Fact-checked by Grok 2 weeks ago

Temperature

Temperature is a that serves as a measure of the of the microscopic particles—such as atoms and molecules—within a substance, reflecting the degree of hotness or coldness relative to a reference point. In thermodynamic terms, it quantifies the total associated with the random motion of these particles, providing an where zero corresponds to the absence of thermal motion. This concept underpins , where two systems in contact reach the same temperature and no net occurs. The most widely used temperature scales include the Celsius (°C), (°F), and (K) scales, each defined by specific reference points such as the freezing and boiling points of under standard . The scale sets the freezing point of at 0°C and the boiling point at 100°C, making it intuitive for everyday applications. In contrast, the scale, common in the United States, assigns 32°F to 's freezing point and 212°F to its boiling point, resulting in finer degree intervals for human-perceived temperatures. The scale, the unit for , starts at (0 K, equivalent to -273.15°C), where molecular motion theoretically ceases, and is essential for scientific calculations involving gases and absolute energy measures. Temperature is measured using various thermometers that exploit physical properties changing with , such as the of or the of metals. Traditional liquid-in-glass thermometers, often using mercury or , rely on the volume of the to indicate temperature on a calibrated . methods, including thermocouples (which generate voltage from temperature-induced metal junctions) and temperature detectors (RTDs, which measure changes in electrical ), offer higher precision and are used in industrial, medical, and . For instance, air temperature is accurately gauged with thermometers capable of resolutions down to fractions of a . In physics and related fields, temperature plays a central role in , influencing phenomena like , phase changes, and chemical reactions, while its variations drive patterns, biological processes, and material behaviors.

Core Concepts

Definition

Temperature is fundamentally defined through the , which states that if two systems are separately in with a third system, then they are in with each other. This law establishes temperature as the property shared by systems in , where no net occurs between them due to the absence of a . Thermal equilibrium thus serves as the empirical basis for measuring and comparing temperatures across thermodynamic systems. As an intensive property, temperature does not depend on the size or amount of the but reflects the average of its microscopic particles, such as atoms or molecules. In thermodynamic contexts, it quantifies the tendency of a to exchange with its surroundings via . Importantly, temperature must be distinguished from : temperature is a characterizing the system's distribution, while is the process of energy driven by a temperature difference between systems. Objects possess temperature but not in isolation; arises only during . In the specific case of an , temperature relates directly to the average translational per , given by the equation \frac{3}{2} k T = \frac{1}{2} m \langle v^2 \rangle, where T is the temperature, k is Boltzmann's , m is the , and \langle v^2 \rangle is the speed of the molecules. This proportionality underscores temperature's role as a macroscopic measure of microscopic agitation, applicable under conditions where intermolecular forces are negligible.

Equilibrium and Non-Equilibrium

In , a achieves a temperature throughout when it is isolated or in contact with a heat bath, such that no net flows between parts of the or with the surroundings, consistent with the . This state implies that macroscopic properties like and are also , and the remains unchanged over time unless perturbed. For instance, two objects in reach the same temperature when is attained, ceasing any energy exchange. In non-equilibrium steady states, temperature is not uniform; instead, persistent gradients drive continuous flow, as seen in steady-state conduction where the properties remain over time despite the imbalance. Here, the overall energy input equals output, maintaining a , but local temperatures vary spatially—for example, in a with fixed and ends, a linear temperature profile develops along its length. Such states are analyzed using extended irreversible , where temperature becomes a local quantity adjusted for dissipative effects like . Non-steady states, or transient conditions, feature time-varying temperatures as the system evolves toward or another steady configuration, such as during initial flow in an insulated body suddenly exposed to a temperature difference. In these scenarios, temperature profiles change dynamically, with diffusion governed by the unsteady , leading to temporary gradients that diminish over time. Unlike steady states, no constant flux persists, and the system's departure from uniformity is both spatial and temporal. Local thermodynamic equilibrium (LTE) approximates equilibrium conditions in specific regions of an otherwise non-equilibrium system, where collision rates among particles are high enough to maintain Maxwell-Boltzmann distributions locally, despite global imbalances. This assumption is valid in dense plasmas or stellar atmospheres when radiative processes are negligible compared to collisions, allowing temperature to be defined via local energy equipartition. For example, in the solar photosphere, LTE holds over small scales where temperature gradients are shallow, enabling equilibrium statistical descriptions amid outward energy transport. An axiomatic approach in treats temperature as a functional of the system's state variables, extending definitions to include fluxes and gradients for consistency with the second law. In extended irreversible , this involves a non-equilibrium depending on and , yielding a local temperature via the ∂s/∂u, where deviations from temperature scale with , such as θ ≈ T (1 - α q²) for small perturbations. This ensures thermodynamic relations hold formally, as in the generalized Gibbs ds = θ⁻¹ du + dissipative terms.

Scales and Units

Empirical Scales

Empirical temperature scales are defined based on reproducible physical phenomena, such as the freezing and points of under standard atmospheric pressure, without reference to any underlying molecular or thermodynamic . These scales emerged in the as practical tools for temperature variations through thermometric fluids like mercury or , prioritizing human-relevant reference points over universal absolutes./08%3A_Entropy_Production_and_Accounting/8.02%3A_Empirical_and_Thermodynamic_Temperature) The scale, proposed by German physicist in , was one of the earliest standardized empirical scales. Fahrenheit calibrated his mercury-in-glass thermometers using three fixed points: the temperature of a brine mixture of ice, , and at 0°F; the freezing point of at 32°F; and the average , initially set near 96°F but later adjusted to 98.6°F. The of was determined to be 212°F, establishing 180 divisions between freezing and boiling. This scale's finer graduations (using 1/96 of the human body temperature interval as a degree) aimed for precision in meteorological and medical applications. In 1731, French naturalist introduced his scale based on the volumetric expansion of in a tube. He defined 0°R as the freezing point of and 80°R as the under standard conditions, dividing the interval into 80 equal parts to reflect alcohol's expansion coefficient, which he measured as expanding by 1/1000 of its volume per degree. This choice made the scale convenient for instruments using alcohol, which expands more than mercury, and it gained popularity in for scientific and industrial uses, such as monitoring processes. The Celsius scale, developed by Swedish astronomer Anders Celsius in 1742, refined earlier proposals by using water's phase changes as fixed points in a decimal system. In his publication Observations of two persistent degrees on a thermometer, Celsius initially proposed 0°C for water's boiling point and 100°C for its freezing point, but this was reversed shortly after his death to the modern convention of 0°C at freezing and 100°C at boiling, dividing the interval into 100 equal degrees. This centigrade (hundred-grade) approach emphasized simplicity and universality for astronomical and everyday measurements, quickly supplanting other scales in scientific contexts. Conversions between these empirical scales account for their differing zero points and degree sizes, derived from the ratios of their intervals between water's freezing and boiling points. The formula to convert Celsius to Fahrenheit is
^\circ\mathrm{F} = ^\circ\mathrm{C} \times \frac{9}{5} + 32
and the inverse is
^\circ\mathrm{C} = (^\circ\mathrm{F} - 32) \times \frac{5}{9}.
For the Réaumur scale, which spans four-fifths the interval of Celsius, the conversions are
^\circ\mathrm{R} = ^\circ\mathrm{C} \times \frac{4}{5}
and
^\circ\mathrm{C} = ^\circ\mathrm{R} \times \frac{5}{4}.
These relations highlight the proportional differences: the Fahrenheit degree is 5/9 of a Celsius degree, while the Réaumur degree is 4/5 of a Celsius degree.
A key limitation of empirical scales is their arbitrary zero points, which are set by convenient references like or rather than any intrinsic physical limit, leading to non-intuitive values for common phenomena across scales. Additionally, intervals are not directly additive between scales without conversion, complicating comparisons; for instance, a 100°R change does not equal a 100°C change in magnitude. These features make empirical scales practical for relative measurements but less suitable for precise scientific calculations involving or ./08%3A_Entropy_Production_and_Accounting/8.02%3A_Empirical_and_Thermodynamic_Temperature)/08%3A_Entropy_Production_and_Accounting/8.02%3A_Empirical_and_Thermodynamic_Temperature)

