fbpx
Wikipedia

Temperature

Temperature is a physical quantity that expresses quantitatively the perceptions of hotness and coldness. Temperature is measured with a thermometer.

Temperature
Thermal vibration of a segment of protein alpha helix. Its amplitude increases with temperature
Common symbols
T
SI unitK
Other units
°C, °F, °R, °Rø, °Ré, °N, °D, °L, °W
Intensive?Yes
Derivations from
other quantities
,
Dimension

Thermometers are calibrated in various temperature scales that historically have relied on various reference points and thermometric substances for definition. The most common scales are the Celsius scale with the unit symbol °C (formerly called centigrade), the Fahrenheit scale (°F), and the Kelvin scale (K), the latter being used predominantly for scientific purposes. The kelvin is one of the seven base units in the International System of Units (SI).

Absolute zero, i.e., zero kelvin or −273.15 °C, is the lowest point in the thermodynamic temperature scale. Experimentally, it can be approached very closely but not actually reached, as recognized in the third law of thermodynamics. It would be impossible to extract energy as heat from a body at that temperature.

Temperature is important in all fields of natural science, including physics, chemistry, Earth science, astronomy, medicine, biology, ecology, material science, metallurgy, mechanical engineering and geography as well as most aspects of daily life.

Effects

 
Average daily variation in human body temperature

Many physical processes are related to temperature; some of them are given below:

Scales

 
Two thermometers showing temperature in Celsius and Fahrenheit

Temperature scales need two values for definition: the point chosen as zero degrees and the magnitudes of the incremental unit of temperature.

The Celsius scale (°C) is used for common temperature measurements in most of the world. It is an empirical scale that developed historically, which led to its zero point 0 °C being defined as the freezing point of water, and 100 °C as the boiling point of water, both at atmospheric pressure at sea level. It was called a centigrade scale because of the 100-degree interval.[3] Since the standardization of the kelvin in the International System of Units, it has subsequently been redefined in terms of the equivalent fixing points on the Kelvin scale, so that a temperature increment of one degree Celsius is the same as an increment of one kelvin, though numerically the scales differ by an exact offset of 273.15.

The Fahrenheit scale is in common use in the United States. Water freezes at 32 °F and boils at 212 °F at sea-level atmospheric pressure.

Absolute zero

At the absolute zero of temperature, no energy can be removed from matter as heat, a fact expressed in the third law of thermodynamics. At this temperature, matter contains no macroscopic thermal energy, but still has quantum-mechanical zero-point energy as predicted by the uncertainty principle, although this does not enter into the definition of absolute temperature. Experimentally, absolute zero can be approached only very closely; it can never be reached (the lowest temperature attained by experiment is 100 pK).[citation needed] Theoretically, in a body at a temperature of absolute zero, all classical motion of its particles has ceased and they are at complete rest in this classical sense. The absolute zero, defined as 0 K, is exactly equal to −273.15 °C, or −459.67 °F.

Absolute scales

Referring to the Boltzmann constant, to the Maxwell–Boltzmann distribution, and to the Boltzmann statistical mechanical definition of entropy, as distinct from the Gibbs definition,[4] for independently moving microscopic particles, disregarding interparticle potential energy, by international agreement, a temperature scale is defined and said to be absolute because it is independent of the characteristics of particular thermometric substances and thermometer mechanisms. Apart from the absolute zero, it does not have a reference temperature. It is known as the Kelvin scale, widely used in science and technology. The kelvin (the unit name is spelled with a lower-case 'k') is the unit of temperature in the International System of Units (SI). The temperature of a body in a state of thermodynamic equilibrium is always positive relative to the absolute zero.

Besides the internationally agreed Kelvin scale, there is also a thermodynamic temperature scale, invented by Lord Kelvin, also with its numerical zero at the absolute zero of temperature, but directly relating to purely macroscopic thermodynamic concepts, including the macroscopic entropy, though microscopically referable to the Gibbs statistical mechanical definition of entropy for the canonical ensemble, that takes interparticle potential energy into account, as well as independent particle motion so that it can account for measurements of temperatures near absolute zero.[4] This scale has a reference temperature at the triple point of water, the numerical value of which is defined by measurements using the aforementioned internationally agreed Kelvin scale.

Kelvin scale

Many scientific measurements use the Kelvin temperature scale (unit symbol: K), named in honor of the physicist who first defined it. It is an absolute scale. Its numerical zero point, 0 K, is at the absolute zero of temperature. Since May, 2019, the kelvin has been defined through particle kinetic theory, and statistical mechanics. In the International System of Units (SI), the magnitude of the kelvin is defined in terms of the Boltzmann constant, the value of which is defined as fixed by international convention.[5][6]

Statistical mechanical versus thermodynamic temperature scales

Since May 2019, the magnitude of the kelvin is defined in relation to microscopic phenomena, characterized in terms of statistical mechanics. Previously, but since 1954, the International System of Units defined a scale and unit for the kelvin as a thermodynamic temperature, by using the reliably reproducible temperature of the triple point of water as a second reference point, the first reference point being 0 K at absolute zero.[citation needed]

Historically, the temperature of the triple point of water was defined as exactly 273.16 K. Today it is an empirically measured quantity. The freezing point of water at sea-level atmospheric pressure occurs at very close to 273.15 K (0 °C).

