Fact-checked by Grok 2 weeks ago

Temperature gradient

The is a quantity in physics and that describes the rate and of spatial variation in within a medium, pointing toward the direction of maximum increase with a equal to that maximum rate of change per unit distance. Mathematically, it is expressed as the operator applied to the , \nabla T = \left( \frac{\partial T}{\partial x} \mathbf{i} + \frac{\partial T}{\partial y} \mathbf{j} + \frac{\partial T}{\partial z} \mathbf{k} \right), where T is and x, y, z are spatial coordinates. In , the gradient drives conductive flow according to Fourier's law, which states that the \mathbf{q} is proportional to the negative of the gradient, \mathbf{q} = -k \nabla T, where k is the thermal conductivity of the material. This relationship underscores its fundamental role in processes where moves from higher to lower regions without motion of the medium. gradients also influence convective , contributing to phenomena like natural convection when exceeding stability thresholds in fluids. Beyond basic thermodynamics, temperature gradients have critical applications across disciplines; in geophysics, the measures the average temperature increase with depth in , typically around 25–30°C per kilometer, influencing volcanic activity and resource exploration. In meteorology, horizontal and vertical temperature gradients drive patterns, such as fronts and jet streams, by determining regions of instability and . Engineering applications include designing thermal barriers in materials to manage stress from uneven heating, as steep gradients can induce mismatches leading to structural failure. In advanced contexts like manufacturing, controlled temperature gradients enable zone refining for purifying crystals by promoting solute migration.

Mathematical Foundations

Definition and Notation

The gradient measures the rate at which changes with respect to in space, quantifying how varies spatially within a medium. In mathematical terms, it is expressed as the spatial of , with the standard notation being \nabla T or \mathrm{grad}\, T, where T represents as a function of . This notation underscores its nature in multi-dimensional contexts, though the detailed directional properties are addressed separately. In one-dimensional scenarios, such as temperature variation along a straight line (e.g., the x-axis), the gradient simplifies to the scalar \frac{dT}{dx}, representing the change in per unit in that direction. For instance, if decreases linearly from 100°C to 80°C over a 2-meter , the is \frac{dT}{dx} = -10^\circ\mathrm{C/m}. The SI units for the are per meter (K/m) or degrees per meter (°C/m), applicable to both one- and multi-dimensional cases, as the scale difference between K and °C does not affect the magnitude. The concept of the temperature gradient originated in Joseph Fourier's seminal 1822 work, Théorie analytique de la chaleur, where he introduced it as a key element in describing heat flow through conduction. Over time, it evolved into a foundational principle in , underpinning analyses of , , and energy transport across physical systems.

Vector Representation

The temperature gradient, denoted as \nabla T, is a vector field that quantifies the spatial variation of temperature T in three-dimensional space. In Cartesian coordinates (x, y, z), it is defined as \nabla T = \left( \frac{\partial T}{\partial x}, \frac{\partial T}{\partial y}, \frac{\partial T}{\partial z} \right), where each component represents the partial derivative of temperature with respect to the respective coordinate. This vector formulation arises from the general definition of the gradient operator applied to the scalar temperature field, capturing both the rate and direction of temperature change. The of the gradient, |\nabla T|, measures the steepness of the variation and is given by |\nabla T| = \sqrt{ \left( \frac{\partial T}{\partial x} \right)^2 + \left( \frac{\partial T}{\partial y} \right)^2 + \left( \frac{\partial T}{\partial z} \right)^2 }, which is the of the . Directionally, \nabla T points in the direction of the steepest increase in at a given point, with the indicating the of that increase per distance; conversely, the direction of -\nabla T corresponds to the steepest decrease. In non-Cartesian coordinate systems, the expression for \nabla T adapts to the geometry of the space. For spherical coordinates (r, \theta, \phi), the gradient is \nabla T = \frac{\partial T}{\partial r} \hat{e}_r + \frac{1}{r} \frac{\partial T}{\partial \theta} \hat{e}_\theta + \frac{1}{r \sin \theta} \frac{\partial T}{\partial \phi} \hat{e}_\phi, where the scale factors account for the varying metric. Similarly, in cylindrical coordinates ( \rho, \phi, z ), it becomes \nabla T = \frac{\partial T}{\partial \rho} \hat{e}_\rho + \frac{1}{\rho} \frac{\partial T}{\partial \phi} \hat{e}_\phi + \frac{\partial T}{\partial z} \hat{e}_z. Transformations between coordinate systems, such as from Cartesian to spherical, involve the chain rule to express partial derivatives in terms of the new variables; for instance, \frac{\partial T}{\partial x} = \frac{\partial T}{\partial r} \frac{\partial r}{\partial x} + \frac{\partial T}{\partial \theta} \frac{\partial \theta}{\partial x} + \frac{\partial T}{\partial \phi} \frac{\partial \phi}{\partial x}, ensuring the vector components remain consistent under orthogonal curvilinear transformations. The temperature gradient is always perpendicular to isothermal surfaces, which are level sets where T is constant. This orthogonality follows from the fact that along such surfaces, the directional derivative of T in the tangential direction is zero, so \nabla T must align normal to the surface to represent the sole direction of variation.