Absolute Scales

Absolute scales of temperature are defined by an invariant reference point at , the lowest conceivable temperature where thermal motion theoretically ceases, providing a universal foundation independent of specific substances or arbitrary fixed points used in empirical scales. This zero point emerges from extrapolations of gas behavior under constant pressure, as observed in , which posits that the volume of a gas is directly proportional to its absolute temperature. At , marked as 0 on these scales, a system's reaches its minimum, and approaches a minimum value for ideal cases, establishing a physical limit to cooling processes. The , the SI unit of denoted by , anchors its to physical constants rather than material properties alone. Prior to 2019, it was realized through the of , fixed at exactly 273.16 , where coexists in , , and vapor phases in . Since the 2019 redefinition, the is defined by assigning the exact value of the as k = 1.380649 \times 10^{-23} J/, linking temperature directly to the average per degree of freedom in a system. This ensures the scale's invariance and precision in scientific measurements, with intervals equivalent to those of the scale but shifted to start at . The (°R), an absolute counterpart to the scale, employs the same degree size as Fahrenheit but sets at 0 °R. Introduced by Scottish engineer William John Macquorn Rankine in the , it places the freezing point of at 491.67 °R and the at 671.67 °R under standard . Primarily used in English-unit engineering contexts, such as in the United States, it facilitates calculations involving absolute temperatures without negative values. Thermodynamic temperature, the core concept underlying absolute scales, is defined independently of any working substance through the efficiency of reversible heat engines, as established by the . The maximum efficiency \eta of such an engine operating between hot reservoir temperature T_h and cold reservoir temperature T_c is given by \eta = 1 - \frac{T_c}{T_h}, where temperatures are measured on an , ensuring the ratio reflects intrinsic thermal properties rather than arbitrary calibrations. This formulation, derived from Sadi Carnot's 1824 analysis, guarantees that all reversible engines between the same reservoirs achieve identical efficiency, defining temperature ratios universally. Gas thermometers calibrate absolute scales using the , PV = nRT, where P is , V is , n is the , R is the , and T is the absolute temperature in kelvins. By maintaining constant volume and measuring pressure changes with temperature, or vice versa, the law extrapolates to where volume or pressure would theoretically vanish. This method provides a practical realization of the thermodynamic scale, approximating ideal behavior with dilute gases like at low pressures.

Theoretical Foundations

Kinetic Theory

The interprets temperature as a measure of the average of particles in random motion, providing a microscopic for macroscopic thermodynamic properties. This approach assumes that a gas consists of a large number of point-like particles with negligible compared to the , moving in lines between collisions that conserve both and , and that intermolecular forces are absent except during collisions. These assumptions idealize the gas as non-interacting except at instantaneous collisions, allowing of properties from particle . To derive the pressure exerted by the gas on the container , consider the momentum change from a particle colliding elastically with a to the x-direction. A particle of m with component v_x imparts an of $2 m v_x upon reversal, and with particles distributed uniformly, the number of collisions per unit time per unit area yields the P = \frac{1}{3} \rho \langle v^2 \rangle, where \rho is the and \langle v^2 \rangle is the speed. For N particles in volume V, this simplifies to P V = \frac{1}{3} N m \langle v^2 \rangle, establishing the kinetic . Linking the average \frac{1}{2} m \langle v^2 \rangle = \frac{3}{2} k T—where k is the and T the temperature—yields the P V = N k T, directly connecting temperature to microscopic motion. The speeds of particles follow the Maxwell-Boltzmann distribution, derived from the assumption of isotropic random motion and conservation laws in collisions. The for speed v is given by f(v) = \left( \frac{m}{2 \pi k T} \right)^{3/2} 4 \pi v^2 \exp\left( -\frac{m v^2}{2 k T} \right), which predicts that the most probable speed scales as \sqrt{T/m} and the root-mean-square speed as \sqrt{3 k T / m}. This distribution emerges from maximizing the number of microstates consistent with fixed under the classical assumptions, ensuring the average per particle is \frac{3}{2} k T. The underpins this energy-temperature relation, stating that in , each quadratic in the contributes \frac{1}{2} k T to the average energy. For a monatomic particle with three translational (kinetic energy terms \frac{1}{2} m v_x^2, \frac{1}{2} m v_y^2, \frac{1}{2} m v_z^2), the total average is thus \frac{3}{2} k T. This theorem arises from the equal weighting of volumes in , explaining why temperature quantifies the total translational energy in gases. While developed for gases, kinetic theory extends to liquids and solids by considering bound particles with vibrational motion. In solids, atoms oscillate around sites, contributing quadratic terms for both kinetic and in approximations, leading to an average energy of k T per mode via equipartition. For a three-dimensional , each atom has (three kinetic, three potential), yielding a total energy of $3 k T per atom and specific heats approaching $3 [R](/page/R) per at high temperatures, as observed in many metals. This vibrational contribution links temperature to dynamics, though quantum effects limit equipartition at low temperatures.