Classification of scales

There are various kinds of temperature scale. It may be convenient to classify them as empirically and theoretically based. Empirical temperature scales are historically older, while theoretically based scales arose in the middle of the nineteenth century.[7][8]

Empirical scales

Empirically based temperature scales rely directly on measurements of simple macroscopic physical properties of materials. For example, the length of a column of mercury, confined in a glass-walled capillary tube, is dependent largely on temperature and is the basis of the very useful mercury-in-glass thermometer. Such scales are valid only within convenient ranges of temperature. For example, above the boiling point of mercury, a mercury-in-glass thermometer is impracticable. Most materials expand with temperature increase, but some materials, such as water, contract with temperature increase over some specific range, and then they are hardly useful as thermometric materials. A material is of no use as a thermometer near one of its phase-change temperatures, for example, its boiling-point.

In spite of these limitations, most generally used practical thermometers are of the empirically based kind. Especially, it was used for calorimetry, which contributed greatly to the discovery of thermodynamics. Nevertheless, empirical thermometry has serious drawbacks when judged as a basis for theoretical physics. Empirically based thermometers, beyond their base as simple direct measurements of ordinary physical properties of thermometric materials, can be re-calibrated, by use of theoretical physical reasoning, and this can extend their range of adequacy.

Theoretical scales

Theoretically based temperature scales are based directly on theoretical arguments, especially those of kinetic theory and thermodynamics. They are more or less ideally realized in practically feasible physical devices and materials. Theoretically based temperature scales are used to provide calibrating standards for practical empirically based thermometers.

Microscopic statistical mechanical scale

In physics, the internationally agreed conventional temperature scale is called the Kelvin scale. It is calibrated through the internationally agreed and prescribed value of the Boltzmann constant,[5][6] referring to motions of microscopic particles, such as atoms, molecules, and electrons, constituent in the body whose temperature is to be measured. In contrast with the thermodynamic temperature scale invented by Kelvin, the presently conventional Kelvin temperature is not defined through comparison with the temperature of a reference state of a standard body, nor in terms of macroscopic thermodynamics.

Apart from the absolute zero of temperature, the Kelvin temperature of a body in a state of internal thermodynamic equilibrium is defined by measurements of suitably chosen of its physical properties, such as have precisely known theoretical explanations in terms of the Boltzmann constant.[citation needed] That constant refers to chosen kinds of motion of microscopic particles in the constitution of the body. In those kinds of motion, the particles move individually, without mutual interaction. Such motions are typically interrupted by inter-particle collisions, but for temperature measurement, the motions are chosen so that, between collisions, the non-interactive segments of their trajectories are known to be accessible to accurate measurement. For this purpose, interparticle potential energy is disregarded.

In an ideal gas, and in other theoretically understood bodies, the Kelvin temperature is defined to be proportional to the average kinetic energy of non-interactively moving microscopic particles, which can be measured by suitable techniques. The proportionality constant is a simple multiple of the Boltzmann constant. If molecules, atoms, or electrons,[9][10] are emitted from material and their velocities are measured, the spectrum of their velocities often nearly obeys a theoretical law called the Maxwell–Boltzmann distribution, which gives a well-founded measurement of temperatures for which the law holds.[11] There have not yet been successful experiments of this same kind that directly use the Fermi–Dirac distribution for thermometry, but perhaps that will be achieved in the future.[12]

The speed of sound in a gas can be calculated theoretically from the molecular character of the gas, from its temperature and pressure, and from the value of the Boltzmann constant. For a gas of known molecular character and pressure, this provides a relation between temperature and the Boltzmann constant. Those quantities can be known or measured more precisely than can the thermodynamic variables that define the state of a sample of water at its triple point. Consequently, taking the value of the Boltzmann constant as a primarily defined reference of exactly defined value, a measurement of the speed of sound can provide a more precise measurement of the temperature of the gas.[13]

Measurement of the spectrum of electromagnetic radiation from an ideal three-dimensional black body can provide an accurate temperature measurement because the frequency of maximum spectral radiance of black-body radiation is directly proportional to the temperature of the black body; this is known as Wien's displacement law and has a theoretical explanation in Planck's law and the Bose–Einstein law.