Physical Mechanisms

Thermal Conduction

Thermal conduction is the process by which heat energy transfers through a material without bulk motion of the substance, driven by a temperature gradient that causes higher-energy particles to share energy with adjacent lower-energy ones. This mechanism occurs in solids and stationary fluids, where the local temperature difference prompts atomic or molecular vibrations and collisions to propagate from hotter to cooler regions. The fundamental relation governing is Fourier's law, first formulated by in 1822, which states that the vector \mathbf{q} is proportional to the negative gradient of temperature: \mathbf{q} = -k \nabla T, where k is the thermal conductivity coefficient. This law emerges from the applied to a differential in the material. Consider a small with faces to the x-direction; the net heat flow into the volume across opposite faces is \frac{\partial q_x}{\partial x} \Delta x \Delta y \Delta z, and by energy balance, this equals the rate of change of \rho c_p \frac{\partial T}{\partial t} \Delta x \Delta y \Delta z, where \rho is density and c_p is . Substituting Fourier's law yields the heat conduction equation \frac{\partial T}{\partial t} = \alpha \nabla^2 T, with \alpha = k / (\rho c_p); in , the time derivative vanishes, simplifying to \nabla \cdot (k \nabla T) = 0. In steady-state conduction, the temperature profile remains constant over time, resulting in a linear temperature gradient for one-dimensional cases with constant k, such as heat flow through a long metal with fixed temperatures at each end. For instance, if one end of a is maintained at 100°C and the other at 0°C, the steady-state temperature varies linearly along the , with q = k \Delta T / L, where L is the . Transient conduction, by contrast, involves time-varying temperatures, as when the initially heats up, with the gradient evolving according to the full until equilibrium is reached. Thermal conductivity k varies significantly with material properties: in metals like , motion enables high values around 400 W/m·K at , facilitating efficient , while insulators such as exhibit low k on the order of 1 W/m·K due to reliance on in vibrations. These differences arise from microscopic mechanisms—electronic conduction dominates in metals, whereas insulators depend on slower vibrations. Boundary conditions define how temperature gradients behave at material interfaces or surfaces, ensuring continuity of temperature and normal heat flux across perfect contacts to satisfy energy conservation. For example, at an interface between two solids, T_1 = T_2 and k_1 \frac{\partial T_1}{\partial n} = k_2 \frac{\partial T_2}{\partial n}, where n is the normal direction; imperfect contacts may introduce thermal resistance. This formulation parallels Fick's first law of diffusion, \mathbf{J} = -D \nabla c, where diffusive flux \mathbf{J} is analogous to heat flux, with diffusion coefficient D playing the role of k, highlighting the shared phenomenological basis for transport processes.

Convection and Advection

In , refers to the transfer of through the bulk motion of a , where warmer, less dense rises and cooler, denser sinks, enhancing heat transport beyond what occurs in stationary media. This process is quantified by , which states that the convective q_{\text{conv}} is proportional to the difference between the surface T_s and the free-stream T_\infty, expressed as q_{\text{conv}} = h (T_s - T_\infty), where h is the convective with units of W/m²K. is distinguished into (or free) , driven solely by forces arising from -induced density variations, and forced , where an external mechanism such as a or imposes motion to augment . Advection, in contrast, describes the passive transport of (or any scalar property) by the mean of the without involving diffusive effects or buoyancy-driven instabilities. In the absence of , this process is governed by the equation \frac{\partial T}{\partial t} + \mathbf{u} \cdot \nabla T = 0, where T is and \mathbf{u} is the vector, indicating that temperature changes result purely from the of existing thermal distributions by the flow. While inherently couples to fluid motion through gradients, isolates the directional transport mechanism, serving as a foundational component in more complex models that include or other forces. A key phenomenon illustrating the onset of convection from an initially stable is , where a layer heated from below develops when the adverse vertical exceeds a critical . This arises as overcomes viscous dissipation and diffusion, leading to organized convective cells or rolls that redistribute heat more efficiently. The transition is characterized by the \text{Ra} = \frac{g \beta \Delta T L^3}{\nu \kappa}, a dimensionless parameter balancing -driven forces (g \beta \Delta T, with g as , \beta as the coefficient, and \Delta T as the temperature difference) against viscous (\nu) and diffusive (\kappa) effects over a L; initiates when \text{Ra} surpasses the critical value of approximately 1708 for typical no-slip boundary conditions. In convective flows, gradients are significantly influenced by effects near solid surfaces, where fluid motion slows and thermal diffusion dominates, resulting in sharpened profiles that concentrate . Within the thermal —a thin region adjacent to —the varies rapidly from value to the free-stream condition, with the layer thickness \delta_t scaling inversely with the speed and , thereby intensifying local gradients and elevating the effective convective coefficient h. This sharpening contrasts with uniform gradients in pure conduction and underscores how amplifies disparities at interfaces in practical applications like exchangers.

Atmospheric and Oceanic Contexts

Meteorology

In , temperature gradients in the atmosphere play a crucial role in determining vertical stability and horizontal weather patterns. The vertical temperature gradient, known as the and denoted as Γ = -dT/dz where T is and z is altitude, describes how temperature changes with height. The environmental lapse rate (ELR) measures the actual atmospheric gradient, while the parcel lapse rate refers to the cooling rate of an ascending air parcel. For dry air, the dry adiabatic lapse rate (DALR) is approximately 9.8 K/km, arising from the adiabatic expansion of unsaturated air. In contrast, the moist adiabatic lapse rate (MALR) for saturated air is lower, typically around 6 K/km, due to the release during that offsets some cooling. Atmospheric stability depends on comparisons between these lapse rates. Absolute stability occurs when the ELR is less than the DALR (or more precisely, less than the MALR for saturated conditions), inhibiting vertical motion as a displaced parcel becomes denser than its surroundings. Neutral stability prevails when the ELR equals the DALR for dry air or MALR for moist air, allowing parcels to maintain . Conditional instability arises when the ELR lies between the DALR and MALR; unsaturated parcels remain stable, but saturation triggers due to the lower MALR. These criteria influence formation, thunderstorms, and , with unstable conditions promoting upward motion and . Horizontal temperature gradients are prominent at weather fronts, where sharp contrasts between air masses drive cyclonic activity. Typical gradients across cold or warm fronts range from 10 to 20 per 100 km, creating baroclinicity that fuels mid-latitude cyclones through shear. These zones mark boundaries between contrasting air masses, with steeper gradients enhancing frontogenesis and associated weather like heavy rain or gusty winds. Temperature inversions represent positive vertical gradients (dT/dz > 0), where temperature increases with height, suppressing and trapping pollutants near the surface. Common in subsiding high-pressure systems or nocturnal cooling, inversions limit vertical mixing, leading to elevated levels. In urban areas, the effect exacerbates this by creating a warmer surface layer overlain by cooler air aloft, further confining emissions from vehicles and industry within the inversion layer. Notable examples include smog episodes in cities like , where persistent inversions have historically worsened air quality.