Thermodynamic Approach

In thermodynamics, temperature emerges as a fundamental parameter through the zeroth law, which establishes that if two systems are each in with a third system, they are in with each other; this allows temperature to be defined as the property shared by systems in mutual . The first law of relates temperature to energy changes in a via the , expressed as dU = \delta Q - \delta W, where dU is the change in , \delta Q is the added to the , and \delta W is the work done by the . For processes at constant , where \delta W = 0, this simplifies to dU = \delta Q, and the C_V is defined as C_V = \left( \frac{\delta Q}{dT} \right)_{V}, quantifying how temperature changes with added while holding fixed. The second law introduces entropy S as a state function that governs irreversible processes, with the differential form for reversible processes given by dS = \frac{\delta Q_{\text{rev}}}{T}, linking temperature directly to the rate of entropy change with reversible heat transfer. From this, temperature can be rigorously defined in terms of fundamental thermodynamic potentials as \frac{1}{T} = \left( \frac{\partial S}{\partial U} \right)_{V,N}, where the partial derivative is taken at constant volume V and particle number N, emphasizing temperature's role as the inverse of the entropy's sensitivity to internal energy. As an intensive property, temperature remains uniform throughout a in and does not depend on the system's size or the amount of matter present, distinguishing it from extensive properties like or . This uniformity ensures that, in equilibrium, all parts of an attain the same temperature regardless of scale. In the context of heat engines, temperature's thermodynamic role is exemplified by the , an idealized reversible cycle comprising two isothermal and two adiabatic processes, whose \eta = 1 - \frac{T_C}{T_H} depends solely on the absolute temperatures of the hot reservoir T_H and cold reservoir T_C, providing a scale-independent upper limit on the conversion of to work. This underscores temperature's as a universal measure of thermal potential in macroscopic systems.

Statistical Mechanics

In statistical mechanics, temperature emerges as a parameter characterizing the distribution of energy among microscopic states in a system at thermal equilibrium. This probabilistic framework bridges the macroscopic thermodynamic properties, such as those defined by the zeroth law, to the underlying microstates, providing a fundamental interpretation of temperature through ensemble theory. The microcanonical ensemble describes an isolated system with fixed energy E, volume V, and particle number N, where all accessible microstates are equally likely. The entropy S is given by S = k \ln \Omega, with k the Boltzmann constant and \Omega the number of microstates corresponding to energy E. Temperature is then defined as the inverse of the rate of change of entropy with energy, T = \left( \frac{\partial S}{\partial E} \right)_{V,N}^{-1} = \frac{1}{k} \left( \frac{\partial \ln \Omega}{\partial E} \right)_{V,N}^{-1}, linking macroscopic temperature directly to the density of states. This relation, originating from Boltzmann's foundational work, ensures consistency with the second law of thermodynamics by maximizing entropy for the given constraints. In the , the system exchanges with a at fixed temperature T, while V and N remain constant. The probability of a with E_i is proportional to e^{-[\beta](/page/Beta) E_i}, where [\beta](/page/Beta) = 1/(k[T](/page/KT)) is the inverse temperature. The partition function Z = \sum_i e^{-[\beta](/page/Beta) E_i} normalizes this distribution and encodes thermodynamic quantities, such as the F = -k[T](/page/KT) \ln Z. This ensemble, formalized by Gibbs, facilitates calculations for systems in contact with a bath, where temperature controls the Boltzmann factor's weighting of states. In non-equilibrium or non-ideal systems, the —measured via criteria like the zeroth law—may diverge from the statistical temperature, defined through local averages or kinetic definitions. For instance, in driven systems or those with spatial gradients, the statistical temperature can reflect microscopic fluctuations differently from the macroscopic value, leading to inconsistencies resolved only in the . Such differences highlight the limitations of equivalence outside ideal conditions. Negative temperatures arise in systems with bounded energy spectra, such as nuclear spin systems, where occurs—more particles occupy higher-energy states than lower ones. Here, the canonical distribution yields T < 0, as \beta < 0, corresponding to states hotter than temperature but unstable against energy exchange with positive-temperature reservoirs. This concept was experimentally realized in spins, where rapid magnetic field reversal induced inversion, confirming the thermodynamic consistency of negative T. Quantum extensions of statistical mechanics incorporate particle indistinguishability via Fermi-Dirac and Bose-Einstein statistics for fermions and bosons, respectively, particularly relevant for degenerate gases at low temperatures where quantum effects dominate. In a degenerate Fermi gas, the Fermi-Dirac distribution f(\epsilon) = [e^{(\epsilon - \mu)/kT} + 1]^{-1} fills states up to the Fermi energy \epsilon_F, with degeneracy setting in when T \ll T_F = \epsilon_F / k, leading to Pauli-blocked excitations and finite pressure even at T=0. For bosons, the Bose-Einstein distribution f(\epsilon) = [e^{(\epsilon - \mu)/kT} - 1]^{-1} allows condensation below T_c \approx (n \lambda^3)^{1/3} h / k (with n density and \lambda thermal wavelength), where a macroscopic ground-state occupation emerges in ideal gases. These statistics, derived by Fermi and Einstein, underpin phenomena like white dwarf stability and superfluidity.

Measurement Methods

Historical Devices

The earliest devices for detecting temperature changes were thermoscopes, which qualitatively indicated variations without quantitative scales. In the 3rd century BCE, Philo of Byzantium described an apparatus consisting of a hollow sphere connected to a tube submerged in water, where heating or cooling caused air expansion or contraction, displacing the water level to show temperature differences. This primitive design relied on the volumetric expansion of air and marked the first recorded attempt to observe thermal effects mechanically. Around 1593, Galileo Galilei improved upon such concepts by inventing an air thermoscope, a sealed bulb attached to a tube in a water reservoir, where rising or falling water levels in the tube visually demonstrated air's expansion with heat and contraction with cold, aiding early meteorological and experimental observations. Advancements in the shifted toward more precise liquid-based instruments suitable for medical applications. In , Italian physician adapted air thermoscopes for clinical use, employing them to monitor patients' body temperatures by observing fluid level changes, thus pioneering quantitative physiological measurements in . By sealing the devices to prevent interference, these early thermometers became more reliable. In 1714, German physicist introduced the first practical , using mercury's high and low freezing point for greater and accuracy compared to alcohol or air variants, enabling finer gradations in temperature readings. The 19th century brought electrical methods that transformed . In 1821, discovered the , observing that a junction of two dissimilar metals, such as and , generated a voltage proportional to the temperature difference (ΔT) between the hot and cold junctions, laying the foundation for thermocouples as robust sensors for high-temperature environments. This Seebeck effect allowed indirect electrical detection of temperature changes, with the voltage output serving as a measurable proxy for ΔT. Complementing this, in 1887, British physicist Hugh Longbourne Callendar developed the resistance thermometer, utilizing a platinum wire coil whose electrical resistance varied predictably with temperature according to the relation R = R_0 (1 + \alpha \Delta T), where R_0 is the resistance at a reference temperature, \alpha is the temperature coefficient of resistance, and \Delta T is the temperature change. This design offered high precision and stability, becoming a standard for calibration due to platinum's consistent properties. Calibration of these historical devices relied on reproducible fixed points, particularly the freezing and points of under , which provided natural benchmarks for scaling temperature intervals. Early thermometers, such as those by , used the point (freezing of at 32°F) and steam point ( of at 212°F) to define gradations, ensuring consistency across instruments despite variations in materials. These points, later refined in the by scientists like who inverted the scale to set freezing at 0°C and at 100°C, allowed for empirical without absolute theoretical foundations.

