposition line). A key feature is the , where all three phases coexist in equilibrium at precisely 0.01°C (273.16 K) and 611.657 Pa. At pressures below the , water can directly from solid to vapor without passing through the liquid phase, a process relevant to certain atmospheric conditions. Phase changes involve the absorption or release of , the required to alter the without changing . The of for , the to melt into at 0°C, is 334 kJ/kg. The of , the to convert to vapor at 0°C, is approximately 2.5 MJ/kg (2501 kJ/kg). For , the direct transition from to vapor at 0°C, the is about 2.83 MJ/kg, which is the sum of the heats of and under these conditions. These values highlight the significant transfers during phase changes, which influence atmospheric stability and cloud development through processes like evaporative cooling. The saturation vapor pressure, the maximum pressure of in equilibrium with its or at a given , follows distinct curves for and due to differences in molecular bonding. Over , the curve rises more steeply with than over , leading to with respect to ice in mixed-phase clouds at temperatures below 0°C. Approximations like the formula provide practical calculations for these curves; for , it is given by e_s(T) = 6.1094 \exp\left(\frac{17.625 T}{T + 243.04}\right) in , where T is in °C, valid from -45°C to 60°C with errors under 0.3%. For , modified coefficients yield e_s(T) = 6.112 \exp\left(\frac{22.46 T}{T + 272.62}\right) , accurate to within 0.1% over -50°C to 0°C. These formulations, derived from empirical fits to thermodynamic data, are widely used in atmospheric models to predict thresholds. In cloud droplets, the curvature of the liquid surface alters the saturation vapor pressure compared to a flat interface, an effect described by the Kelvin equation. For a spherical droplet of radius r, the vapor pressure e exceeds the flat-surface value e_s according to \ln\left(\frac{e}{e_s}\right) = \frac{2 \sigma M_w}{r \rho_w R T}, where \sigma is the surface tension (about 0.072 N/m for water at 20°C), M_w is the molar mass of water (0.018 kg/mol), \rho_w is the liquid density (1000 kg/m³), R is the gas constant (8.314 J/mol·K), and T is temperature in K. This increase, known as the Kelvin effect, makes small droplets (e.g., r < 1 \mum) harder to form or evaporate more readily, counteracting condensation in humid air.

Nucleation processes

Cloud condensation nuclei

Cloud condensation nuclei (CCN) are atmospheric aerosol particles that serve as the initial sites for the condensation of water vapor into liquid droplets during cloud formation. These particles lower the energy barrier required for droplet nucleation by providing surfaces upon which water molecules can condense, enabling the formation of cloud droplets at low supersaturations (RH slightly above 100%), much lower than required for homogeneous nucleation on pure water surfaces. CCN activation occurs when the ambient supersaturation exceeds a critical threshold specific to each particle's size, composition, and solubility, leading to the formation of a stable droplet embryo. The number and properties of CCN directly influence cloud droplet number concentration, size distribution, and optical properties, thereby affecting precipitation efficiency and . Common types of CCN include sea salt, sulfates, and dust particles, each characterized by varying degrees of solubility and hygroscopicity that determine their activation potential. Sea salt aerosols, primarily from ocean spray, are highly soluble and exhibit strong water uptake due to their ionic composition, making them effective CCN in marine environments. Sulfate particles, such as ammonium sulfate, are also highly hygroscopic, with the hygroscopicity parameter κ quantifying their ability to attract water; for ammonium sulfate, κ ≈ 0.61, reflecting moderate solubility compared to more ionic salts. Dust particles from arid regions, often mineral-based like silicates, show lower hygroscopicity (κ ≈ 0.01–0.5) unless coated with soluble material, but their large size can still facilitate activation under sufficient supersaturation. The hygroscopicity parameter κ, introduced as a volume-based mixing rule, simplifies predictions of particle-water interactions across compositions, with values ranging from near 0 for insoluble organics to 1.28 for sodium chloride as a proxy for highly soluble salts. Sources of CCN are divided into natural and anthropogenic categories, with global concentrations typically ranging from 10 to 1000 cm⁻³ in the boundary layer, varying by location and season. Natural sources dominate in remote areas and include sea salt from wave breaking, sulfates from volcanic eruptions and biogenic emissions (e.g., dimethyl sulfide from phytoplankton), and dust from desert winds, contributing baseline CCN levels of 10–100 cm⁻³ over oceans. Anthropogenic sources, such as pollution from fossil fuel combustion, biomass burning, and industrial processes, release sulfates, black carbon, and organics, elevating concentrations to 100–1000 cm⁻³ or higher in continental regions and increasing global mean CCN by 60–80% relative to preindustrial levels. These emissions alter the CCN budget, with anthropogenic contributions comprising up to 50% of total CCN in polluted marine boundary layers. The activation of CCN is governed by Köhler theory, which balances the Kelvin effect (increased vapor pressure over curved surfaces) and the Raoult effect (lowered vapor pressure due to soluble solutes), predicting a critical supersaturation s_c for each particle. In Köhler theory, the critical supersaturation s_c for activation is approximated by s_c = \sqrt{ \frac{4 A^3}{27 B} }, where A = \frac{2 M_w \sigma}{R T \rho_w} is the Kelvin term coefficient related to surface tension, and B = \frac{3 i m_s M_w}{4 \pi \rho_w r_d^3} is the Raoult term coefficient depending on the number of moles of solute i m_s and dry particle radius r_d; this highlights the interplay of surface tension (favoring smaller droplets) and solute effects (favoring larger ones), with a maximum supersaturation marking the activation point. Smaller particles require higher s_c due to dominant curvature effects, while soluble coatings reduce s_c by enhancing the solute term. The κ parameterization integrates into an analytical form of Köhler theory, allowing efficient computation of s_c from composition: S_c = \sqrt{ \frac{4 A^3}{27 B \kappa D_d^3} }, where A and B are constants related to surface tension and solute properties, and D_d is dry diameter. The activation spectrum describes the cumulative number of CCN that activate as a function of supersaturation, often parameterized as N(S) = C S^k, where C reflects total aerosol loading and k (typically 0.5–1) indicates spectrum steepness, influenced by particle size distribution and composition. Higher updraft velocities increase the maximum supersaturation achieved during ascent (via adiabatic cooling), broadening the activated fraction and shifting the spectrum toward smaller, less hygroscopic particles; for example, updrafts of 0.1–1 m s⁻¹ can double the number of activated CCN at low supersaturations (0.1–0.5%). In aerosol-limited regimes (low CCN concentrations), activation is highly sensitive to updraft fluctuations, while updraft-limited regimes (high concentrations) show reduced variability. This dependence underscores the role of vertical motion in determining effective cloud droplet number, with global models showing up to 60% variability in droplet concentrations attributable to updraft effects on CCN activation.

Ice nucleation

Ice nucleation in clouds primarily occurs through two mechanisms: homogeneous and heterogeneous freezing. Homogeneous freezing involves the spontaneous formation of ice crystals within supercooled liquid water droplets in the absence of solid inclusions, typically requiring temperatures around -38°C, which represents the homogeneous freezing limit for pure water. This process is significant in the upper troposphere where cirrus clouds form under strong supercooling conditions. Heterogeneous ice nucleation, in contrast, is facilitated by ice-nucleating particles (INPs) that lower the energy barrier for ice formation, allowing it to occur at warmer temperatures between 0°C and -38°C. INPs serve as templates for ice embryos and include a variety of atmospheric aerosols such as mineral dust from deserts, black carbon from combustion sources like biomass burning and aircraft exhaust, and biological particles including bacteria and fungal spores. Heterogeneous nucleation proceeds via specific modes: immersion freezing, where an INP is fully immersed within a supercooled droplet; contact freezing, involving collision between an interstitial INP and a droplet; and deposition nucleation, where water vapor directly deposits as ice onto an INP surface, particularly relevant under ice-subsaturated conditions. To model these processes in atmospheric simulations, empirical parameterizations are employed, such as the DeMott scheme for immersion freezing, which estimates the INP concentration as n_{INP} = \exp[ a (T - 273.15) + b ] \times N_{aer > 0.5 \mu m}, where N_{aer > 0.5 \mu m} is the concentration of particles larger than 0.5 μm diameter, T is in , and a ≈ -0.616, b ≈ 12.0 are fitted parameters derived from and field data. This scheme relates INP concentrations to total particles greater than 0.5 μm and , enabling predictions of ice formation rates in mixed-phase and clouds. In clouds, heterogeneous by INPs like mineral dust and competes with homogeneous freezing, influencing cloud optical properties and by altering number and size distributions. Similarly, in contrails, INPs promote rapid ice formation via deposition in the cold upper , potentially leading to persistent that enhances climate warming. Recent laboratory studies in 2025 have highlighted the efficiency of bacterial INPs, demonstrating their role in elevating INP concentrations in marine aerosols through correlations with specific microbial taxa like Formosa and Lewinella, which enhance ice formation at warmer temperatures and contribute to regional microphysics.