Oceanography

In oceanography, temperature gradients play a crucial role in structuring water masses, driving circulation, and influencing marine ecosystems and global . Vertical temperature gradients are particularly prominent in the form of the , a transition layer where temperature decreases rapidly with depth, typically separating the warmer, wind-mixed surface layer from the colder deep below. This layer acts as a barrier to vertical mixing, limiting exchange and affecting biological productivity. Horizontal temperature gradients, meanwhile, arise from uneven heating and large-scale currents, creating sharp boundaries that steer flows and modulate atmospheric patterns. The exhibits a sharp vertical , often ranging from 0.1 to 1 K per 100 m, which isolates the surface —where temperatures can reach 20–30°C in tropical regions—from the deep ocean's near-constant 2–4°C. In tropical and subtropical waters, a permanent thermocline persists year-round at depths of 100–1,000 m, maintained by consistent heating at the surface and of cold water masses. By contrast, in mid-latitude temperate zones, a seasonal thermocline forms above the permanent one during spring and summer, as surface warming reduces mixing and creates a steeper gradient; this layer deepens or erodes in winter due to enhanced storm-driven turbulence. In polar regions, the thermocline is shallow or absent, as cold temperatures extend uniformly through the . These gradients contribute to , influencing sound propagation, oxygen distribution, and habitat zones for . Horizontal temperature gradients in the ocean are often amplified by major current systems, such as the Gulf Stream, where contrasts can reach up to 13 K per 100 km across its frontal zone, transporting warm subtropical waters northward and moderating climates along eastern North America and Europe. These gradients arise from latitudinal heating differences and Ekman transport, fostering instabilities that generate eddies and meanders, which in turn mix heat and nutrients. Such features not only drive regional weather variability but also contribute to broader heat redistribution, with the Gulf Stream's warm core elevating sea surface temperatures by 5–10°C relative to surrounding waters. Temperature gradients are integral to , the global "" that connects surface and deep currents through differences governed by both (thermo) and (haline). Colder, saltier waters in high latitudes become denser and sink, forming that flows southward, while warmer, less saline surface waters move poleward; this overturning transports heat equivalent to about 1 petawatt, regulating Earth's . gradients from variations—typically 0.1–0.5 kg/m³ over thousands of kilometers—interact with to sustain this slow (centimeters per second) but voluminous circulation, spanning from the to and back to the over centuries. Recent studies as of 2025 indicate an ongoing slowdown of the Atlantic Meridional Overturning Circulation (AMOC), a key component, due to increased freshwater influx from melting ice, with projections suggesting a potential by 2100 that could drastically alter global heat distribution and weather patterns. El Niño events dramatically weaken equatorial Pacific temperature gradients, flattening the east-west sea surface temperature contrast from a normal 5–8 K over 10,000 km to near zero in extreme cases, as warm waters shift eastward and suppress upwelling. The 1997–98 event, one of the strongest on record, saw eastern Pacific sea surface temperatures rise by over 4 K above average, deepening the thermocline by 50–100 m and reducing the vertical gradient, which persisted into mid-1998. More recently, the 2023–2024 El Niño, also among the five strongest events, peaked with similar anomalies, contributing to record global temperatures in 2023 and widespread weather disruptions including droughts and floods. This reconfiguration halted normal trade winds, triggering global weather anomalies including droughts in Indonesia and floods in South America, while elevating worldwide temperatures by about 0.13 K. Recovery involved Kelvin waves that restored the gradient, highlighting El Niño's role in interannual climate variability.

Geological and Terrestrial Applications

Geothermal Gradient

The refers to the rate of increase in with depth in and , typically driven by internal heat sources and conductive . In , the average is approximately 25–30 °C/km near the surface. This gradient generally decreases with increasing depth due to changes in rock thermal conductivity, as deeper rocks become less porous and thus more conductive, allowing heat to flow more efficiently for a given flux. The primary heat sources sustaining the geothermal gradient include radiogenic decay of elements such as , , and , which contribute roughly 50% of Earth's surface . The remaining flux arises from primordial heat retained since Earth's formation and from core solidification. Conductive heat flow in the subsurface is described by Fourier's law: q = -k \frac{dT}{dz} where q is the (W/m²), k is the thermal conductivity of the rock (W/m·K), and \frac{dT}{dz} is the (K/m). This relationship highlights how variations in rock properties and heat production influence temperature profiles. Geothermal gradients exhibit significant regional variations, often higher in tectonically active or volcanic areas; for instance, in , gradients can reach 80–100 °C/km due to proximity to mantle hotspots and rift zones. These gradients are measured primarily through logs in boreholes, where thermistors record downhole temperatures to compute the rate of change with depth. Such measurements are essential for assessing potential and understanding geodynamic processes. The plays a critical role in driving , where temperature differences create buoyancy forces that facilitate the upwelling of hot material and the overall engine of . Historically, the foundational understanding of subsurface heat flow emerged from Joseph Fourier's 1822 work on heat conduction, with 19th-century surveys in mines and wells providing early empirical data on gradients, later refined by Lord Kelvin's applications to Earth's thermal history.