Modern Techniques

Modern temperature measurement techniques leverage advanced electronic, optical, and superconducting principles to achieve high precision across diverse environments, from cryogenic conditions to extreme high temperatures, often enabling non-contact and . These methods surpass traditional mechanical devices by offering resolutions down to millikelvin scales and response times in microseconds, essential for applications in , materials processing, and scientific . Recent advances include atom-based thermometers using Rydberg atoms, which provide ultra-high accuracy for fundamental , as demonstrated in developments reported in 2025. Infrared thermometry, particularly through pyrometers, measures temperature by detecting emitted from objects, assuming blackbody behavior for ideal cases. Pyrometers operate on the principle that all bodies above emit radiation, with intensity governed by , allowing non-contact measurements up to 3000°C without interference from the . For blackbody radiators, relates the peak wavelength of emission to temperature via \lambda_{\max} T = 2897.77 \, \mu\text{m} \cdot \text{K}, enabling temperature determination from . This technique is widely used in industrial furnaces and , though accuracy depends on correcting for variations in real materials. Optical methods provide versatile non-contact sensing for gases, fluids, and solids. excites fluorescent molecules with a laser, measuring the intensity or spectral shift of emitted light to infer temperature, as fluorescence yield decreases with rising thermal energy. Planar LIF enables two-dimensional mapping in combustion chambers and microfluidic devices, with resolutions below 1 K in flows up to 2000 K. detects temperature via shifts in scattered light wavelengths from molecular vibrations, using the anti-Stokes to Stokes intensity ratio for calibration. This approach achieves microscale resolution (~1.5 μm) in biological samples and harsh environments, with sensitivities up to 1.20 %/K at 300 K using titanium dioxide probes. Both techniques excel in transient, high-speed scenarios like engine testing, where physical probes would disrupt the medium. For cryogenic applications near , superconducting transition edge sensors (TES) offer exceptional sensitivity in the millikelvin range. TES devices consist of thin superconducting films, such as molybdenum-gold bilayers, biased at their critical temperature (~100 mK), where a small temperature rise induces a sharp resistance change due to the superconductor-normal metal transition. This enables energy resolutions of ~1.4 eV for detection at 100 mK, far superior to alternatives, and is pivotal in bolometers for and quantum computing cryostats. The sensitivity parameter \alpha = T/R \cdot dR/dT quantifies performance, with noise minimized at low base temperatures. High-temperature measurements in extreme environments, such as plasmas, employ robust optical and acoustic approaches. Optical fiber Bragg gratings (FBG) inscribed in silica or sapphire fibers detect temperature through shifts in reflected Bragg wavelength, caused by thermal expansion and refractive index changes, with sensitivities of 10–15 pm/K. Regenerated FBGs withstand up to 1173 K in radiation-heavy settings like nuclear reactors, while sapphire variants reach 2173 K with 1 K resolution, outperforming thermocouples in corrosive conditions. For plasmas, acoustic thermometry uses laser-induced breakdowns to generate sound waves, measuring their propagation speed—dependent on gas temperature—to infer values up to 1000 K with ±16 K accuracy. This method, validated against thermocouples, suits fusion and combustion diagnostics where optical access is limited. The International Temperature Scale of 1990 (ITS-90) standardizes these measurements using 17 defining fixed points from 0.65 K (³He vapor pressure) to 1357.77 K (Cu freezing point), ensuring global consistency through reproducible phase transitions of pure substances. Key points include the triple point of equilibrium hydrogen (13.8033 K), water (273.16 K), and freezing points of gallium (302.9146 K), indium (429.7485 K), tin (505.078 K), zinc (692.677 K), aluminum (933.473 K), silver (1234.93 K), and copper (1357.77 K), interpolated via resistance thermometers or radiation laws in subranges. This scale, adopted by the International Committee for Weights and Measures, underpins calibrations for all modern sensors, with uncertainties below 0.001 K at many points.
SubstanceTemperature (K)Type
³He0.65
e-H₂13.8033
Ne24.5561
O₂54.3584
Ar83.8058
234.3156
H₂O273.16
302.9146
In429.7485Freezing point
Sn505.078Freezing point
Zn692.677Freezing point
Al933.473Freezing point
Ag1234.93Freezing point
1337.33Freezing point
1357.77Freezing point

Physical Effects

Thermal Expansion

Thermal expansion refers to the tendency of materials to change their dimensions in response to temperature variations, a arising from the increased vibrational of atoms and molecules at higher temperatures. This effect is quantified by the coefficient of thermal expansion, which measures the fractional change in or per unit temperature change at constant . For linear expansion in solids, the coefficient α is defined as α = (1/L) (∂L/∂T)_P, where L is the and T is the temperature. In practical applications, for small temperature changes ΔT, the change in is approximated by ΔL = L₀ α ΔT, with L₀ as the initial . For volumetric expansion, the coefficient β is defined similarly as β = (1/V) (∂V/∂T)_P, where V is . In isotropic solids, where expansion occurs uniformly in all directions, β ≈ 3α. This relationship holds because volume changes result from the combined linear expansions along three perpendicular axes. In gases, follows , which states that at constant , V is directly proportional to the temperature T, or V ∝ T, implying β = 1/T for an . While most materials exhibit positive thermal expansion, some display anomalous behavior. Water, for instance, shows negative thermal expansion between 0°C and , where its reaches a maximum at due to structural changes in hydrogen bonding. Similarly, rubber exhibits anomalous expansion above its temperature, where the material shifts from a rigid glassy state to a flexible rubbery state, resulting in a significantly higher thermal expansion coefficient—typically about 3 times that of the glassy phase—owing to increased mobility. These properties find practical use in devices exploiting differential expansion. Bimetallic strips, composed of two metals with different α values bonded together, bend upon heating due to unequal expansion rates, enabling applications in thermostats to control circuits and in thermometers to indicate temperature via mechanical deflection.