Cloud formation

Cooling mechanisms

Cooling mechanisms in cloud physics refer to the processes that reduce air , leading to and subsequent cloud formation. These mechanisms lower the of an air parcel until its reaches the , enabling . The primary pathways are adiabatic and non-adiabatic cooling, each driven by distinct physical principles. Adiabatic cooling occurs when an air parcel ascends and expands without heat exchange with its surroundings, resulting in a temperature decrease due to the work done against . This is the dominant mechanism for cloud formation, particularly through vertical motion in updrafts where the ascent velocity w = \frac{dz}{dt} determines the rate of expansion and cooling. In unsaturated air, the parcel cools at the dry adiabatic of approximately 9.8 K/km, but once is reached, release modifies the rate to the moist adiabatic . The moist adiabatic is given by \Gamma_m = g \frac{1 + \frac{L q}{c_p T}}{1 + \frac{L^2 q}{c_p R_v T^2}}, where g is gravitational acceleration, L is the latent heat of vaporization, q is the water vapor mixing ratio, c_p is the specific heat capacity at constant pressure for dry air, T is temperature, and R_v is the gas constant for water vapor. This rate typically ranges from 4 to 7 K/km, depending on temperature and humidity, and is less steep than the dry rate due to the warming effect of condensation. The height at which saturation occurs during ascent, known as the lifting condensation level (LCL), can be approximated using the dew point depression, the difference between air temperature T and dew point temperature T_d. The formula is z_{LCL} \approx 125 (T - T_d) meters, with T and T_d in degrees Celsius; for example, a 5°C depression yields an LCL at about 625 meters above the surface. Orographic lifting exemplifies adiabatic cooling when air is forced upward over terrain, such as mountains, leading to cloud formation on windward slopes. Similarly, frontal lifting occurs as warmer air rises over a denser cold air mass at atmospheric fronts, promoting widespread cloud development. Non-adiabatic cooling involves heat exchange with the and contributes to in horizontally stratified or slowly moving air. Radiative cooling primarily arises from longwave emission at cloud tops or clear skies, where outgoing infrared radiation exceeds incoming , cooling the air at rates up to several degrees per hour overnight. Mixing with drier air from above entrains unsaturated parcels into moist layers, diluting and causing evaporative cooling as excess moisture condenses or evaporates to equilibrate. of cloud droplets or into subsaturated air further cools the through the absorption of , enhancing instability in boundary layers. These processes often complement adiabatic mechanisms, particularly in stratiform clouds.

Moisture dynamics

Moisture dynamics in cloud physics refers to the processes by which is supplied to the atmosphere, leading to and formation primarily through the addition of moisture rather than temperature changes. Evaporative sources play a central role, with surface fluxes from oceans, lakes, and vegetated land providing the primary input of into the lower atmosphere. For instance, over tropical oceans, fluxes from evaporation can exceed 100 W/m², directly contributing to the moistening of air parcels that later form . from moist regions, such as warm currents or humid continental interiors, transports this vapor horizontally, enhancing local and facilitating development in otherwise drier areas. Relative humidity (RH), defined as RH = \frac{e}{e_s} \times 100\%, where e is the actual and e_s is the , serves as a key indicator of proximity to , with cloud formation occurring near or above 100% RH. Absolute humidity profiles, representing the total mass per unit volume, often show vertical gradients where lower-level enrichment from surface creates layers prone to upon uplift. In the , these profiles can exhibit sharp increases in moisture content near the surface, setting the stage for cloud initiation when combined with dynamic processes. Observations from radiosondes and satellites confirm that such profiles are critical for predicting heights, as higher absolute humidity aloft reduces the need for extensive cooling to reach . Isentropic uplift contributes to moisture dynamics by allowing moist air to ascend along surfaces of constant potential temperature without significant mixing, preserving high humidity levels during the process. This mechanism is particularly evident in synoptic-scale flows, where air parcels from low-level moist sources rise adiabatically, leading to gradual . In extratropical cyclones, — the horizontal influx of toward a low-pressure center—amplifies this effect, with vertically integrated often exceeding 10^{-7} kg kg^{-1} s^{-1} in active storm systems, directly fueling and formation. These convergent patterns draw in vapor from surrounding regions, elevating local and enabling widespread . In fog and boundary layer clouds, moisture dynamics dominate through nocturnal buildup, where reduced turbulent mixing overnight traps evaporated moisture near the surface, gradually increasing RH to saturation levels. Radiation fog, for example, forms through radiative cooling of the surface and near-surface air on clear, calm nights, which cools moist air to its dew point; evaporation contributes to the available moisture in the shallow layer, often in valleys or coastal areas. Stratus clouds in the boundary layer similarly arise from this persistent moisture supply, with advection sustaining the vapor field and preventing rapid dissipation. This process highlights how moisture addition alone can drive cloud formation in stable nocturnal environments.

Microphysical growth processes

Warm cloud processes

Warm cloud processes refer to the microphysical mechanisms governing the growth and precipitation formation within clouds where temperatures remain above 0°C, involving exclusively liquid water droplets. These processes dominate in low-level clouds such as stratocumulus and trade wind cumuli, where initial droplet activation occurs on (CCN) under supersaturated conditions. The primary mode of initial growth for cloud droplets is condensational growth, driven by the diffusion of toward the droplet surface when the ambient S exceeds unity. The radial growth rate is described by the approximate : \frac{dr}{dt} = \frac{S - 1}{r \cdot F_k}, where r is the droplet and F_k is a factor accounting for kinetic and thermal resistances to . This formulation highlights the inverse dependence on droplet , meaning smaller droplets grow faster relative to their size, leading to a narrowing of the size distribution over time in the absence of other effects. Growth rates typically range from 0.1 to 1 μm/ for droplets of 10 μm at of 0.1–1%, establishing the scale for subsequent processes. As droplets grow beyond approximately 20–30 μm, condensational growth becomes inefficient due to the reduced , and collision-coalescence emerges as the dominant mechanism for further enlargement and raindrop formation. In this process, droplets collide due to differential gravitational settling velocities, with larger droplets falling faster and collecting smaller ones. The collection efficiency E, which quantifies the fraction of collisions resulting in coalescence, varies from 0.1 to 1 depending on the () of the relative motion, typically in the range of 1–100 for cloud droplets. At low (<1), electrostatic and hydrodynamic effects enhance E, while at higher , turbulence can further increase collision rates by broadening relative velocities. This process is crucial for warm rain development, with coalescence rates scaling with the third moment of the droplet size distribution. Precipitation in warm clouds is characterized by droplet size distributions that evolve from narrow, quasi-monodisperse cloud droplet spectra to broader raindrop distributions. A widely used empirical form for raindrop sizes is the Marshall-Palmer distribution: N(D) = N_0 \exp(-\Lambda D), where N(D) is the number concentration per diameter interval D, N_0 = 8000 \, \mathrm{m^{-3} \, mm^{-1}} is the intercept parameter, and \Lambda \approx 4.1 R^{-0.21} \, \mathrm{mm^{-1}} depends on rainfall rate R in mm/h. This exponential form captures the decrease in larger drop concentrations observed in natural rainfall, with typical mean diameters of 1–2 mm for moderate rates. The onset of precipitation requires autoconversion, the initial transfer of mass from cloud droplets to raindrops via self-coalescence of similar-sized droplets, often parameterized with thresholds to represent this instability. In the seminal Kessler scheme, autoconversion activates when the cloud liquid water mixing ratio exceeds a threshold of approximately 0.3–1 g/kg, or equivalently when mean droplet diameters surpass 20–25 μm, reflecting the point where collision probabilities become significant. Alternative formulations, such as those extending , incorporate CCN concentration to modulate the threshold, with higher CCN leading to smaller droplets and delayed autoconversion. These thresholds ensure realistic warm rain formation times of 20–60 minutes in models.