Weathering Processes

Temperature gradients at Earth's surface play a crucial role in initiating and accelerating processes by inducing mechanical stresses and facilitating chemical reactions in rocks. These gradients arise from diurnal, seasonal, or episodic variations in , leading to , , or changes that contribute to the physical and chemical breakdown of rock materials. In particular, such gradients are prominent in exposed surface environments where rocks interact directly with atmospheric conditions, promoting the formation of and over geological timescales. Thermal expansion weathering occurs when temperature gradients cause uneven heating or cooling across a rock's surface, resulting in internal stresses that generate microcracks. Daily temperature fluctuations of 10-20 between day and night, common in arid regions, drive this process by expanding the outer layers of rocks more than the cooler interior, leading to spalling or granular disintegration. For instance, in outcrops, repeated cycles of such gradients propagate fractures along boundaries, enhancing rock breakdown over time. This mechanism is particularly effective in deserts, where clear skies amplify diurnal swings and minimal exposes rocks to intense solar radiation. Freeze-thaw cycles represent another weathering pathway intensified by gradients, especially in temperate zones where rocks experience repeated crossings of the 0°C threshold. During freezing, in pore spaces or fractures expands by about 9% upon formation, but the key gradient forms across the - , where subzero temperatures in contrast with , generating wedging pressures up to 200 that exceed rock tensile strength. This process accelerates in regions with frequent winter thaws, such as mid-latitude mountains, where diurnal gradients promote and refreezing, progressively widening cracks and dislodging fragments. Temperature gradients also enhance chemical by driving and reactions in minerals, particularly clays, which alter rock structure and . In environments with fluctuating and , gradients promote the uptake or loss of water molecules in expandable clays like , leading to swelling and shrinkage that weaken the rock matrix. For example, diurnal gradients in semi-arid areas can induce cyclic of clay interlayers, facilitating and dissolution of surrounding silicates, thereby increasing and potential. This effect is more pronounced in humid climates, where higher overall temperatures accelerate alongside gradients. Quantitatively, the thermal stress generated by these gradients can be expressed as \sigma = \alpha E \Delta T, where \sigma is the thermal stress, \alpha is the linear thermal expansion coefficient (typically $5-10 \times 10^{-6} K^{-1} for rocks like granite), E is Young's modulus (around 50-100 GPa), and \Delta T is the temperature change. In arid climates, larger \Delta T values (e.g., 20-30 K daily) produce stresses exceeding 10-20 MPa, sufficient to initiate cracking in surface rocks, whereas in humid areas, smaller gradients (5-10 K) combine with moisture to favor chemical over purely mechanical breakdown, resulting in slower but more pervasive alteration.

Engineering and Human Environments

Indoor Thermal Gradients

Indoor thermal gradients refer to spatial variations in temperature within enclosed building spaces, primarily driven by currents that cause warm air to rise and cooler air to settle. Vertical gradients typically range from 1 to 3 / in standard rooms, resulting from natural effects where heated air ascends, creating a warmer layer near the and cooler conditions at floor level. The American Society of Heating, Refrigerating and Air-Conditioning Engineers () Standard 55 (as of the 2023 edition) specifies that the vertical air temperature difference between head height (1.1 ) and ankle level (0.1 ) should not exceed 3 for seated individuals to ensure occupant comfort, corresponding to a gradient limit of 3 /. Exceeding this threshold can lead to local discomfort, such as , though whole-body dissatisfaction remains low (under 10%) even at gradients up to 8 / under neutral thermal conditions. Recent research suggests that gradients up to 4–5 / may be acceptable in such conditions with low predicted percentage dissatisfied. Horizontal gradients often arise near building envelopes, such as windows or doors, where drafts introduce cooler outdoor air, resulting in temperature differences of up to 3 across a room section. These variations are exacerbated by air infiltration, creating uneven zones that affect overall indoor uniformity. In enclosed spaces, such convection-driven patterns contribute to these horizontal disparities without relying on external . Health impacts from indoor thermal gradients primarily manifest as thermal discomfort and potential respiratory aggravation, with draughts from steep gradients causing sensations of chill that may trigger symptoms or exacerbate chronic conditions by inflaming airways. Post-1970s energy crises, building codes evolved to prioritize through tighter envelopes, inadvertently heightening gradient-related issues like draughts, which prompted updated standards emphasizing comfort limits to mitigate health risks. guidelines address draught risk by limiting air speeds to 0.15 m/s at low temperatures to prevent local cold discomfort. Measurement of indoor thermal gradients commonly employs thermocouples for precise point-wise air profiling along vertical or horizontal traverses, offering high accuracy for convection-influenced zones. () imaging provides non-contact mapping of surface and air temperatures, enabling visualization of gradient patterns across rooms, though it requires for to quantify absolute values.