Measurement of the spectrum of noise-power produced by an electrical resistor can also provide accurate temperature measurement. The resistor has two terminals and is in effect a one-dimensional body. The Bose-Einstein law for this case indicates that the noise-power is directly proportional to the temperature of the resistor and to the value of its resistance and to the noise bandwidth. In a given frequency band, the noise-power has equal contributions from every frequency and is called Johnson noise. If the value of the resistance is known then the temperature can be found.[14][15]

Macroscopic thermodynamic scale

Historically, till May 2019, the definition of the Kelvin scale was that invented by Kelvin, based on a ratio of quantities of energy in processes in an ideal Carnot engine, entirely in terms of macroscopic thermodynamics.[citation needed] That Carnot engine was to work between two temperatures, that of the body whose temperature was to be measured, and a reference, that of a body at the temperature of the triple point of water. Then the reference temperature, that of the triple point, was defined to be exactly 273.16 K. Since May 2019, that value has not been fixed by definition but is to be measured through microscopic phenomena, involving the Boltzmann constant, as described above. The microscopic statistical mechanical definition does not have a reference temperature.

Ideal gas

A material on which a macroscopically defined temperature scale may be based is the ideal gas. The pressure exerted by a fixed volume and mass of an ideal gas is directly proportional to its temperature. Some natural gases show so nearly ideal properties over suitable temperature range that they can be used for thermometry; this was important during the development of thermodynamics and is still of practical importance today.[16][17] The ideal gas thermometer is, however, not theoretically perfect for thermodynamics. This is because the entropy of an ideal gas at its absolute zero of temperature is not a positive semi-definite quantity, which puts the gas in violation of the third law of thermodynamics. In contrast to real materials, the ideal gas does not liquefy or solidify, no matter how cold it is. Alternatively thinking, the ideal gas law, refers to the limit of infinitely high temperature and zero pressure; these conditions guarantee non-interactive motions of the constituent molecules.[18][19][20]

Kinetic theory approach

The magnitude of the kelvin is now defined in terms of kinetic theory, derived from the value of the Boltzmann constant.

Kinetic theory provides a microscopic account of temperature for some bodies of material, especially gases, based on macroscopic systems' being composed of many microscopic particles, such as molecules and ions of various species, the particles of a species being all alike. It explains macroscopic phenomena through the classical mechanics of the microscopic particles. The equipartition theorem of kinetic theory asserts that each classical degree of freedom of a freely moving particle has an average kinetic energy of kBT/2 where kB denotes the Boltzmann constant.[citation needed] The translational motion of the particle has three degrees of freedom, so that, except at very low temperatures where quantum effects predominate, the average translational kinetic energy of a freely moving particle in a system with temperature T will be 3kBT/2.

Molecules, such as oxygen (O2), have more degrees of freedom than single spherical atoms: they undergo rotational and vibrational motions as well as translations. Heating results in an increase of temperature due to an increase in the average translational kinetic energy of the molecules. Heating will also cause, through equipartitioning, the energy associated with vibrational and rotational modes to increase. Thus a diatomic gas will require more energy input to increase its temperature by a certain amount, i.e. it will have a greater heat capacity than a monatomic gas.

As noted above, the speed of sound in a gas can be calculated from the molecular character of the gas, from its temperature and pressure, and from the value of the Boltzmann constant. Taking the value of the Boltzmann constant as a primarily defined reference of exactly defined value, a measurement of the speed of sound can provide a more precise measurement of the temperature of the gas.[13]

It is possible to measure the average kinetic energy of constituent microscopic particles if they are allowed to escape from the bulk of the system, through a small hole in the containing wall. The spectrum of velocities has to be measured, and the average calculated from that. It is not necessarily the case that the particles that escape and are measured have the same velocity distribution as the particles that remain in the bulk of the system, but sometimes a good sample is possible.

Thermodynamic approach

Temperature is one of the principal quantities in the study of thermodynamics. Formerly, the magnitude of the kelvin was defined in thermodynamic terms, but nowadays, as mentioned above, it is defined in terms of kinetic theory.

The thermodynamic temperature is said to be absolute for two reasons. One is that its formal character is independent of the properties of particular materials. The other reason is that its zero is, in a sense, absolute, in that it indicates absence of microscopic classical motion of the constituent particles of matter, so that they have a limiting specific heat of zero for zero temperature, according to the third law of thermodynamics. Nevertheless, a thermodynamic temperature does in fact have a definite numerical value that has been arbitrarily chosen by tradition and is dependent on the property of particular materials; it is simply less arbitrary than relative "degrees" scales such as Celsius and Fahrenheit. Being an absolute scale with one fixed point (zero), there is only one degree of freedom left to arbitrary choice, rather than two as in relative scales. For the Kelvin scale since May 2019, by international convention, the choice has been made to use knowledge of modes of operation of various thermometric devices, relying on microscopic kinetic theories about molecular motion. The numerical scale is settled by a conventional definition of the value of the Boltzmann constant, which relates macroscopic temperature to average microscopic kinetic energy of particles such as molecules. Its numerical value is arbitrary, and an alternate, less widely used absolute temperature scale exists called the Rankine scale, made to be aligned with the Fahrenheit scale as Kelvin is with Celsius.