Phase Changes

Phase changes in matter are fundamentally tied to temperature, where certain transitions between states occur at specific temperatures under equilibrium conditions, involving the absorption or release of without altering the system's temperature. These are known as phase transitions, characterized by a discontinuous change in volume and . During such transitions, the added or removed heat, termed , facilitates the reorganization of molecular structure while maintaining constant temperature. For instance, the melting of into at 0°C requires the absorption of 334 J/g of , allowing the to break into a state without temperature rise. The thermodynamics of these transitions is encapsulated in the Clausius-Clapeyron equation, which quantifies how pressure varies with temperature along the phase boundary: \frac{dP}{dT} = \frac{L}{T \Delta V} Here, L represents the molar of the transition, T is the equilibrium temperature, and \Delta V is the molar volume change between phases. This relation, derived from the equality of chemical potentials across phases, is particularly useful for vapor-liquid equilibria, where it predicts the increase in with temperature, explaining phenomena like under pressure. A special case arises at the , the unique combination of temperature and where , , and gas phases coexist in stable equilibrium, marking the intersection of the three boundaries. For , this point is precisely defined at 273.16 and 611.657 , serving as a fundamental reference for temperature scales due to its reproducibility. Beyond the critical point, phase distinctions blur, leading to s above the critical temperature, where no forms between liquid-like and gas-like states, and properties vary continuously with . , for example, reaches its critical temperature at 304.1 ; above this, it exists solely as a , exhibiting unique solvent properties useful in industrial applications like processes. Certain materials display hysteresis during phase changes, where the transition temperature depends on the direction of temperature change, resulting from kinetic barriers that delay the forward or reverse transformation. In shape-memory alloys, such as those based on NiTi, this hysteresis manifests in the martensitic phase transition, with the austenite-to-martensite conversion occurring at a lower temperature than the reverse upon heating, enabling the shape-memory effect but requiring careful control in applications like actuators.

Advanced Topics

Negative Temperature

Negative temperatures arise in isolated thermodynamic systems where the total is bounded from above, such as certain spin systems or laser media, allowing the to decrease as increases beyond a certain point. In these systems, the standard thermodynamic definition of temperature, given by T = \left( \frac{\partial E}{\partial S} \right)_{V,N}, where E is , S is , V is volume, and N is particle number, becomes negative when \frac{\partial S}{\partial E} < 0, corresponding to a state of where higher-energy states are more occupied than lower-energy ones. A classic example is a system of non-interacting particles in a , with two energy levels: a lower state at energy -\mu B and an upper state at +\mu B, where \mu is the and B is the field strength. At with equal populations in both states, the temperature is infinite, as is maximized; if the population in the upper state exceeds that in the lower state due to selective excitation, the temperature becomes negative, reflecting the inverted distribution. Similarly, in lasers, in the atomic energy levels of the gain medium—pumped by external energy—creates a state, enabling and coherent light output. Systems at negative temperatures are hotter than any system at positive temperature because heat naturally flows from negative-T to positive-T systems, as the former have higher per of freedom and tend to release to reach more stable positive-temperature states. This instability implies that negative-temperature states are transient and require from the environment to persist, such as by minimizing interactions with lattice vibrations in spin systems. The concept was first experimentally realized by Purcell and Pound in 1951, who observed negative temperatures in lithium fluoride nuclear spins subjected to radio-frequency fields, achieving population inversion without significant lattice coupling. Modern implementations routinely produce negative temperatures using (NMR) techniques, where radiofrequency pulses invert spin populations in isolated ensembles, allowing precise study of these states in controlled magnetic fields.

Relativistic Temperature

In , the concept of temperature becomes anisotropic when observed from a frame moving relative to the system's , due to the of the underlying particle distribution. For a system at rest with isotropic temperature T, an observer moving with v relative to it measures a direction-dependent temperature T' = T \gamma (1 - \beta \cos \theta), where \gamma = 1 / \sqrt{1 - \beta^2} is the , \beta = v/c, c is the , and \theta is the angle between the velocity vector and the direction of measurement. This transformation arises from the on the momenta of particles, making the higher in the forward direction and lower in the backward direction compared to the transverse case. Such anisotropy is crucial in contexts like high-velocity plasmas, where assuming a scalar temperature would lead to inconsistencies in thermodynamic relations. A significant extension occurs in accelerated frames through the , where an observer undergoing uniform a perceives the Minkowski vacuum as a thermal bath with temperature T = \frac{[\hbar](/page/H-bar) a}{2\pi k c}, with \hbar as the reduced Planck's constant and k as Boltzmann's constant. This effect, derived from the response of quantum detectors in , links acceleration to a fictitious temperature without any real heat source, highlighting the frame-dependence of thermal perception in quantum field theory. In general relativity, this manifests in black hole thermodynamics via the Hawking temperature at the event horizon, given by T_H = \frac{\hbar c^3}{8\pi G M k}, where G is the gravitational constant and M is the black hole mass. Smaller black holes thus exhibit higher temperatures, leading to evaporation through particle emission, with the formula establishing the scale where quantum effects dominate classical gravity. These relativistic notions apply in , where the (CMB) temperature scales as T \propto (1 + z) with z, reflecting the as a relativistic Doppler-like shift. This relation has been observationally confirmed up to z ≈ 6.3, providing a test of the standard model's adiabatic evolution. In particle accelerators, such as the (RHIC), collisions of heavy ions at near-light speeds produce quark-gluon plasmas at temperatures around 4 trillion , where relativistic transformations ensure consistent thermodynamic descriptions across frames. Ongoing debate in relativistic thermodynamics centers on whether temperature behaves as a scalar ( under boosts) or part of a tensor (direction-dependent or frame-variant). Proponents of the scalar view argue for invariance to preserve the second law across frames, while tensor approaches, incorporating four-vectors for and , resolve inconsistencies like spurious by yielding covariant transformations such as T' = T. This controversy underscores the need for a unified framework, with recent proposals favoring scalar invariance for thermodynamic consistency in accelerated systems.