Thermal Insulation Design

Thermal insulation design in focuses on minimizing gradients across building envelopes to reduce heat loss, thereby enhancing and maintaining structural integrity. By incorporating materials with high thermal resistance, designs lower the conductive while preserving a stable indoor against external variations. This approach is essential for achieving compliance with building codes that emphasize reduced . The core principle of thermal insulation involves increasing the R-value, a measure of thermal resistance defined as the ratio of temperature difference to , with units of m²·K/W; higher R-values correspond to greater resistance to heat flow and thus reduced temperature gradients for a given thermal load. Insulation materials trap still air or use low-conductivity structures to impede conduction, the primary mode of across solid barriers, effectively lowering the overall heat loss from heated or cooled spaces. For instance, proper insulation can reduce heating energy demands by 10–30% in cold climates by flattening the temperature profile through walls and roofs. Common insulation materials are selected based on their thermal conductivity k, where lower k values enable higher R-values per unit thickness. Fiberglass, composed of fine glass fibers, has a thermal conductivity of approximately 0.04 W/m·K and provides an R-value of 2.5 to 3.8 per inch, making it suitable for batt installations in attics and walls due to its affordability and ease of handling. Rigid foam boards, such as polyurethane or polystyrene, offer superior performance with k around 0.02 to 0.03 W/m·K and R-values of 5 to 7 per inch, ideal for continuous exterior applications where space is limited. These materials are chosen for specific projects by balancing k, resistance, and ratings to optimize the effective reduction without compromising durability. Design calculations for insulation assemblies rely on steady-state models to predict and minimize gradients in . The heat flux through a wall is given by q = \frac{\Delta T}{R_{\text{total}}}, where \Delta T is the temperature difference across the assembly and R_{\text{total}} is the sum of individual layer resistances; this equation ensures that added layers proportionally reduce and the associated internal gradient. The U-factor, or overall , is calculated as U = \frac{1}{R_{\text{total}}} in W/m²·K, providing a metric for entire assemblies including framing and air films—values below 0.3 W/m²·K are targeted for energy-efficient to limit heat loss to under 20 W/m² in moderate climates. These computations guide layer sequencing to avoid thermal bridges, such as metal studs, which can increase effective U-factors by 20-50%. Advanced techniques, such as vacuum insulation panels (VIPs), push the boundaries of gradient minimization by evacuating air to achieve thermal conductivities as low as 4 mW/m·K, resulting in near-zero temperature drops across panels as thin as 20-40 mm and U-values below 0.2 W/m²·K. These panels, encased in metallic or polymeric barriers, suppress gaseous conduction and , offering up to five times the insulation of traditional foams in compact spaces like retrofitted walls. VIPs have been integrated into projects adhering to standards, established in 2000, which award credits in the Energy and Atmosphere category for enhanced thermal envelopes that reduce building energy use by at least 10-20% over baseline codes. Such applications support by enabling thinner walls while maintaining low indoor thermal gradients.