Cold cloud processes

Cold cloud processes encompass the formation, growth, and precipitation mechanisms of ice particles in clouds at temperatures below 0°C, where water vapor and supercooled liquid droplets interact to produce ice crystals that dominate precipitation in many mid-latitude and polar regions. These processes are essential for understanding snowfall, hail formation, and cloud radiative properties, as ice crystals scatter and absorb radiation differently from liquid droplets. In subzero environments, ice development begins with nucleation and proceeds through diffusional growth, accretion, and collision, often leading to rapid glaciation of mixed-phase clouds. Supercooling can persist to thresholds around -38°C before homogeneous freezing dominates, but heterogeneous mechanisms typically initiate ice earlier. The Bergeron-Wegener-Findeisen (BWF) process is a primary mechanism driving ice crystal growth in cold clouds, where ice crystals grow rapidly by vapor diffusion at the expense of surrounding supercooled droplets due to the lower saturation vapor pressure over ice compared to liquid water below 0°C. This vapor density difference, Δρ_v = ρ_vl - ρ_vi (where ρ_vl is the saturation vapor density over liquid and ρ_vi over ice), creates a gradient that favors net deposition onto ice, often leading to cloud glaciation between -40°C and 0°C. Originally conceptualized by in 1911 and elaborated by in 1928 and in 1938 through cloud chamber experiments, the BWF process is crucial for precipitation initiation and is incorporated into numerical weather models to simulate weather and climate impacts. Ice crystal habits, such as plates forming between -8.1°C and -22.4°C or columns between -4.0°C and -8.1°C, influence growth rates and fallout speeds in cold clouds. Growth occurs primarily via vapor deposition, modeled using the capacitance approach, where the mass growth rate is given by \frac{dm}{dt} = 4\pi C D_v (\rho_v - \rho_{vs}) with m as crystal mass, C as the capacitance (dependent on size and habit, e.g., ≈ radius/2 for spheres but shape-specific for plates or columns), D_v as vapor diffusivity, \rho_v as ambient vapor density, and \rho_{vs} as saturation vapor density over the crystal surface. This model accounts for nonspherical shapes, with ventilation effects enhancing growth for falling crystals, as observed in supercooled cloud tunnel studies. Plates and dendrites promote broader size distributions, while columns favor vertical growth. Riming and aggregation further accelerate ice particle development into precipitation-sized elements, such as snowflakes or ice pellets. Riming involves supercooled droplets freezing upon collision with ice crystals, adding mass and altering shape; it is most efficient for crystals larger than 100-200 μm, with onset depending on crystal habit—columns rim along their length, while plates capture droplets on broad faces—leading to or hail in intense updrafts. Aggregation occurs through collisions between ice particles, forming branched structures that fall faster and collect more droplets, contributing to snow formation via accretion in stratiform clouds. These processes broaden particle size distributions and are key to precipitation efficiency, as simulated in cloud-resolving models. Ice multiplication in cold clouds arises from both homogeneous and heterogeneous pathways, amplifying initial nucleation to produce abundant crystals. Homogeneous multiplication, such as rime splintering during droplet freezing between -3°C and -8°C, generates fragments without external nuclei, while heterogeneous mechanisms include ice-ice collisions causing breakup, especially for rimed particles exceeding 1 mm. These processes explain observed high ice concentrations in mixed-phase clouds, with heterogeneous breakup dominating in turbulent conditions to form ice pellets. Competition between pathways affects cloud lifetime and precipitation. Recent airborne observations from HALO research flights during the 2021 CIRRUS-HL campaign, analyzed in 2025 studies, provide updated insights into cirrus microphysics over Europe. High-latitude cirrus showed lower ice water content (by 42%) and fewer but larger crystals compared to mid-latitudes, with frequent ice supersaturations of 105%-122% RH_i driving deposition growth. These findings highlight reduced aerosol influences post-COVID, refining models of ice habit evolution and radiative forcing in cold cirrus.

Cloud classification

Morphological systems

Morphological systems in cloud physics refer to the traditional classification of clouds based on their visual appearance, shape, and vertical position in the atmosphere, providing insights into underlying physical processes such as convection, stability, and phase changes. This approach emphasizes observable structures that reflect dynamic and thermodynamic conditions, distinguishing clouds by their form rather than quantitative measurements. The foundations of modern cloud morphology trace back to the early 19th century, when Luke Howard, a British pharmacist and meteorologist, introduced a systematic nomenclature in his 1803 essay "On the Modifications of Clouds." Howard identified three primary cloud forms—cirrus, cumulus, and stratus—drawing on Latin terms to describe their characteristic shapes: cirrus for fibrous or hair-like structures, cumulus for heaped or piled masses, and stratus for extended horizontal layers. His work established a binomial system that accounted for hybrid forms and transitional states, influencing subsequent international standards and enabling consistent observation of cloud evolution. The World Meteorological Organization (WMO) formalized this heritage in its International Cloud Atlas, recognizing ten principal genera based on shape and structure, with , , and serving as archetypal categories. Cirrus clouds appear as delicate, wispy filaments composed primarily of ice crystals, often indicating upper-level divergence and cold temperatures. Cumulus clouds exhibit heaped, dome-shaped outlines with flat bases, driven by buoyant updrafts that promote vertical growth and reflect daytime heating over land or water. Stratus clouds form uniform, blanket-like layers with horizontal extent, typically resulting from gentle lifting or radiative cooling in stable air masses. These genera capture essential morphological distinctions tied to atmospheric dynamics, such as instability in cumulus versus laminar flow in stratus. Within genera, species further refine morphological details; for instance, cumulus congestus describes a towering variant of cumulus with pronounced vertical development, resembling a cauliflower due to accelerated condensation and precipitation formation within vigorous updrafts. Such species highlight transitional morphologies that signal intensifying convection. Microphysical parameters, like droplet concentration, subtly influence these shapes by affecting growth rates and optical properties. Clouds are also categorized by altitude levels, which correlate with temperature regimes and moisture availability: high-level clouds (e.g., cirrus) form above 6 km in temperate latitudes, middle-level clouds (e.g., altocumulus) between 2 and 7 km, and low-level clouds (e.g., stratus) below 2 km. These divisions reflect vertical stratification in the troposphere, where decreasing temperatures favor ice-phase clouds at higher altitudes. Supplementary features augment primary morphologies, providing additional clues to local instabilities or environmental conditions. Mammatus, or mamma, consists of pouch-like protrusions hanging from cloud undersides, often on cumulonimbus anvils, arising from sinking pockets of cold air in turbulent outflows. Virga appears as pendant streaks of precipitation evaporating before reaching the ground, common in arid regions and indicating subsaturated lower layers. Contrails, anthropogenic linear clouds from aircraft exhaust, persist as cirrus-like forms (Cirrus homogenitus) when ice crystals form in humid upper air, demonstrating human influence on natural cloud morphology.

Physical parameter classifications

Cloud physics employs physical parameter classifications to categorize clouds based on measurable attributes such as thermodynamic phase, precipitation characteristics, and dynamic processes, providing insights into their formation, evolution, and radiative impacts. These classifications differ from morphological ones by emphasizing quantifiable properties like temperature thresholds, particle composition, and vertical motion rather than visual appearance. Phase-based classifications distinguish clouds by the dominant form of water substance: liquid, ice, or a mixture. Warm clouds, typically found in temperatures above 0°C, consist primarily of liquid water droplets and are prevalent in tropical and subtropical regions. Cold clouds, occurring below 0°C, are dominated by ice crystals, often forming through deposition or freezing processes in high-altitude or winter stratiform layers. Mixed-phase clouds, which contain both supercooled liquid droplets and ice particles, are critical in mid-latitude weather systems; for instance, at temperatures between -20°C and 0°C often exhibit this phase partitioning, where ice growth via the depletes surrounding liquid. These phase distinctions influence cloud lifetime and precipitation efficiency, with mixed-phase clouds particularly sensitive to entrainment mixing that alters liquid-ice mass ratios. Precipitation potential classifies clouds by the intensity, duration, and type of fallout they produce. Nimbostratus clouds generate steady, widespread precipitation such as prolonged rain or snow due to their layered structure and slow ascent rates, often covering large areas with uniform drizzle. In contrast, cumulonimbus clouds produce intense, short-duration showers or hail through vigorous updrafts that support rapid droplet coalescence and riming. Virga, a phenomenon where precipitation evaporates in subsaturated subcloud layers, is observed in both types but is more common in arid environments or altocumulus decks, reducing surface rainfall despite active hydrometeor formation aloft. These distinctions highlight how microphysical pathways—coalescence in warm sectors versus aggregation in cold—affect precipitation reach and volume. Dynamic classifications focus on motion and stability, dividing clouds into convective and stratiform regimes. Convective clouds feature strong vertical turbulence driven by buoyancy, leading to towering structures with high updrafts exceeding 10 m/s, as in thunderstorms where instability fosters rapid energy release. Stratiform clouds, conversely, exhibit gentle horizontal flows and subsidence, forming extensive layers with weak vertical motion under stable conditions. Stability indices like Convective Available Potential Energy (CAPE), which quantifies buoyant energy as the positive area on a sounding diagram between the parcel and environmental temperature profiles, predict convective potential; values above 2000 J/kg indicate severe storm risk. These metrics underscore how dynamic forcing modulates cloud organization and precipitation distribution. Modern classification schemes, such as those employed by the International Satellite Cloud Climatology Project (ISCCP), categorize clouds by properties including optical depth, phase, and top pressure to infer microphysical and radiative characteristics. These approaches account for aerosol indirect effects, where increased particle concentrations can narrow cloud droplet size distributions and enhance albedo, particularly in marine stratocumulus. By combining satellite-derived parameters with aerosol loading data, the ISCCP-H series resolves mixed-phase boundaries and dispersion impacts on cloud radiative properties, aiding improvements in climate model parameterizations.