References

  1. [1]
    Temperature - HyperPhysics Concepts
    We perceive temperature as the average translational kinetic energy associated with the disordered motion of atoms and molecules.
  2. [2]
    Kelvin: Thermodynamic Temperature | NIST
    May 15, 2018 · Thermodynamic temperature, by contrast, is an absolute measure of the average total internal energy of an object or objects—namely its kinetic ...
  3. [3]
    Thermodynamic Equilibrium
    Thermodynamic equilibrium leads to the large scale definition of temperature. When two objects are in thermal equilibrium they are said to have the same ...<|control11|><|separator|>
  4. [4]
    1.2 Thermometers and Temperature Scales - UCF Pressbooks
    Thermometers measure temperature according to well-defined scales of measurement. The three most common temperature scales are Fahrenheit, Celsius, and Kelvin.
  5. [5]
    13.1 Temperature – College Physics - University of Iowa Pressbooks
    Temperature Scales. Thermometers are used to measure temperature according to well-defined scales of measurement, which use pre-defined reference points to ...
  6. [6]
    Temperature Scales
    There are three temperature scales in use today, Fahrenheit, Celsius and Kelvin. Fahrenheit temperature scale is a scale based on 32 for the freezing point ...
  7. [7]
    SI Units – Temperature | NIST
    The kelvin temperature scale begins at absolute zero, the coldest possible temperature and the point at which even atoms would stand perfectly still.<|control11|><|separator|>
  8. [8]
    13.1 Temperature – College Physics chapters 1-17 - UH Pressbooks
    Section Summary · Temperature is the quantity measured by a thermometer. · Temperature is related to the average kinetic energy of atoms and molecules in a ...
  9. [9]
    [PDF] TEMPERATURE 6.1
    Temperature 6.1 describes methods for measuring water and air temperatures, essential for data collection, using Celsius scale. Liquid-in-glass and thermistor ...
  10. [10]
    How Do You Measure Air Temperature Accurately? | NIST
    Mar 26, 2025 · Air temperature is most often measured using electronic thermometers, which are accurate down to a fraction of a degree.
  11. [11]
    Heat and temperature - PMC - PubMed Central - NIH
    Jul 19, 2022 · Temperature and thermal energy dynamics are important concepts in the natural sciences, forming core properties in both physics and chemistry.
  12. [12]
    Zeroth Law - Thermal Equilibrium | Glenn Research Center - NASA
    May 2, 2024 · The zeroth law of thermodynamics is an observation. When two objects are separately in thermodynamic equilibrium with a third object, they are in equilibrium ...
  13. [13]
  14. [14]
  15. [15]
    Heat vs temperature - Energy Education
    Dec 20, 2021 · The core difference is that heat deals with thermal energy, whereas temperature is more concerned with molecular kinetic energy.
  16. [16]
    Kinetic Temperature, Thermal Energy - HyperPhysics
    Comparison with the ideal gas law leads to an expression for temperature sometimes referred to as the kinetic temperature.
  17. [17]
    1.1 Temperature and Thermal Equilibrium - UCF Pressbooks
    Temperature is operationally defined as the quantity measured by a thermometer. It is proportional to the average kinetic energy of atoms and molecules in a ...
  18. [18]
    Thermodynamic Equilibrium
    Thermodynamic equilibrium occurs when objects in contact stop changing properties, reaching the same temperature. If two objects are separately in equilibrium ...
  19. [19]
    [PDF] HEAT CONDUCTION EQUATION
    Heat conduction in a medium is said to be steady when the temperature does not vary with time, and unsteady or transient when it does.
  20. [20]
    18.3 Transient Heat Transfer (Convective Cooling or Heating) - MIT
    We examine a lumped parameter analysis of an object cooled by a stream. This will allow us to see what the relevant non-dimensional parameters are.
  21. [21]
    Local Thermodynamic Equilibrium - an overview - ScienceDirect.com
    Local thermodynamic equilibrium (LTE) is defined as a state in which the properties of a system, such as temperature and density, allow for the use of ...
  22. [22]
    [PDF] 2. Basic assumptions for stellar atmospheres
    Local Thermodynamic Equilibrium. A star as a whole (or a stellar atmosphere) is far from being in thermodynamic equilibrium: energy is transported from the ...
  23. [23]
    Thermodynamic Equilibrium, Local and otherwise
    It will turn out that one can define a "temperature" in some very strict sense only under special conditions, called thermodynamic equilibrium, which do not ...
  24. [24]
    [PDF] Temperature in non-equilibrium states: a review of open problems ...
    Oct 14, 2003 · The conceptual problems arising in the definition and measurement of temperature in non- equilibrium states are discussed in this paper in ...
  25. [25]
    [PDF] Temperature Measurement - MST.edu
    Since the device is a resistor, the only viable method of measuring the sensed temperature is to apply a small known current across the device and measure the ...
  26. [26]
    Bimetallic Pocket Thermometer | National Museum of American History
    The one marked "FAHR" refers to Daniel Gabriel Fahrenheit, a Polish scientist who in 1724 propsed a temperature scale according to which water froze at 32 ...
  27. [27]
    Chrono-Biographical Sketch: René-Antoine de Réaumur
    --1722: publishes his L'Art de Convertir le Fer Forgé en Acier... --1731: invents the Réaumur thermometer and devises a temperature scale with zero degrees ...
  28. [28]
    History of the Celsius temperature scale
    The French Réaumur scale had zero at the freezing point but ... Most probably Linné abandoned the Fahrenheit scale when he learned about Celsius' research.
  29. [29]
    [PDF] TEMPERATURE SCALES AND CONVERSIONS
    Aug 9, 2023 · We want in this brief note to define these scales and provide the necessary formulas relating them to each other. CELSIUS- Also known as ...
  30. [30]
    Charles' Law (Jacques-Alexandre-César Charles)
    Absolute zero is the temperature at which the volume of a gas becomes zero when the a plot of the volume versus temperature for a gas are extrapolated. As ...
  31. [31]
    Absolute Zero Temperature | Research Starters - EBSCO
    Absolute zero is defined as the lowest possible temperature, marking a theoretical limit at 0 kelvins, or -273.15 degrees Celsius. At this temperature ...
  32. [32]
    kelvin - BIPM
    The definition of the unit of thermodynamic temperature was given by the 10th CGPM which selected the triple point of water, TTPW, as a fundamental fixed point ...
  33. [33]
    Kelvin: Present Realization | NIST
    May 15, 2018 · Today, the kelvin is defined by taking the fixed numerical value of the Boltzmann constant k to be 1.380 649 ×10−23 when expressed in the unit ...
  34. [34]
    Rankine temperature scale | Description, Symbol, Conversion, & Facts
    The Rankine (°R) scale is the absolute temperature scale related to the Fahrenheit (°F) scale, and both scales have the same size unit of temperature.
  35. [35]
    Rankine - Energy Education
    Jul 21, 2018 · The Rankine (°R) is an absolute unit of temperature, named after physicist William John Macquorn Rankine. It is related to the Fahrenheit temperature scale.
  36. [36]
    6.2 The Thermodynamic Temperature Scale - MIT
    For any two temperatures , , the ratio of the magnitudes of the heat absorbed and rejected in a Carnot cycle has the same value for all systems. Figure 6.2: ...
  37. [37]