References

  1. [1]
    8.5 Gradients: How to Find Them | METEO 300 - Dutton Institute
    For instance, the temperature gradient gives the maximum amount of temperature change in space and the direction of that maximum temperature change. Thus, a ...
  2. [2]
    [PDF] HEAT CONDUCTION EQUATION
    The driving force for any form of heat transfer is the temperature difference, and the larger the temperature difference, the larger the rate of heat transfer.
  3. [3]
    Temperature and heat
    Heat is said to pass from one system to another if the two systems are initially at different temperatures and are then placed in contact with each other.
  4. [4]
    [PDF] 12.002 Physics and Chemistry of the Earth and Terrestrial Planets
    This means that if the thermal gradient is greater than the adiabatic gradient, there is energy available to drive vertical motion of rock packages in the.
  5. [5]
    Geothermal Glossary | Department of Energy
    Geothermal Gradient​​ The rate of temperature increase in the Earth as a function of depth. Temperature increases an average of 1° Fahrenheit for every 75 feet ...
  6. [6]
    Engineering: The challenge of temperature - The Open University
    Clearly, very steep temperature gradients make for high stress, and the gradient depends on how much energy has to flow from the inside to the surface and how ...
  7. [7]
    Temperature Gradients - an overview | ScienceDirect Topics
    Temperature gradient is defined as the direction and rate of the most rapid temperature change within a sample, indicating the unequal distribution of ...
  8. [8]
    Fourier's Law of Thermal Conduction | Calculation - Nuclear Power
    This law states that the time rate of heat transfer through a material is proportional to the negative gradient in the temperature and the area.
  9. [9]
    Define temperature gradient. - Heat Transfer - CK-12
    Mathematically, it is expressed as a vector quantity and is denoted by the symbol ∇ T , where T refers to the temperature. ... temperature changes per unit length ...Missing: notation | Show results with:notation
  10. [10]
    Thermal Gradient: Definition & Calculation - Lesson - Study.com
    Thermal gradient is defined as the ratio of the temperature difference and the distance between two points.
  11. [11]
    Temperature Gradient Conversion - Units Converters
    The temperature gradient is a dimensional quantity expressed in units of degrees (on a particular temperature scale) per unit length.. There are various units ...Missing: notation | Show results with:notation<|control11|><|separator|>
  12. [12]
    Thermal conductivity through the 19th century - Physics Today
    Aug 1, 2010 · From 1807 to 1811, Joseph Fourier conducted experiments and devised mathematical techniques that together yielded the first estimate of a ...
  13. [13]
    [PDF] Lecture 5 Vector Operators: Grad, Div and Curl
    dimensions, then its gradient at any point is defined in Cartesian co-ordinates by. gradU = ∂U. ∂x. ˆı +. ∂U. ∂y. ˆ +. ∂U. ∂z k . (5.1). 1. Page 2. 5/2.
  14. [14]
    [PDF] Review of Vector Analysis in Cartesian Coordinates - Research
    Scalars are usually represented by italic letters: for example, T for temperature or p for pressure. Vector: A quantity that has a magnitude and a direction.
  15. [15]
    [PDF] 2.2+ Gradient extras
    Mar 14, 2008 · Gradient in cartesian coordinates. The position vector reads r = xx + yy + zz. Picking dr = x dx, dr = ydy and dr = zdz in (1), respectively ...
  16. [16]
    14.5 Directional Derivatives
    In other words, the gradient ∇f points in the direction of steepest ascent of the surface, and |∇f| is the slope in that direction.
  17. [17]
    2.7 Directional Derivatives and the Gradient
    The principal interpretation of d f d x ( a ) is the rate of change of , f ( x ) , per unit change of , x , at . x = a . The natural analog of this ...
  18. [18]
    [PDF] 19 Lecture: pp. 130-133
    Gradient in cylindrical and spherical coord. Use transforms and chain rule ... do the same thing for spherical coordinates: φ φ θ θ θ. ˆ sin. 1. ˆ. 1ˆ.
  19. [19]
    [PDF] 18.02 Multivariable Calculus - MIT OpenCourseWare
    For fixed φ, θ we are slicing our region by rays straight out of the origin; ρ ranges from its value on the plane z = 1/. √. 2 to its value on the sphere ρ = 1.
  20. [20]
    2 Differential Calculus of Vector Fields - Feynman Lectures - Caltech
    For a temperature field the contours are called “isothermal surfaces” or isotherms. Figure 2–1 illustrates a temperature field and shows the dependence of T on ...
  21. [21]
    Joseph Fourier, Théorie analytique de la chaleur (1822)
    Joseph Fourier's book, Théorie Analytique de la Chaleur contains the first extended mathematical account of heat diffusion.
  22. [22]
    Heat Transfer: Conservation of Energy - COMSOL
    Jun 29, 2018 · The final step is to define the conduction heat flux vector, , using Fourier's law of conduction, , where is the thermal conductivity. The ...
  23. [23]
    [PDF] The 1-D Heat Equation
    Sep 8, 2006 · ... metal rod with non-uniform temperature, heat (thermal energy) is transferred from regions of higher temperature to regions of lower temperature.<|control11|><|separator|>
  24. [24]
    Steady-State Vs Transient Thermal Analysis In FEA
    Steady-state thermal analysis has no meaningful time scale, while transient analysis considers time significantly. Use steady-state when time behavior is ...
  25. [25]
    Thermal Conductivity - HyperPhysics
    Thermal Conductivity ; Silver. 1.01. 406.0 ; Copper. 0.99. 385.0 ; Gold ... 314 ; Brass ... 109.0 ; Aluminum. 0.50. 205.0.
  26. [26]
    [PDF] thermal conductivity of polymers, glasses & ceramics by MDSC, TA ...
    Solid conductors (such as metals) typically have thermal conductivities in the range of 10 to 400 W/°C m while insulators (such as polymers, glasses and ...
  27. [27]
    Interface Boundary Condition - Heat Transfer - Nuclear Power
    The heat conduction equation is a partial differential equation that describes heat distribution (or the temperature field) in a given body over time.
  28. [28]
    Revised Formulation of Fick's, Fourier's, and Newton's Laws for ...
    Derivations. Fick deduced his first law of diffusion (eq 1) by analogy with Fourier's law of heat conduction (and Ohm's law of electrical conduction). (1) ...Introduction · Results and Discussion · Conclusions · References
  29. [29]
    17. Convective Heat Transfer - MIT
    The convective heat transfer coefficient is defined by. $\displaystyle \dot{q} = \frac{\dot{Q, (17..2). Equation (17.2) is often called Newton's Law of Cooling.
  30. [30]