Observation techniques

Remote sensing methods

Remote sensing methods in cloud physics enable the observation of cloud properties from afar, providing large-scale, non-intrusive data on structure, microphysics, and dynamics without direct sampling. These techniques utilize electromagnetic waves to infer cloud characteristics such as particle size distributions, phase, and vertical extent, often from ground-based, airborne, or spaceborne platforms. Instruments like radars and lidars profile clouds vertically, while satellite sensors offer global coverage, facilitating studies of cloud evolution and interactions over remote regions. Radar systems, particularly millimeter-wave cloud profiling radars operating at frequencies like 35 GHz or 94 GHz, detect hydrometeors by measuring the backscattered power from cloud particles. The equivalent radar reflectivity factor Z, a key observable, quantifies the scattering intensity and is defined as Z = \int_0^\infty N(D) D^6 \, dD, where N(D) is the particle size distribution function and D is the equivalent diameter; this sixth-power dependence makes radar highly sensitive to larger particles in precipitation-laden clouds. Doppler radar capabilities further reveal particle fall velocities and internal air motions by analyzing the frequency shift in returned signals, allowing differentiation of updrafts, downdrafts, and sedimentation processes within clouds. These measurements support retrievals of precipitation rates and microphysical parameters, though attenuation in heavy rain limits penetration into denser regions. Lidar instruments complement radar by probing finer-scale features using laser pulses to measure attenuated backscatter from aerosols and cloud particles, enabling detailed studies of aerosol-cloud interactions that influence droplet activation and cloud albedo. Backscatter profiles reveal aerosol loading and its role in modifying cloud condensation nuclei availability, with enhanced backscatter indicating polluted environments that suppress precipitation formation. Depolarization techniques, which assess the change in polarization state of the returned light, discriminate cloud phase by exploiting the irregular shapes of ice crystals versus spherical droplets; high depolarization ratios (>0.3) typically signify ice-dominated clouds, aiding in the identification of mixed-phase regions critical for . Satellite-based provides synoptic views of properties, with instruments like the (MODIS) on NASA's and Aqua satellites retrieving \tau, defined as the vertical integral of the , \tau = \int_0^H \beta_{\text{ext}}(z) \, dz, where \beta_{\text{ext}} is the and H is height; this parameter gauges thickness and radiative impact from multispectral reflectance in visible and near-infrared bands. The CloudSat mission's Profiling (CPR), operating at 94 GHz, delivers high-resolution (240 m vertical) profiles of vertical structure, detecting weak echoes down to -30 dBZ to map hydrometeor layers and distinguish liquid, ice, and precipitating phases globally. These combined observations from the enhance understanding of geometry and its role in Earth's energy balance. Recent advancements as of 2025 include hyperspectral imagers that improve retrievals of cloud droplet effective radius r_e, the intensity-weighted mean radius influencing and efficiency. Instruments like the Enhanced MODIS Airborne Simulator (eMAS) leverage finer to derive r_e with reduced biases in -influenced scenes, achieving accuracies within 2-3 μm compared to in-situ validations during campaigns like ORACLES. Such techniques, building on bispectral methods, enable more precise assessments of indirect effects by resolving subtle signatures in reflectance. Derived parameters, such as , can be inferred from these profiles to contextualize microphysical states.

In-situ and laboratory measurements

In-situ measurements of cloud microphysics involve direct sampling within cloud environments using instrumented , balloons, and increasingly drones, providing high-resolution data on particle sizes, concentrations, and dynamics that complement broader . These methods capture empirical details essential for understanding cloud formation and evolution, such as droplet spectra and habits, under natural conditions. Key instruments include the Forward Scattering Spectrometer Probe (FSSP), an optical particle counter that measures cloud droplet sizes from approximately 2 to 50 μm in diameter by detecting light scattering angles, enabling precise sizing and concentration estimates in liquid clouds. For ice particles, two-dimensional imaging probes like the 2D-C (for cloud-sized particles) and 2D-P (for precipitation-sized particles) capture images to determine particle habits, orientations, and sizes ranging from tens to thousands of micrometers, revealing diverse shapes such as plates, columns, and aggregates. These probes, often mounted on , account for flow distortions and shattering artifacts to ensure accurate in-cloud sampling. Aircraft-based campaigns exemplify in-situ applications, such as the Clouds, Radiation, Aerosol Transport Experimental Study () conducted in January-February 2018, which deployed the NSF/NCAR HIAPER aircraft to probe boundary layer clouds, measuring updraft speeds up to 2 m/s in supercooled layers and linking them to droplet activation and precipitation formation. More recent efforts, like the goSouth-2 in 2025, used instrumented flights from to contrast clean-air cloud microphysics with polluted regimes, quantifying influences on droplet number concentrations. Laboratory experiments replicate cloud conditions to isolate processes unfeasible in the field. The AIDA (Aerosol Interaction and Dynamics in the Atmosphere) cloud chamber at Karlsruhe Institute of Technology simulates cirrus formation through adiabatic expansions, measuring heterogeneous ice nucleation rates on aerosols like mineral dust, with rates varying from 10^{-6} to 10^{-3} cm^{-3} s^{-1} depending on temperature and cooling rates of 0.5-3 K min^{-1}. Wind tunnels, such as the vertical facility at the University of Mainz, assess collision efficiencies between droplets by levitating pairs in controlled airflows, revealing efficiencies of 0.1-0.6 for radii 10-100 μm under turbulent conditions relevant to warm rain development. Balloon-borne sondes provide vertical profiles through clouds, equipped with sensors for , relative humidity, and to detect supersaturation zones critical for , as demonstrated in campaigns like Balloon-borne Aerosol-Cloud Interaction Studies (BACIS), which resolved humidity gradients within mixed-phase layers to ±0.5% accuracy. Emerging integrations of drones in 2025 enhance sampling, with lightweight probes measuring , humidity, and particle concentrations at altitudes up to 1 km, offering flexible, high-spatial-resolution data (e.g., 1 Hz sampling) in shallow cumuli without the logistical demands of manned flights.

Key cloud properties

Microphysical parameters

Microphysical parameters provide quantitative descriptions of cloud particle populations, focusing on their size distributions, concentrations, and mass contents, which are essential for understanding formation, evolution, and radiative impacts. These parameters characterize the ensemble properties of droplets and ice particles within , influencing processes such as light scattering and development. (LWC) measures the mass of liquid per volume of air and is a fundamental indicator of density and optical thickness in warm . It is calculated as \mathrm{LWC} = \frac{\pi}{6} \rho_w \int_0^\infty N(D) D^3 \, dD, where \rho_w is the density of liquid (approximately 1000 kg m^{-3}), N(D) is the droplet size distribution function representing the number of droplets per volume with diameters between D and D + dD, and the integral sums the volume contributions over all sizes assuming spherical droplets. water content (IWC) analogously quantifies the mass of particles per volume, often approximated by integrating the mass-dimensional relationship m(D) over the particle size distribution: \mathrm{IWC} = \int_0^\infty m(D) N(D) \, dD, where m(D) accounts for non-spherical habits and effective densities typically lower than \rho_w (around 500–900 kg m^{-3} for various crystals). These contents typically range from 0.01 to 1 g m^{-3} in non-precipitating , establishing the scale of substance available for microphysical processes. The effective r_e represents an area-weighted mean , crucial for parameterizing in radiative transfer models. For liquid clouds, it is defined as r_e = \frac{\int_0^\infty r^3 N(r) \, dr}{\int_0^\infty r^2 N(r) \, dr}, where the integrals are over the (equivalent for ). This links the third and second moments of the size to properties, with extinction efficiency Q \approx 2 for droplets much smaller than visible wavelengths due to . Typical values of r_e are 5–15 \mum in continental clouds and 10–20 \mum in , directly affecting shortwave and . Cloud particle size distributions N(D) describe the spectrum of particle sizes and are commonly parameterized using the three-parameter gamma distribution: N(D) = N_0 D^\alpha \exp(-\lambda D), where N_0 is the intercept parameter (in m^{-3} \mum^{-1-\alpha}), \alpha is the shape parameter (often 0 for exponential or 2–7 for broader distributions), and \lambda is the slope parameter (in \mum^{-1}) controlling the tail. This form captures the skewness observed in droplet spectra, with \alpha > 0 indicating narrower distributions than the exponential case (\alpha = 0). Alternatively, the lognormal distribution is used for its ability to model multimodal spectra: N(D) = \frac{N_t}{\sqrt{2\pi} \sigma D} \exp\left( -\frac{(\ln D - \ln D_m)^2}{2\sigma^2} \right), where N_t is the total number concentration, D_m the median diameter, and \sigma the geometric standard deviation (typically 1.2–2.0). The gamma distribution is preferred for its simplicity in analytical integrations for moments like LWC, while lognormal better fits aerosol-influenced clouds. The total number concentration N_\mathrm{tot} = \int_0^\infty N(D) \, dD quantifies the abundance of particles, ranging from 10–100 cm^{-3} in clean marine to 1000–10,000 cm^{-3} in polluted environments, reflecting aerosol activation. The of the size distribution, defined as the relative \epsilon = \sigma_r / r_m (where \sigma_r is the standard deviation and r_m the mean radius of the distribution), modulates radiative effects; for fixed LWC, lower \epsilon (narrower distributions) yields smaller r_e and higher by increasing , with studies showing significant modulation up to 10–30% albedo enhancement in low- regimes compared to broader ones. These parameters are derived from in-situ probes and , providing for cloud models.