    Carnot efficiency - Energy Education
    Jul 22, 2020 · Carnot efficiency describes the maximum thermal efficiency that a heat engine can achieve as permitted by the Second Law of Thermodynamics.
  38. [38]
    2.9: Temperature and the Ideal Gas Thermometer
    Apr 27, 2023 · The current real-world standard temperature scale is the International Temperature Scale of 1990 (ITS-90). This defines temperature over a wide ...Missing: PV= nRT
  39. [39]
    9.2 Relating Pressure, Volume, Amount, and Temperature: The Ideal ...
    Feb 14, 2019 · Note that temperatures must be on the kelvin scale for any gas law calculations (0 on the kelvin scale and the lowest possible temperature is ...
  40. [40]
    [PDF] History of the Kinetic Theory of Gases* by Stephen G. Brush** Table ...
    Herapath submitted his first paper on kinetic theory to the Royal Society of London in. 1820, hoping to get it published in the Philosophical Transactions ...
  41. [41]
    [PDF] The scientific papers of James Clerk Maxwell
    before the University in 1878, complete this collection of Clerk Maxwell's scientific writings. The diagrams in this work have been re-produced by a ...
  42. [42]
    0th Law of Thermodynamics - Chemistry LibreTexts
    Jan 29, 2023 · The Zeroth Law of Thermodynamics states that if two systems are in thermodynamic equilibrium with a third system, the two original systems are in thermal ...
  43. [43]
    first law of thermodynamics: conservation of energy - MIT
    The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system.
  44. [44]
    5.13 The First Law of Thermodynamics – Douglas College Physics ...
    The first law of thermodynamics states that the change in internal energy of a system equals the net heat transfer into the system minus the net work done by ...
  45. [45]
    The First Law of Thermodynamics - Internal Energy, Work, and Heat
    Sep 10, 2025 · The first law of thermodynamics states that the energy of the universe is constant. The change in the internal energy of a system is the sum of the heat ...
  46. [46]
    19.2: Entropy and the Second Law of Thermodynamics
    Jul 7, 2023 · The second law of thermodynamics states that in a reversible process, the entropy of the universe is constant.Learning Objectives · Entropy · The Relationship between...
  47. [47]
    12.3 Second Law of Thermodynamics: Entropy - Physics | OpenStax
    Mar 26, 2020 · The second law of thermodynamics states that the total entropy of a system either increases or remains constant in any spontaneous process; it ...
  48. [48]
    The Thermodynamic Identity - HyperPhysics
    A useful summary relationship called the thermodynamic identity makes use of the power of calculus and particularly partial derivatives.
  49. [49]
    7.5: Partial Derivatives with Respect to \(T\), \(p\), and \(V\)
    Apr 12, 2022 · The tables in this section collect useful expressions for partial derivatives of the eight state functions T , p , V , U , H , A , G , and S in ...
  50. [50]
    1.3 Extensive and intensive properties – Introduction to Engineering ...
    All specific properties are intensive properties, as they refer to the corresponding extensive properties per unit mass, eg, specific volume and specific ...
  51. [51]
    1.4 Muddiest Points on Chapter 1 - MIT
    Intensive properties are properties that do not depend on the quantity of matter. For example, pressure and temperature are intensive properties. Energy, ...
  52. [52]
    2.4: Extensive and Intensive Properties - Chemistry LibreTexts
    Mar 20, 2025 · An intensive property is a property of matter that depends only on the type of matter in a sample and not on the amount. Color, temperature, and ...
  53. [53]
    4.6: The Carnot Cycle - Physics LibreTexts
    Mar 2, 2025 · A Carnot engine has an efficiency of 0.60 and the temperature of its cold reservoir is 300 K. (a) What is the temperature of the hot reservoir?Example 4 . 6 . 1 : The Carnot... · Example 4 . 6 . 2 : A Carnot...
  54. [54]
    3.3 The Carnot Cycle - MIT
    Second, the efficiency of a Carnot cycle is given compactly by. $\displaystyle \eta = 1 - \frac{T_1}{T_2}, (3..7). The efficiency can be 100% only if the ...
  55. [55]
    Carnot Cycle - Chemistry LibreTexts
    Feb 16, 2025 · The Carnot cycle is the most efficient engine possible based on the assumption of the absence of incidental wasteful processes such as friction.The Carnot Cycle · P-V Diagram · T-S Diagram · Efficiency<|control11|><|separator|>
  56. [56]
    [PDF] Lecture 7: Ensembles
    ensemble. In the microcanonical ensemble, the temperature is a derived quantity, with 1. T. = @S. @E . So far, we have only been using the microcanonical ...
  57. [57]
    Temperature in and out of equilibrium: a review of concepts, tools ...
    Nov 10, 2017 · We review the general aspects of the concept of temperature in equilibrium and non-equilibrium statistical mechanics.
  58. [58]
  59. [59]
    History of the Thermometer - PMC - NIH
    Aug 23, 2019 · The expansion and contraction of air with changes in temperature was noted as early as 220 BC by Philo of Byzantium.
  60. [60]
    The History of the Thermometer - ThoughtCo
    Jan 3, 2021 · In 1593, Galileo Galilei invented a rudimentary water thermoscope, which for the first time allowed temperature variations to be measured.
  61. [61]
    Kelvin: History | NIST - National Institute of Standards and Technology
    May 14, 2018 · The famous Italian astronomer and physicist Galileo, or possibly his friend the physician Santorio, likely came up with an improved thermoscope ...
  62. [62]
    Introduction | SpringerLink
    Mar 18, 2022 · Santorio instead had used the thermometer for medical practice and showed it publicly to his students and colleagues in Padua since 1611 ...
  63. [63]
    The Weight of the Air: Santorio's Thermometers and the Early History ...
    Medical Uses of the Thermometer. The last point leads us naturally to reflect on the application of Santorio's thermometer to medical diagnosis. Obizzi and ...
  64. [64]
    Daniel Gabriel Fahrenheit | Biography, Thermometer, & Facts
    Oct 28, 2025 · He is best known for inventing the mercury thermometer (1714) and developing the Fahrenheit temperature scale (1724), which is still commonly ...
  65. [65]
    Daniel Gabriel Fahrenheit and the Birth of Precision Thermometry
    In 1714, at the age of 28, he achieved his goal: developing a pair of thermometers that gave the same temperature reading.
  66. [66]
    Seebeck effect | Thermoelectricity, Temperature Gradients & Heat
    Oct 14, 2025 · The German physicist Thomas Johann Seebeck discovered (1821) the effect. The Seebeck effect is used to measure temperature with great ...
  67. [67]
    Thermocouples - HyperPhysics
    This temperature dependence of the junction potential was discovered by Seebeck in 1821, so is often called the "Seebeck effect".<|separator|>
  68. [68]
    Callendar's platinum thermometer | Opinion - Chemistry World
    Jul 7, 2022 · As Callendar pointed out in his 1887 paper, a new thermometric standard must be a highly reproducible portable device, stable enough not to ...
  69. [69]
    [PDF] Platinum resistance thermometry
    platinum resistance thermometers as useful precision instruments occurred in 1887 when H. L. Callendar. [16]' reported that platinum resistance thermom-.
  70. [70]
    The origins of fixed temperature points - Brannan
    Feb 17, 2021 · In 1742, Swedish scientist Anders Celsius chose 0 degrees for the boiling point of water, and 100 degrees for the freezing point. A year later, ...