    Advection Term - an overview | ScienceDirect Topics
    Advection refers to the transport of molecules (solute) by the movement of bulk fluid (solvent). It is affected by the sum of the i) surface force active on ...
  31. [31]
    [PDF] Fluid Dynamics - UCSD CSE
    • Advection is the transport of a fluid property through the macroscopic ... • The advection equation specifies a scalar field. ds. dt which is the rate ...
  32. [32]
    [PDF] Fluid Dynamics and Rayleigh-Benard Convection
    Oct 24, 2022 · Ra is the Rayleigh number. A detailed stability calculation reveals that the critical constant is 1708. Our derivation of the Rayleigh ...
  33. [33]
    [PDF] Velocity & Thermal Boundary Layers
    The temperature gradients result in heat fluxes (Fourier's Law) and, therefore, heat transfer between the fluid and the wall. Steady-state thermal boundary ...
  34. [34]
    Lapse Rates | Learning Weather at Penn State Meteorology
    The lapse rate is the change in temperature with altitude in any given layer of air. As a general rule, the greater the decrease in temperature with height, ...Missing: formula | Show results with:formula
  35. [35]
    [PDF] Atmospheric Stability
    •The environmental lapse rate is smaller than the dry but larger than the moist adiabatic lapse rate. Page 16. Neutral Stability. •Dry air is neutrally stable ...
  36. [36]
    Global Climatologies of Fronts, Airmass Boundaries, and Airstream ...
    Feb 1, 2019 · In this case, the threshold of 2.0 K (100 km)−1 most closely matches the operational surface analysis by the German weather service Deutscher ...
  37. [37]
    Inversion | US EPA
    Oct 1, 2025 · In a temperature inversion, the situation “inverts,” and cold air at the surface gets trapped under a layer of warmer air.Missing: trapping | Show results with:trapping
  38. [38]
    [PDF] effects of the urban heat island upon meteorological parameters
    May 15, 1972 · This inversion in turn leads to higher concentration of pollutants at lower levels - a vicious cycle in which air pollution is further enhanced.
  39. [39]
    How You Control a Day in Smog City 2
    Inversion A temperature inversion is a layer of warm air above the ground that traps particle pollution and ground-level ozone below it. This "lid" prevents ...
  40. [40]
    What is a thermocline? - NOAA's National Ocean Service
    Jun 16, 2024 · A thermocline is the transition layer between warmer mixed water at the ocean's surface and cooler deep water below.
  41. [41]
    6.2 Temperature – Introduction to Oceanography
    Below the mixed layer there is a rapid decline in temperature over a fairly narrow increase in depth. This is called the thermocline. Below the thermocline the ...
  42. [42]
    [PDF] 1 Lecture 3: Temperature, Salinity, Density and Ocean Circulation
    In mid latitudes a seasonal thermocline often starts to develop in the spring above the permanent thermocline, as surface temperature rise and mixing by.
  43. [43]
    Seasonal Variability of Horizontal Gradients in the North Atlantic ...
    Keywords. frontal zone, horizontal gradients, temperature gradient ... Here, the front with temperature gradients exceeding 2 °C/100 km narrows in ...Temperature Frontal Zones · Density Frontal Zones · Seasonal Variability Of The...
  44. [44]
    [PDF] FLORIDA CURRENT, GULF STREAM, AND LABRADOR CURRENT
    The maximum temperature gradient across the Stream is located near a depth of 500m and amounts to around 101C. Maximum surface temperature gradients occur in ...
  45. [45]
    Thermohaline Circulation - Fact Sheet by Stefan Rahmstorf
    The salinity and temperature differences arise from heating/cooling at the sea surface and from the surface freshwater fluxes (evaporation and sea ice formation ...<|separator|>
  46. [46]
    9.8 Thermohaline Circulation – Introduction to Oceanography
    A temperature-salinity (T-S) diagram is used to examine how temperature, salinity, and density change with depth, and to identify the vertical structure of ...Missing: gradients | Show results with:gradients
  47. [47]
    Genesis and Evolution of the 1997-98 El Niño - NOAA/PMEL
    Termination of the 1997–98 El Niño was preconditioned by low-frequency ocean wave processes which elevated the thermocline in the central and eastern Pacific.
  48. [48]
    The Termination of the 1997–98 El Niño. Part I - AMS Journals
    Warm eastern equatorial Pacific (EEqP) sea surface temperature anomalies (SSTAs) exceeded 4°C at the event peak and lasted well into boreal spring of 1998, even ...
  49. [49]
    Geothermal Gradient - an overview | ScienceDirect Topics
    Geothermal gradients in continental areas are commonly about 25 °C km−1, but are lower for Archean cratons and much higher in areas of Cenozoic volcanism and ...
  50. [50]
    Potential for deep geothermal energy in Scotland: study volume 2
    Nov 13, 2013 · The average geothermal gradient in continental areas is 25-30 ºC/km . 2.1.4 Factors affecting the supply and transfer of heat in the crust. Heat ...
  51. [51]
    Should you rely on geothermal gradients? - GeoExpro
    May 7, 2024 · ... gradient generally decreases with depth. Why? Because the rocks are less porous and therefore more conductive with increasing depth. This is ...
  52. [52]
    Radiogenic heating sustains long-lived volcanism and magnetic ...
    Roughly half of the heat that Earth loses through its surface today comes from the three long-lived, heat-producing elements (potassium, thorium, and uranium).
  53. [53]
    Heat flow in Wyoming
    Heat flow is related to the geothermal gradient in a purely conductive medium: q=-k dT/dz. where k is the thermal conductivity of the material and is free to ...Missing: formula | Show results with:formula
  54. [54]
    [PDF] Geothermal exploitation in Iceland – Success and Challenges
    The typical background values of the temperature gradient in Iceland is 80-100°C/km at the boarder of the volcanic rift zone to 40-50°C/km in the oldest crust ...
  55. [55]
    [PDF] Determination of geothermal gradient from borehole temperature ...
    Geothermal gradient is a useful parameter for constraining models of heat flow and petroleum generation in sedimentary basins.
  56. [56]
    9.2 The Temperature of Earth's Interior – Physical Geology
    As we'll see in Chapter 10, a convecting mantle is an key feature of plate tectonics. The convection of the mantle is a product of the upward transfer of heat ...
  57. [57]
    Kelvin, Perry and the Age of the Earth | American Scientist
    By inverting—or turning inside out—Fourier's diffusion calculation, Kelvin could solve for the age of the Earth in terms of the geothermal gradient at the ...Missing: surveys | Show results with:surveys