Dynamic and thermodynamic properties

Cloud dynamics are characterized by vertical motions, including updrafts and downdrafts, which drive the of air, , and particles within clouds. speeds (w) typically range from 0.5 to 5 m/s in shallow , increasing to 10 m/s or more in deep convective systems, while downdrafts often exhibit comparable magnitudes but are influenced by evaporative cooling and loading. These motions contribute to , quantified by (TKE), defined as \text{TKE} = \frac{1}{2} (u'^2 + v'^2 + w'^2), where u', v', and w' are the fluctuating components of the velocity in the horizontal and vertical directions, respectively. In boundary layer clouds, TKE values commonly range from 0.1 to 1 m²/s², reflecting the intensity of small-scale eddies that mix air parcels. Thermodynamic properties govern cloud stability through buoyancy forces, where positively buoyant air parcels rise due to warmer temperatures relative to the environment. A key measure is convective available potential energy (CAPE), which quantifies the energy available for convection and is calculated as \text{CAPE} = \int_{LFC}^{EL} g \frac{\theta_e - \theta_{env}}{\theta} \, dz, with g as gravitational acceleration, θ_e the equivalent potential temperature of the parcel, θ_env the environmental potential temperature, θ the reference temperature, and the integral from the level of free convection (LFC) to the equilibrium level (EL). Typical CAPE values in thunderstorm environments range from 1000 to 2500 J/kg for moderate instability, supporting updraft intensification. Entrainment introduces dry environmental air into , diluting their and , with rates (μ) typically on the order of 0.1–1 km⁻¹ in shallow cumuli, though observations show averages around 0.6 km⁻¹ and ranges up to 2.8 km⁻¹ depending on depth and . This process leads to mixing, visualized in mixing diagrams that plot droplet properties like size versus ; homogeneous mixing assumes uniform dilution across the parcel, while inhomogeneous mixing preferentially evaporates smaller droplets, altering microphysical evolution near edges. Radiative processes influence cloud , particularly through longwave cooling, which drives and changes. In clouds, radiative cooling rates approximate dT/dt ≈ −10 K/day, primarily due to emission from ice particles in the cold upper , though net effects include partial compensation by of upwelling .

Modeling approaches

Parameterization schemes

Parameterization schemes in cloud physics provide simplified mathematical representations of sub-grid-scale microphysical processes within and climate models, enabling efficient computation of cloud formation, evolution, and without resolving individual particles. These schemes approximate complex interactions such as , , and fallout by relating bulk variables like mixing ratios and number concentrations to process rates, often tuned to observations for realism. Bulk microphysics schemes, a of these parameterizations, categorize hydrometeors into a few classes (e.g., cloud , , cloud , ) and predict their evolution through prognostic equations. Single-moment bulk schemes predict only the mass mixing ratio for each hydrometeor category, assuming fixed intercept parameters in the , which limits their ability to capture variability in particle number and size. A seminal example is the Lin scheme, which prognostically treats (LWC) alongside , , , and , incorporating processes like , , and collection while assuming gamma distributions for simplicity. In contrast, double-moment schemes predict both mass mixing ratios and total number concentrations, allowing dynamic evolution of mean particle sizes and improving representations of influences and efficiency; studies show they outperform single-moment schemes in simulating droplet spectra and rainfall accumulation by up to 20-30% in mid-latitude storms. Autoconversion, the collision and coalescence of droplets to form , is parameterized in schemes to initiate ; the Sundqvist scheme, widely adopted in models, uses a form P = k q (1 - \exp(-b q)) where q is the mixing , providing a smooth threshold behavior tuned to observed warm processes (typical b \approx 10^3 kg g^{-1}). Accretion terms extend this by representing the collection of by larger drops, often using continuous collection equations with kernels derived from stochastic collision theory; in the Lin scheme, accretion rates scale with rain mixing and fall speed differences, contributing significantly to warm production in stratiform clouds. Ice processes in parameterization schemes address mixed-phase and cold clouds through and mechanisms. The Meyers scheme parameterizes heterogeneous ice by or biological particles, with ice nucleating particle (INP) concentration scaling exponentially with ice as n_{\text{INP}} = \exp\left[12.96(S_i - 1) - 0.639\right] L^{-1} (or \approx 10^{3} \exp\left[12.96(S_i - 1)\right] m^{-3}), where S_i is ice relative humidity, calibrated to and deposition freezing observations and reducing numbers by factors of 10-100 compared to homogeneous alone. of particles is modeled using power-law terminal velocities v = a D^b, where D is maximum , a and b are empirical coefficients (e.g., a \approx 10^{-3} m^{1-b} s^{-1}, b \approx 0.5-1 for aggregates), derived from data to account for and effects, influencing optical by up to 20% in global simulations. Aerosol-aware schemes integrate (CCN) activation to link aerosols with droplet numbers, enhancing indirect effect representations. In the 2025 update to the ECMWF Integrated Forecasting System (IFS), prognostic CCN treatment via Köhler theory improves low-cloud susceptibilities, revealing regional biases in aerosol-cloud interactions consistent with satellite observations of cloud droplet number concentrations.

Numerical simulations

Numerical simulations in cloud physics employ high-resolution computational models to explicitly resolve the and microphysical processes within clouds, bridging the gap between observational and theoretical understanding. These approaches simulate turbulent flows, droplet growth, and phase changes at scales that capture individual structures, often requiring sophisticated numerical schemes to handle the multiscale nature of atmospheric phenomena. Unlike coarser models that rely on parameterizations, numerical simulations aim to resolve key physical processes directly, enabling detailed studies of formation, evolution, and interactions with and aerosols. Detailed microphysics can also employ spectral bin schemes that resolve full distributions with multiple bins. Large-eddy simulation (LES) is a cornerstone method for modeling cloud dynamics, resolving turbulent eddies larger than the grid spacing Δx, typically around 100 m, while parameterizing subgrid-scale effects. In LES of marine stratocumulus , for instance, the model captures -layer and cloud-top entrainment, using subgrid turbulent kinetic energy (TKE) closures to represent unresolved motions. This technique has been instrumental in the diurnal cycle of stratocumulus and indirect effects, providing insights into cloud susceptibility to perturbations. Seminal work on LES for cloud-topped boundary layers emphasized the need for accurate subgrid models to handle buoyancy-driven . Cloud-resolving models (CRMs) extend this resolution to explicitly simulate convective clouds over larger domains, incorporating detailed microphysics schemes such as bin methods that track distributions () with 100 to 1000 size bins. These bins evolve the through processes like , , and collision-coalescence, allowing for realistic representation of formation in cumulus or systems. CRMs have been used to mesoscale convective systems, revealing sensitivities to environmental conditions like and moisture. A comprehensive highlights their role in advancing understanding of tropical without cumulus parameterizations. Direct numerical simulation (DNS) resolves all scales of turbulence down to the Kolmogorov length, but its application to cloud physics is severely limited by computational demands, typically confined to Reynolds numbers < 10^4 and small domains representing only portions of a cloud. In cloud-edge mixing studies, DNS has quantified droplet size distributions under decaying , but scaling to realistic atmospheric (often >10^6) remains infeasible without approximations. Hybrid approaches combine DNS for microscale processes with for larger eddies, enhancing accuracy in particle-turbulence interactions. By 2025, advancements in numerical simulations include (ML) acceleration for modeling, where parameterize and radiative effects, improving efficiency in global-scale simulations constrained by satellite data. Additionally, GPU-based global CRMs have achieved unprecedented resolutions, embedding high-fidelity cloud physics in multiscale frameworks to simulate stratocumulus coverage at 200 m scales on supercomputers. These developments, such as ML-driven emulators for particle-resolved DNS, address computational bottlenecks and enable hybrid resolved-parameterized modeling for climate projections.