  71. [71]
    [PDF] Understanding Radiation Thermometry – Part I
    Jul 8, 2015 · The wavelength at the maximum radiant intensity decreases as the temperature increases (Wien's displacement law). ... blackbody (Kirchhoff's Law).
  72. [72]
    None
    ### Summary of Principles of Noncontact IR Temperature Measurement Using Pyrometers
  73. [73]
    A state-of-the-art review on laser-induced fluorescence (LIF) method ...
    Laser-induced fluorescence (LIF) is a robust and vigorous method for non-intrusive measurement of temperature, pressure, concentration, and pH in fluids.
  74. [74]
    Contactless Temperature Sensing at the Microscale Based on ...
    Apr 2, 2021 · It follows that temperature can be measured from Raman spectra by determining the degree of the shift position of a defined peak at different ...
  75. [75]
    Transition-Edge Sensors - Figueroa Group - Northwestern University
    A transition-edge sensor is a thermometer made from a superconducting film operated near its transition temperature Tc. While in its transition from ...
  76. [76]
    [PDF] Superconducting transition edge sensor using dilute AlMn alloys
    In this letter, we report the results from a superconduct- ing transition-edge sensor (TES) bolometer fabricated using manganese impurities to reduce the Tc of ...
  77. [77]
    Recent advancements in fiber Bragg gratings based temperature ...
    Distributed FBGs based temperature sensor can measure the temperature in the range of 473 K to 773 K under a challenging operating environment consisting of ...
  78. [78]
    Laser-induced breakdown thermometry via time-of-arrival ...
    Sep 20, 2018 · In this study, laser-induced breakdown thermometry based on time-of-arrival measurements of associated acoustic waves is proposed.
  79. [79]
    [PDF] Guidelines for realizing the International Temperature Scale of 1990 ...
    Recommendation of the CIPM that the numerical value of 273.16 °K, exactly, be assigned to the triple point of water on the thermodynamic scale, ...
  80. [80]
    [PDF] Guide to the Realization of the ITS-90 - BIPM
    Oct 20, 2021 · The defining fixed points of the ITS-90 concerned in this section are given in Table 1. Above the triple point of water, the assigned values of ...
  81. [81]
    [PDF] Thermal Expansion - Rice University
    10–6/K. 1.8. The coefficient of thermal expansion is also often defined as the fractional increase in length per unit rise in temperature.
  82. [82]
    1.3 Thermal Expansion – General Physics Using Calculus I
    Use the equation for linear thermal expansion Δ L = α L Δ T to calculate the change in length, Δ L . Use the coefficient of linear expansion α for steel from ...
  83. [83]
    coefficient of linear expansion. - Physics
    Basic features: The length change is proportional to the temperature change. Here α is the coefficient of linear expansion. where β = 3 α is the coefficient of ...
  84. [84]
    Charles's Law - Chemistry 301
    Charles's Law states that the Volume (V) of a gas is directly proportional to the temperature (T). This law is valid as long as the pressure and the amount of ...
  85. [85]
    [PDF] Modeling the Effect of Plasticizer on the Viscoelastic Response of ...
    Mar 31, 2015 · Above Tg,t , the material has a rubbery thermal expansion coefficient, and below Tg,t the material is in a glassy state with a glassy thermal ...
  86. [86]
    VT Physics Lecture Demo H14 - Virginia Tech Physics
    H14: Thermal Expansion of Metals. Purpose: To show the thermal expansion of metals, and how this effect is used in bimetallic thermometers and thermostats.
  87. [87]
    To drink or to pour: How should athletes use water to cool themselves?
    ... heat required to melt ice, known as the latent heat of fusion, is much greater than the specific heat capacity of water at 334 J/g. It is this much greater ...
  88. [88]
    23.4: The Clausius-Clapeyron Equation - Chemistry LibreTexts
    Mar 4, 2025 · The Clapeyron equation, which tells us that the slope of the coexistence curve is related to the ratio of the molar enthalpy between the phases to the change ...Evaporation · The Clapeyron Equation · Example 23.4.1 · Learning Objectives
  89. [89]
    [PDF] Highaccuracy measurements of the vapor pressure of ice referenced ...
    Dec 6, 2013 · the water triple-point vapor pressure of Guildner et al. [1976] (pi. (Tt = 273.16 K) = 611.657 Pa, which has a specified relative uncertainty of.
  90. [90]
    Carbon dioxide - the NIST WebBook
    Temperature (K), Method, Reference, Comment. 16.7, 288. A, Stephenson and Malanowski, 1987, Based on data from 273. to 304. K.; AC. 16.4, 258. A, Stephenson and ...
  91. [91]
    (PDF) Hysteresis in shape-memory alloys - ResearchGate
    Aug 8, 2025 · We present an overview of hysteresis phenomena in the martensitic transformation, and their relevance in the thermomechanical behaviour of shape-memory alloys.
  92. [92]
    Thermodynamics and Statistical Mechanics at Negative Absolute ...
    Negative temperatures are hotter than positive temperatures. When account is taken of the possibility of negative temperatures, various modifications of ...
  93. [93]
    Population Inversion – gain, upper laser level - RP Photonics
    Formally, population inversion is sometimes described as a state with a negative temperature. This is because a Boltzmann distribution would lead to such a ...
  94. [94]
    Physics of negative absolute temperatures | Phys. Rev. E
    Jan 17, 2017 · Thus the Boltzmann entropy formula (9) predicts that in the new equilibrium state, the energy will be shared between the two groups of spins.
  95. [95]
    A Nuclear Spin System at Negative Temperature | Phys. Rev.
    A nuclear spin system at negative temperature. EM Purcell and RV Pound Department of Physics, Harvard University, Cambridge, Massachusetts.
  96. [96]
    [PDF] Temperature and Special Relativity 1 T = T /γ - Kirk T. McDonald
    Dec 2, 2020 · That temperature is a Lorentz invariant was argued for a few years, starting in 1966, by. Landsberg [35, 38, 43, 44, 49]. 4. This view attracted ...
  97. [97]
    What is the temperature of a moving body? | Scientific Reports - Nature
    Dec 15, 2017 · Starting from Planck and Einstein, several authors have proposed their own reasoning, concluding that a moving body could appear cooler, hotter ...
  98. [98]
    [PDF] On relativistic temperature | PhilSci-Archive
    Jul 1, 2023 · Abstract. I revisit the long-running controversy as to the transformation proper- ties of temperature under Lorentz transformations, and argue ...
  99. [99]
    Unruh effect - Scholarpedia
    Oct 10, 2014 · Because the acceleration necessary to reach a temperature of 1\, {\rm K} through the Unruh effect is of order 10^{20} \, {\rm m/s^2}, one might ...
  100. [100]
  101. [101]
    [1405.1277] Tests of the CMB temperature-redshift relation ... - arXiv
    May 6, 2014 · In the expanding Universe, the average temperature of the cosmic microwave background (CMB) is expected to depend like TCMB~(1+z) on redshift z.Missing: relativistic | Show results with:relativistic
  102. [102]
  103. [103]
    Brookhaven's RHIC Sets Record | BNL Newsroom
    Guinness World Records has recognized our 2.4-mile particle accelerator, the Relativistic Heavy Ion Collider (RHIC), for generating a record temperature of 7.2 ...
  104. [104]
    Resolving inconsistencies in Relativistic Thermodynamics with four-entropy and four-heat current
    ### Summary of Debate/Inconsistencies in Relativistic Thermodynamics Regarding Temperature Transformation