  58. [58]
    Weathering and Soils - Tulane University
    Sep 10, 2015 · Thermal Expansion - Although daily heating and cooling of rocks do not seem to have an effect, sudden exposure to high temperature, such as in ...Missing: gradient | Show results with:gradient
  59. [59]
    Chapter 8 - Weathering & Erosion - GotBooks.MiraCosta.edu
    Thermal expansion—expansion and contraction caused by daily heating and cooling, particularly effective in arid environments. Heat from wildfires can also ...
  60. [60]
    [PDF] Mechanical weathering and rock erosion by climate-dependent ...
    Our calculations also indicate that climate strongly influences subcritical cracking—and thus rock weathering rates—irrespective of the source of the stress ( ...<|control11|><|separator|>
  61. [61]
    [PDF] Rockfall triggering by cyclic thermal stressing of exfoliation fractures
    Mar 28, 2016 · The role of thermal effects (temperature and insolation) on initiating rock deformation, where rock surfaces expand, contract, and eventually ...
  62. [62]
    Water–Rock Interaction and Freeze–Thaw Cycles as Drivers of Acid ...
    Nov 13, 2024 · Freeze–thaw action, also referred to as frost weathering, is the dominant physical process operating wherever rocks experience temperature ...
  63. [63]
    [PDF] Climatic controls on frost cracking and implications for the evolution ...
    Jun 14, 2007 · Freeze/thaw weathering requires temperature oscil- lations about 0°C, whereas segregation ice grows at a temperature range that maximizes the ...
  64. [64]
    Chapter 4 Weathering and Erosion – *Introduction to World Geography
    Frost wedging is most effective in mountainous climates. In warm areas where freezing is infrequent, in very cold areas where thawing is infrequent, or in arid ...
  65. [65]
    [PDF] Chemical Weathering of Basalts and Andesites
    These observa- tions support the theory of clay-mineral formation by dehydration and crystallization of allophane. Direct alteration of primary minerals to ...
  66. [66]
    Clay mineral formation under oxidized conditions and implications ...
    Nov 1, 2017 · Fe-rich clay minerals, including nontronite were rapidly synthesized within 2 days at 150 °C, following 1 day at room temperature and within 60 ...
  67. [67]
    [PDF] Chapter 3 Weathering and Soils - Find People
    Chemical weathering involves reactions that change primary rock-forming minerals into secondary minerals, such as clays. In the process, some elements are lost.
  68. [68]
    [PDF] Impact of Thermo-Mechanical Stimulation on the Reservoir Rocks of ...
    Feb 12, 2018 · σt = induced tensile thermal stress (MPa) α = linear expansion coefficient (m/ (m K)) E = Young's modulus (MPa) ΔT = temperature difference (°C ...
  69. [69]
    Thermal effect of temperature gradient in a field environment ...
    Nominal vertical air temperature gradients between 0.1 and 1.1 m heights were 1, 3 and 5 K/m while nominal room air temperatures at 0.6 m height were 20, 23 and ...
  70. [70]
    Thermal Environmental Conditions for Human Occupancy - ashrae
    Standard 55 is designed to specify those combinations of factors that result in satisfactory thermal conditions for a majority of occupants.
  71. [71]
    Predicted percentage dissatisfied with vertical temperature gradient
    Aug 1, 2020 · The limit of 3 °C/m between head and feet is stipulated in ASHRAE 55 and ISO 7730 (Category C buildings), based on the work of Olesen et al. [3] ...
  72. [72]
    [PDF] Thermal Conditions in a Simulated Office Environment with ...
    Due to the horizontal temperature difference, the air and operative temperature near the window was about 0.4-0.9 °C higher than room design temperature (in the ...
  73. [73]
    [PDF] INDOOR AIRFLOW WITH COOLING PANEL AND RADIATIVE ...
    The horizontal temperature difference along the x direction, observed only in the region near the floor, as shown in Figure 7a, is about 3 °C. The con ...
  74. [74]
    Low indoor temperatures and insulation - WHO Housing and ... - NCBI
    Cold air inflames lungs and inhibits circulation, increasing the risk of respiratory conditions, such as asthma attacks or symptoms, worsening of chronic ...Missing: draughts gradients discomfort
  75. [75]
    Climate Change and Indoor Air Quality: Lessons from the Energy ...
    Jul 24, 2013 · First, energy conservation in housing and other buildings is imperative, but can affect indoor environmental quality both negatively and positively.
  76. [76]
    [PDF] A Basic Guide to Thermocouple Measurements - Texas Instruments
    Choosing a thermocouple often is a function of the measurement temperature range required in the application. Other considerations include the temperature ...Missing: indoor | Show results with:indoor
  77. [77]
    Using quantitative infrared thermography to determine indoor air ...
    Infrared thermography (IR) has proven to be an effective alternative method for determining room temperature within buildings. Nevertheless, IR is limited ...
  78. [78]
    Insulation | Department of Energy
    An insulating material's resistance to conductive heat flow is measured or rated in terms of its thermal resistance or R-value -- the higher the R-value, the ...Types of Insulation · Where to Insulate in a Home · Adding Insulation
  79. [79]
    Building Science Introduction - Heat Flow
    Jan 1, 2011 · The ability of a material to resist heat flow is measured in R-Value. R-Value is the inverse of U-Factor (R=1/U). The higher the R-Value, the ...
  80. [80]
    [PDF] Insulating with Exterior Rigid Foam
    Oct 14, 2014 · Principle of Thermal Bridging. •Thermal Conductivity @ ~ 70o F. – Wood (pine) = 0.14 (W/mK). – Fiberglass insulation = 0.04 (W/mK). – Air = ...
  81. [81]
    5 Most Common Thermal Insulation Materials - Thermaxx Jackets
    Oct 20, 2021 · Fiberglass is an excellent non-flammable insulation material, with R-values ranging from R-2.9 to R-3.8 per inch.Missing: conductivity | Show results with:conductivity
  82. [82]
    A review of the challenges posed by the use of vacuum panels in ...
    Jan 1, 2020 · Due to their low thermal conductivity, high insulation levels can be achieved with thinner walls than is possible with conventional thermal ...2. Vacuum Insulation Panels... · 2.2. Case Studies · 3. Challenges Of Vip Use In...
  83. [83]
    LEED v4: Building Design + Construction Guide - USGBC
    This reference guide is designed to elaborate upon and work in conjunction with the rating system. Written by expert users of LEED, it serves as a roadmap.