References

  1. [1]
  2. [2]
  3. [3]
    [PDF] Atmospheric Physics - An Introduction to Clouds
    ▷ Definition: A cloud is an aggregate of cloud droplets or ice crystals, or a combination of both, suspended in air. ▷ Clouds are a major factor in the Earth's ...
  4. [4]
    Meteorology by Aristotle - The Internet Classics Archive
    When the cloud burns as it is drawn downwards, that is, when the exhalation becomes rarer, it is called a fire-wind, for its fire colours the neighbouring air ...
  5. [5]
    The exhalations (Chapter 3) - Structure and Method in Aristotle's ...
    For this reason, Aristotle sought principles that are special and appropriate to meteorology and found them in the exhalations of early Ionian physical ...
  6. [6]
    The Science and Art of Meteorology - National Geographic Education
    Dec 9, 2024 · Evangelista Torricelli, an Italian physicist, observed that changes in air pressure were connected to changes in weather. In 1643, Torricelli ...<|control11|><|separator|>
  7. [7]
    Physical and Meteorological Observations, Conjectures, and Sup …
    The cold air descending from above, as it penetrates our warm region full of watry particles, condenses them, renders them visible, forms a cloud thick and dark ...
  8. [8]
    [PDF] Eighteenth Century Theories of Evaporation and Rain
    'Benjamin Franklin. uPhysical and Meteorological Observations, Con- jectures, and Suppositions," Phil. TrelM. Roy. Soc. London LV (1766),. 182-192. ·Ibid.~ p ...Missing: condensation | Show results with:condensation
  9. [9]
    On The Modification Of Clouds - Met Office
    Howard, an amateur meteorolgist and professional pharmacist, devised the cloud classification scheme the basis of which is still in use today.
  10. [10]
    Accurate Expressions for the Dewpoint and Frost Point Derived from ...
    John Dalton first introduced a method for reliably measuring the dewpoint ... Errors are calculated relative to laboratory measurements of vapor pressure ...
  11. [11]
    MEASUrEMENtS Of AtMOSPHErIc NANOPArtIclES (1875–1980)
    1974) using a thermal gradient diffusion cloud chamber (working fluid ethylene glycol). ... Vincent and S. Watanabe. From molecules to meteorology via turbulent ...
  12. [12]
    Review of sub-3 nm condensation particle counters, calibrations ...
    The first CPC experiments were conducted by Coulier (Coulier Citation1875a, Citation1875b) and Aitken (Aitken Citation1880a, Citation1880b), and further ...
  13. [13]
    [PDF] The Wegener-Bergeron-Findeisen process - Semantic Scholar
    Abstract. The Wegener-Bergeron-Findeisen process refers to the rapid growth of ice crystals at the expense of surrounding cloud droplets,.
  14. [14]
    [PDF] HISTORY OF PROJECT CIRRUS - DTIC
    It would be far better to study cloud particle growth in airplane flights. That would require the development of now Instruments. This was late in-1946. 'They ...Missing: WWII | Show results with:WWII
  15. [15]
    [PDF] A State‐of‐the‐Art Experimental Laboratory for Cloud and ... - CORE
    Nov 1, 2011 · Besides cloud chambers, other facilities such as specially designed vertical wind tunnels have been used to simulate the dynamics of raindrops ...
  16. [16]
    [PDF] 19820004236.pdf - NASA Technical Reports Server (NTRS)
    During the 1980's the Space Shuttle will provide a plat- form for ... through air; cloud drop nucleation and growth; ice crystal growth from the ...
  17. [17]
  18. [18]
  19. [19]
    [PDF] Atmospheric Thermodynamics - ECMWF
    ... derive the dry adiabatic lapse rate of the atmosphere pd. dT g dz c. = -. (38) which amounts to a temperature decrease of 0.98 K per 100 m. 9. Reversible ...
  20. [20]
    [PDF] 6. Potential Temperature ( )
    The pressure and temperature can be combined into a single variable, the potential temperature, and it will be shown that the atmosphere is stable if the ...
  21. [21]
    [PDF] NOAA Technical Memorandum NWS SR-109
    The isotherm value at this point is equal to the equivalent temperature. process, and were. Equivalent potential temperature The temperature a parcel of air.
  22. [22]
    [PDF] 4 Water Vapor - UBC EOAS
    How do the vapor pressure values from Tetens' formula differ from those of the Clausius-Clapeyron equation, as a function of temperature? Find the answer.
  23. [23]
    [PDF] Vapor pressure of water at its triple point
    It is our purpose to measure with the highest possible accuracy the vapor pressure of water for - 5°C < t 68 < 100°C. We chose to concentrate ini- tially on ...
  24. [24]
  25. [25]
    A Simple Accurate Formula for Calculating Saturation Vapor ...
    Jun 1, 2018 · A Simple Accurate Formula for Calculating Saturation Vapor Pressure of Water and Ice ... 1996: Improved Magnus's form approximation of saturation ...Abstract · Introduction · Methods · Results
  26. [26]
    [PDF] improved magnus' form approximation of saturation - OSTI.gov
    The saturation vapor pressure of pure water vapor is denoted by E, or E, for saturation with respect to water or ice, respectively. When moist air exists in a ...
  27. [27]
    5.4 Curvature Effect: Kelvin Effect | METEO 300 - Dutton Institute
    Dependence of the ratio of the saturation vapor pressure over a curved surface to the saturation vapor pressure over a flat surface on the drop radius.
  28. [28]
    [PDF] The Kelvin equation - Rutgers University, GLASS |
    The Kelvin equation tells us how the vapor pressure depends on the curvature of the liquid. The cause for this change in vapor pressure is the Laplace pressure.
  29. [29]
    [PDF] A single parameter representation of hygroscopic growth and cloud ...
    Apr 18, 2007 · Values of the hygroscopicity parameter are between 0.5 and 1.4 for highly-CCN-active salts such as sodium chlo- ride, between 0.01 and 0.5 for ...
  30. [30]
    Anthropogenic contribution to cloud condensation nuclei and the ...
    May 29, 2013 · We show that anthropogenic emissions increase the global mean CCN in the lower troposphere by ∼60–80% and cloud droplet number concentration by ∼40%.
  31. [31]
    [PDF] Global cloud condensation nuclei influenced by carbonaceous ...
    Assuming that the carbonaceous combustion aerosol does not act as CCN, modelled CCN concentrations do not exceed ∼1000 cm−3 even in polluted regions. When ...
  32. [32]
    The nucleus in and the growth of hygroscopic droplets
    The nucleus in and the growth of hygroscopic droplets. Hilding Köhler. Abstract. The first page of this article is displayed as the abstract.
  33. [33]
    [PDF] Aerosol- and updraft-limited regimes of cloud droplet formation ...
    Sep 24, 2009 · The aerosol-limited regime of CCN activation is characterized by a relatively high ratio between the updraft velocity and particle number ...
  34. [34]
    Role of updraft velocity in temporal variability of global cloud ... - PNAS
    Increased competition for water vapor by CCN promotes the sensitivity to updraft fluctuations. For ice crystals, updraft velocity is the primary driver of ...
  35. [35]
  36. [36]
    How Clouds Form | National Oceanic and Atmospheric Administration
    Jul 1, 2025 · Those objects are called cloud condensation nuclei. These nuclei are minute solid and liquid particles found in abundance in the atmosphere ...Missing: sulfates | Show results with:sulfates
  37. [37]
    5.3 How can supersaturation be achieved? | METEO 300
    Three basic mechanisms for cooling the air are RUM: Radiation, Uplift, and Mixing. Radiation and mixing happen at constant pressure (isobaric); uplift happens ...Missing: non- | Show results with:non-
  38. [38]
    [PDF] The Moist Adiabatic Lapse Rate
    The gravity term of (1) represents the effects of buoyancy. Eq. (2) describes dry adiabatic motion. We assume for simplicity that. ∂s. ∂z is ...Missing: (c_p R_v
  39. [39]
    [PDF] SKEW-T, LOG-P DIAGRAM ANALYSIS PROCEDURES
    Sep 11, 2007 · Convection Condensation Level, and the Convection Temperature. L. Lifting Condensation Level (LCL) ... H(meters) = 125(T-Td). T and Td in oC. H( ...<|control11|><|separator|>
  40. [40]
    CLOUD DEVELOPMENT
    Certain types and shapes of dust and salt particles, such as sea salts and clay, make the best condensation nuclei. With proper quantities of water vapor and ...Missing: sulfates | Show results with:sulfates
  41. [41]
    [PDF] Lecture 7a: Cloud Development and Forms - UCI ESS
    When boundaries between air of unlike temperatures (fronts) migrate, warmer air is pushed aloft. ❑ This results in adiabatic cooling and cloud formation.
  42. [42]
    Surface‐Atmosphere Decoupling Prolongs Cloud Lifetime Under ...
    May 17, 2023 · This cloud-prolonging effect by warm-air advection is caused by the warm-advection-induced stabilization/decoupling of the surface-atmosphere ...
  43. [43]
    A Lagrangian View of Moisture Dynamics During DYNAMO - PMC
    The LCT is almost always positive (red and blue curves), indicating a net moistening by surface evaporation minus precipitation and vertical advection.
  44. [44]
    Moist Orographic Convection: Physical Mechanisms and Links to ...
    Feb 25, 2018 · This paper reviews the current understanding of moist orographic convection and its regulation by surface-exchange processes.
  45. [45]
    Discussion on Humidity - National Weather Service
    Absolute and specific humidity are quite similar in concept. RELATIVE HUMIDITY: Relative humidity (RH) (expressed as a percent) also measures water vapor, but ...Missing: profiles physics
  46. [46]
    Relative Humidity as an Indicator for Cloud Formation over ...
    Abstract. The influence of land surface heterogeneity on potential cloud formation is investigated using relative humidity as an indicator.
  47. [47]
    Diurnal variations of cloud and relative humidity profiles across the ...
    Nov 5, 2019 · Here we combine recent satellite observations of clouds profiles and relative humidity profiles to document the diurnal variations of the water vapor and ...
  48. [48]
    [PDF] Isentropic Analysis and Interpretation - DTIC
    Isentropic uplifting typi- cally represents not only vertical motion but moisture advection as well. Simi- larly, subsidence and drying out of air behind ...
  49. [49]
    The Use of Moisture Flux Convergence in Forecasting Convective ...
    Moisture flux convergence (MFC) is a term in the conservation of water vapor equation and was first calculated in the 1950s and 1960s as a vertically integrated ...Missing: uplift | Show results with:uplift
  50. [50]
    Uplift mechanisms - The Physical Environment
    Convergent uplift occurs when air enters a center of low pressure. As air converges into the center of a cyclone it is forced to rise off the surface. As the ...Missing: isentropic physics<|control11|><|separator|>
  51. [51]
    A review on factors influencing fog formation, classification ...
    Fog is formed when air approaches saturation by adding moisture (by evaporation or mixing two air parcels of contrasting temperatures) or by removing heat.
  52. [52]
    Formation and Development of Nocturnal Boundary Layer Clouds ...
    The formation and evolution of nocturnal boundary layer clouds over land are studied using a simple well-mixed boundary layer theory.Missing: buildup | Show results with:buildup
  53. [53]
    Effect of the boundary layer low-level jet on fast fog spatial propagation
    Nov 9, 2023 · The boundary layer low-level jet (BLLJ) causes fast fog propagation by transporting moisture and warm air, increasing humidity and enhancing ...Missing: buildup | Show results with:buildup
  54. [54]
    Observational Constraints on Warm Cloud Microphysical Processes ...
    We introduce new parameterizations for autoconversion and accretion rates that greatly improve representation of the growth processes of warm rain.
  55. [55]
    [PDF] 1. Droplet Growth by Condensation - atmo.arizona.edu
    Initially, droplets that form on large condensation nucleus grow faster, but after a certain radius they homogenize. We can also use the equation for ...
  56. [56]
    [PDF] SIO 217A: Cloud Droplet Growth
    Dec 6, 2013 · If S < 1, then dr/dt < 0, which describes the evaporation of a cloud ... Similarily, increased values of S -1 increase the droplet growth rate.
  57. [57]
    [PDF] 4. Droplet Growth in Warm Clouds
    the supersaturation S, the rate of increase dr/dt is inversely proportional to the radius r of the droplet. We write r dr = G`. S dt which can be integrated ...
  58. [58]
    Collision Efficiency of Drops in a Wide Range of Reynolds Numbers
    An approach is developed enabling one to calculate the collision efficiency and the collision kernel within a wide range of the Reynolds numbers.
  59. [59]
    Rate of loss of cloud droplets by coalescence in warm clouds
    Nov 8, 2006 · Loss rates are found to be surprisingly high even for the modest precipitation rates found in drizzling boundary layer clouds and can be used to ...
  60. [60]
    THE DISTRIBUTION OF RAINDROPS WITH SIZE in - AMS Journals
    The distribution of raindrops with size. JS Marshall and W. Mc K. Palmer. Article History. Get Permissions. Full access. Abstract. No Abstract Available.
  61. [61]
    Statistical theory of the Marshall-Palmer distribution of raindrops
    Single-drop fragmentation determines size distribution of raindrops. 2009, Nature Physics. Physically based two-moment bulkwater parametrization for warm ...
  62. [62]
    Parameterization of the Autoconversion Process.Part I - AMS Journals
    The Kessler-type autoconversion parameterizations that have been widely used in cloud-related modeling studies are theoretically derived and analyzed by ...
  63. [63]
    Extending Twomey's Analytical Estimate of Nucleated Cloud Droplet ...
    The present scheme has been incorporated in a newly developed bulk scheme to study warm precipitating clouds. It is currently being tested to examine the ...
  64. [64]
    Evaluation of autoconversion schemes in a single model framework ...
    Oct 21, 2015 · We examined the performance of autoconversion (mass transfer from cloud water to rainwater by the coalescence of cloud droplets) schemes in warm rain.
  65. [65]
    (PDF) The Wegener-Bergeron-Findeisen process - Its discovery and ...
    May 10, 2025 · The Wegener-Bergeron-Findeisen process refers to the rapid growth of ice crystals at the expense of surrounding cloud droplets.
  66. [66]
    the representation of heterogeneous ice activation in cloud models
    Oct 13, 2021 · The homogeneous nucleation of ice in supercooled liquid-water clouds is characterized by time-dependent freezing rates.
  67. [67]
    The Growth of Atmospheric Ice Crystals: A Summary of Findings in ...
    Abstract Measurements of ice crystal growth under free fall in a generation of vertical supercooled cloud tunnels and some static cloud chambers as well as ...
  68. [68]
    Ice Crystal Habits and Growth Processes in Stratiform Clouds with ...
    Both riming and aggregation processes played central roles in the broadening of particle size distributions (PSDs), and these processes were more active in ...
  69. [69]
    Ice Multiplication by Breakup in Ice–Ice Collisions. Part II: Numerical ...
    Breakup in ice–ice collisions is by far the most prolific process of ice initiation in the simulated clouds (95%–98% of all nonhomogeneous ice).
  70. [70]
  71. [71]
    Genera | International Cloud Atlas
    The classification of clouds has ten main groups, called genera. Each observed cloud is a member of one, and only one, genus.
  72. [72]
    The Useful Pursuit of Shadows | American Scientist
    Figure 1. Clouds were first described systematically by Luke Howard, a London pharmacist and amateur meteorologist, in an 1802 address to the Askesian Society.
  73. [73]
    Genera | International Cloud Atlas
    Consideration of the most typical forms of clouds leads to the recognition of ten genera. The definitions of the genera given below do not cover all possible ...Cirrus · Cumulonimbus · Stratocumulus · Cirrocumulus<|separator|>
  74. [74]
    Ten Basic Clouds | National Oceanic and Atmospheric Administration
    Mar 28, 2023 · Cirrus (Ci), cirrocumulus (Cc), and cirrostratus (Cs) are high level clouds. They are typically thin and white in appearance, but can appear in a magnificent ...NWS Cloud Chart · The Four Core Types of Clouds · A 'Hole' Lot of Clouds 2
  75. [75]
    Genera and species | International Cloud Atlas
    Genera and species ; Altocumulus. Ac ; Altostratus. As ; Nimbostratus. Ns ; Stratocumulus. Sc.
  76. [76]
    Levels | International Cloud Atlas
    The troposphere can be vertically divided into three levels, formerly known as “étages”: high, middle and low. Each level is defined by the range of heights.
  77. [77]
    Supplementary features | International Cloud Atlas
    Clouds sometimes have supplementary features attached to or partly merged with them. P.10.5 (Cb) incus P.10.6 (Cb) mamma P.4.14 (Ac) virga P.9.9 (Cb) ...
  78. [78]
    Aircraft condensation trails | International Cloud Atlas
    Aircraft condensation trails (contrails) that have persisted for at least 10 minutes will be given the name of the genus, Cirrus, followed only by the special ...
  79. [79]
    ISCCP Definition of Cloud Types - NASA
    The standard ISCCP products assume that clouds warmer than 260 K are liquid clouds composed of spherical droplets with an effective radius of 10 microns and ...
  80. [80]
    Microphysical Processes in Ice and Mixed-Phase Clouds (Chapter 6)
    Aug 22, 2018 · Clouds containing liquid drops and ice particles belong to the mixed-phase cloud type. Most deep convective clouds, even in the tropics, are ...
  81. [81]
    Liquid‐Ice Mass Partitioning Across the Edge of Mixed‐Phase ...
    Nov 30, 2023 · One of the major factors controlling the phase partitioning in mixed-phase cloud is entrainment mixing, but it is still poorly understood.Missing: classifications | Show results with:classifications
  82. [82]
    [PDF] A FOCUS ON MIXED-PHASE CLOUDS
    Mixed-phase clouds are complete cloud systems containing both liquid and ice water in the vertical that interact via microphysical processes.<|separator|>
  83. [83]
    Cloud Classification - National Weather Service
    Cloud Classification and Characteristics. Clouds are classified according to their height above and appearance (texture) from the ground.
  84. [84]
    [PDF] THE TEN DIFFERENT TYPES OF CLOUDS - Amazon AWS
    Of the ten main cloud types, only two are consistent precipitation producers: nimbostratus and cumulonimbus. Nimbostratus are responsible for continuous.
  85. [85]
    Cloud classification summary | International Cloud Atlas
    Cloud classification summary ; Genera · (listed by frequency of observation) ; Cirrus · fibratus · uncinus · spissatus · castellanus · floccus · intortus
  86. [86]
    THUNDERSTORMS
    Dynamic precipitation tends to have a less intense rain rate than convective precipitation and also tends to last longer. While stratiform rain is the product ...
  87. [87]
    Stability Indices
    Stability indices are a group of indices that have been developed to illustrate the potential for convection within an air mass based upon a radiosonde or ...Missing: dynamic stratiform
  88. [88]
    Atmospheric stability | Atmospheric Physics Class Notes - Fiveable
    Convective available potential energy (CAPE) indicates thunderstorm potential; Stability indices (Lifted Index, K-Index) help forecast thunderstorm likelihood ...
  89. [89]
    Regional variability of aerosol impacts on clouds and radiation in ...
    Jul 23, 2025 · Anthropogenic aerosols are a primary source of uncertainty in future climate projections. Changes to aerosol concentrations modify cloud ...
  90. [90]
    International Satellite Cloud Climatology Project (ISCCP)-H Series
    The International Satellite Cloud Climatology Project (ISCCP) provides data on various aspects of cloud variables like cloud amount, optical thickness, and ...
  91. [91]
    Global quantification of the dispersion effect with POLDER satellite ...
    Aug 2, 2025 · Investigation of aerosol indirect effects on monsoon clouds using ground-based measurements over a high-altitude site in Western Ghats.
  92. [92]
    [PDF] Chapter 13 Remote Sensing of Precipitation from Airborne and ...
    Weather radar measurements from airborne or satellite platforms can be an effective remote sensing tool for examining the three-dimensional structures of ...
  93. [93]
    8.2: Weather Radars - Geosciences LibreTexts
    Dec 14, 2024 · Combining these 8 equations, and using eq. (8.24) to define the reflectivity factor Z gives the traditional form of the radar equation:.
  94. [94]
    On retrieving the microphysical properties of cirrus clouds using the ...
    Specifically, we present an algorithm that uses the zeroth and first moments of the Doppler spectrum (radar reflectivity and mean Doppler velocity) to retrieve ...<|separator|>
  95. [95]
    Peering into Cloud Physics Using Ultra-Fine-Resolution Radar and ...
    The WCR-QPC measures radar Doppler spectra along with the first three moments, i.e., reflectivity, Doppler velocity, and spectrum width. Radar reflectivity is ...Missing: ∫ | Show results with:∫
  96. [96]
    Understanding Aerosol–Cloud Interactions through Lidar Techniques
    This review paper provides an overview of recent advancements in the study of ACI using lidar techniques.
  97. [97]
    Depolarization ratio–effective lidar ratio relation: Theoretical basis ...
    Jun 14, 2007 · This paper introduces a technique for cloud phase discrimination based on linear polarization measurements made by space-based lidar.
  98. [98]
    MOD06_L2 - MODIS/Terra Clouds 5-Min L2 Swath 1km and 5km
    The MODIS Cloud Product will be used to investigate seasonal and inter-annual changes in cirrus (semi-transparent) global cloud cover and cloud phase.Missing: depth τ = ∫ β_ext dz CloudSat CPR 94 GHz vertical structure
  99. [99]
    Spaceborne Cloud and Precipitation Radars: Status, Challenges ...
    Existing and planned spaceborne radar missions provide cloud and precipitation information over the oceans and land difficult to access in remote areas.
  100. [100]
    CloudSat – aeris
    CPR (Cloud Profiling Radar) · Nominal frequency 94 GHz · Pulse width 3.3 µsec · PRF 4300 Hz · Minimum detectable reflectivity factor < -29 dBZ · Antenna size 1.85 m.Missing: MODIS τ = β_ext dz
  101. [101]
    Evaluating spectral cloud effective radius retrievals from the ... - AMT
    Feb 27, 2025 · This study evaluates cloud effective radius (CER) retrievals from eMAS using data from ORACLES, comparing eMAS and RSP with in situ data, and ...Missing: advancements | Show results with:advancements
  102. [102]
    [PDF] Retrieving Vertical Profiles of Cloud Droplet Effective Radius using ...
    Feb 13, 2025 · Cloud droplet effective radius is retrieved using MODIS spectral measurements, calculated by the minimum of the difference between two MODIS ...Missing: advancements | Show results with:advancements
  103. [103]
    Cloud Ice Properties: In Situ Measurement Challenges in
    The imaging probes that are most frequently employed in aircraft measurements are the Two-Dimensional Cloud and Precipitation spectrometers (2D-C, 2D-P), the ...
  104. [104]
    In Situ Cloud Instruments - Centre for Atmospheric Science
    A forward scattering spectrometer probe (FSSP) is an optical particle counter, suitable for counting and sizing particles in the size range of 0.5 - 47 micron ...
  105. [105]
    On measurements of small ice particles in clouds - Heymsfield - 2007
    The University of North Dakota Citation and the NASA WB57 research aircraft sampled cloud layers from 0 to −70°. The aircraft carried FSSPs, 2D imaging probes ...
  106. [106]
    [PDF] Chapter 11 Processing of Ice Cloud In Situ Data Collected by Bulk ...
    Feb 5, 2025 · These instruments include the Forward Scat- tering Spectrometer Probe (FSSP), a legacy probe originally manufactured by PMS and Particle Metrics.
  107. [107]
    SOCRATES - Earth Observing Laboratory
    SOCRATES will be conducted to improve our understanding of clouds, aerosols, air-sea exchanges, and their interactions over the Southern Ocean.
  108. [108]
    Preparations for goSouth-2
    Apr 18, 2025 · The aim of goSouth-2 is a detailed contrast study of cloud properties in clean Antarctic air compared to clouds in aerosol-polluted Australian air.
  109. [109]
    [PDF] Probing ice clouds by broadband mid-infrared extinction spectroscopy
    Oct 24, 2006 · The technical equipment of the AIDA chamber as well as its operation as an expansion cloud chamber for ice nucleation studies has been described ...
  110. [110]
    Collision efficiencies empirically determined from laboratory ...
    In laboratory experiments at the vertical wind tunnel of the University of Mainz the collisional growth of drops with radii between 70 and 170 μm in radius ...
  111. [111]
    Balloon-borne aerosol–cloud interaction studies (BACIS) - AMT
    Aug 19, 2022 · In situ measurements of aerosol and cloud profiles and background meteorological parameters (temperature, relative humidity, wind speed and ...
  112. [112]
    Data collected by a drone backpack for air quality and atmospheric ...
    Oct 9, 2025 · Additionally, high-resolution TBS profiling of the boundary layer by Chávez-Medina et al. (2025) and aerosol and cloud measurements by Le et al.
  113. [113]
    Updraft and downdraft characterization with Doppler lidar - ACP
    The average value of the updraft mean vertical velocities is 0.5–0.7 m/s or 0.40 w*–0.45 w* (convective velocity scale w*), and the negative downdraft mean ...Missing: typical speeds
  114. [114]
    Estimating Turbulence Kinetic Energy Dissipation Rates in the ...
    May 1, 2019 · We focus on recently proposed methods of the TKE dissipation-rate retrievals based on zero crossings and recovering the missing part of the spectrum.
  115. [115]
    Severe Weather Topics
    In general, CAPE values of less than 1000J/kg represent weak instability, 1000 to 2500J/kg moderate instability, 2500-4000J/kg strong instability, and greater ...
  116. [116]
    Factors Governing Cloud Growth and Entrainment Rates in Shallow ...
    Jun 5, 2021 · We estimate ERs ranging from 0.16 to 2.8 km−1 with an average of 0.58 ± 0.10 km−1 in 469 clouds. Entrainment rate is best correlated with cloud ...
  117. [117]
    Homogeneous and Inhomogeneous Mixing in Cumulus Clouds
    Dec 1, 2009 · The data suggest that homogeneous mixing is more likely to occur in the vicinity of the cloud core, whereas inhomogeneous mixing dominates in more diluted ...
  118. [118]
    Simple Spectral Models for Atmospheric Radiative Cooling in
    At the same time, basic aspects of longwave radiative cooling, such as its characteristic value of 2 K day−1, its sharp decline (or “kink”) in the upper ...
  119. [119]
    Bulk Parameterization of the Snow Field in a Cloud Model in
    Abstract. A two-dimensional, time-dependent cloud model has been used to simulate a moderate intensity thunderstorm for the High Plains region.
  120. [120]
    Make It a Double? Sobering Results from Simulations Using Single ...
    This study adds to the large body of work that has shown that double-moment schemes generally outperform single-moment schemes.Abstract · Introduction · Simulations with single... · Single- versus double-moment...Missing: seminal | Show results with:seminal
  121. [121]
    Refinements to Ice Particle Mass Dimensional and Terminal Velocity ...
    Abstract. This two-part study addresses the development of reliable estimates of the mass and fall speed of single ice particles and ensembles.
  122. [122]
    Introducing aerosol-cloud interactions in the ECMWF model reveals ...
    Abstract. Realistic number concentrations of aerosol cloud condensation nuclei (CCN) are critical to simulate the Earth's atmosphere radiative budget.Missing: aware schemes
  123. [123]
    [PDF] A Review on the Uses of Cloud-(System-)Resolving Models
    A CRM was used to study the sensitivity of cumulus convection to various parameters such as dimensionality, resolution, domain size, and microphysics ( ...
  124. [124]
    [PDF] Turbulence modelling in large-eddy simulations of the cloud-topped ...
    This paper discusses turbulence modelling in large-eddy simulations of the cloud- topped atmospheric boundary layer. While our primary focus is on ...
  125. [125]
    A New Cloud Physics Parameterization in a Large-Eddy Simulation ...
    In this paper, we present a new computationally efficient and relatively simple bulk microphysical scheme refined for LES models of marine stratocumulus using ...Missing: seminal | Show results with:seminal
  126. [126]
    [PDF] Global Cloud-Resolving Models
    Abstract. Purpose of Review Global cloud-resolving models (GCRMs) are a new type of atmospheric model which resolve nonhydrostatic.
  127. [127]
    Direct numerical simulation of turbulence and microphysics in the Pi ...
    Feb 4, 2022 · Due to computational constraints, however, the DNS in the current simulations is restricted to Ra = 7.9 × 10 6 , which is accomplished by ...
  128. [128]
    Direct Numerical Simulation of a Warm Cloud Top Model Interface
    Therefore, due to computational limitations, only a small portion of a cloud could be considered in a DNS study. The basic model used in all such works was ...
  129. [129]
    Cloud‐edge mixing: Direct numerical simulation and observations in ...
    Jan 10, 2017 · A direct numerical simulation (DNS) with the decaying turbulence setup has been carried out to study cloud-edge mixing and its impact on the droplet size ...Introduction · DNS Model and Model Setup · Analogy Between DNS and In...
  130. [130]
    Aerosol–cloud interactions in cirrus clouds based on global-scale ...
    Jul 10, 2025 · Machine learning (ML) models showed that using relative humidity with respect to ice (RHi) as a predictor significantly increased the accuracy ...
  131. [131]
    Improving Stratocumulus Cloud Amounts in a 200‐m Resolution ...
    Mar 6, 2024 · Compared to other available global modeling tools for studying cloud feedback, today's GPU-accelerated Multiscale Modeling Frameworks ...
  132. [132]
    Emulator of PR‐DNS: Accelerating Dynamical Fields With Neural ...
    Feb 18, 2024 · Particle-resolved direct numerical simulations (PR-DNS) play an increasing role in investigating aerosol-cloud-turbulence interactions at ...