The thermodynamic definition of temperature is due to Kelvin. It is framed in terms of an idealized device called a Carnot engine, imagined to run in a fictive continuous cycle of successive processes that traverse a cycle of states of its working body. The engine takes in a quantity of heat Q1 from a hot reservoir and passes out a lesser quantity of waste heat Q2 < 0 to a cold reservoir. The net heat energy absorbed by the working body is passed, as thermodynamic work, to a work reservoir, and is considered to be the output of the engine. The cycle is imagined to run so slowly that at each point of the cycle the working body is in a state of thermodynamic equilibrium. The successive processes of the cycle are thus imagined to run reversibly with no entropy production. Then the quantity of entropy taken in from the hot reservoir when the working body is heated is equal to that passed to the cold reservoir when the working body is cooled. Then the absolute or thermodynamic temperatures, T1 and T2, of the reservoirs are defined such that[21]

 

 

 

 

 

(1)

The zeroth law of thermodynamics allows this definition to be used to measure the absolute or thermodynamic temperature of an arbitrary body of interest, by making the other heat reservoir have the same temperature as the body of interest.

Kelvin's original work postulating absolute temperature was published in 1848. It was based on the work of Carnot, before the formulation of the first law of thermodynamics. Carnot had no sound understanding of heat and no specific concept of entropy. He wrote of 'caloric' and said that all the caloric that passed from the hot reservoir was passed into the cold reservoir. Kelvin wrote in his 1848 paper that his scale was absolute in the sense that it was defined "independently of the properties of any particular kind of matter". His definitive publication, which sets out the definition just stated, was printed in 1853, a paper read in 1851.[22][23][24][25]

Numerical details were formerly settled by making one of the heat reservoirs a cell at the triple point of water, which was defined to have an absolute temperature of 273.16 K.[26] Nowadays, the numerical value is instead obtained from measurement through the microscopic statistical mechanical international definition, as above.

Intensive variability

In thermodynamic terms, temperature is an intensive variable because it is equal to a differential coefficient of one extensive variable with respect to another, for a given body. It thus has the dimensions of a ratio of two extensive variables. In thermodynamics, two bodies are often considered as connected by contact with a common wall, which has some specific permeability properties. Such specific permeability can be referred to a specific intensive variable. An example is a diathermic wall that is permeable only to heat; the intensive variable for this case is temperature. When the two bodies have been connected through the specifically permeable wall for a very long time, and have settled to a permanent steady state, the relevant intensive variables are equal in the two bodies; for a diathermal wall, this statement is sometimes called the zeroth law of thermodynamics.[27][28][29]

In particular, when the body is described by stating its internal energy U, an extensive variable, as a function of its entropy S, also an extensive variable, and other state variables V, N, with U = U (S, V, N), then the temperature is equal to the partial derivative of the internal energy with respect to the entropy:[28][29][30]

 

 

 

 

 

(2)

Likewise, when the body is described by stating its entropy S as a function of its internal energy U, and other state variables V, N, with S = S (U, V, N), then the reciprocal of the temperature is equal to the partial derivative of the entropy with respect to the internal energy:[28][30][31]

 

 

 

 

 

(3)

The above definition, equation (1), of the absolute temperature, is due to Kelvin. It refers to systems closed to the transfer of matter and has a special emphasis on directly experimental procedures. A presentation of thermodynamics by Gibbs starts at a more abstract level and deals with systems open to the transfer of matter; in this development of thermodynamics, the equations (2) and (3) above are actually alternative definitions of temperature.[32]

Local thermodynamic equilibrium

Real-world bodies are often not in thermodynamic equilibrium and not homogeneous. For the study by methods of classical irreversible thermodynamics, a body is usually spatially and temporally divided conceptually into 'cells' of small size. If classical thermodynamic equilibrium conditions for matter are fulfilled to good approximation in such a 'cell', then it is homogeneous and a temperature exists for it. If this is so for every 'cell' of the body, then local thermodynamic equilibrium is said to prevail throughout the body.[33][34][35][36][37]

It makes good sense, for example, to say of the extensive variable U, or of the extensive variable S, that it has a density per unit volume or a quantity per unit mass of the system, but it makes no sense to speak of the density of temperature per unit volume or quantity of temperature per unit mass of the system. On the other hand, it makes no sense to speak of the internal energy at a point, while when local thermodynamic equilibrium prevails, it makes good sense to speak of the temperature at a point. Consequently, the temperature can vary from point to point in a medium that is not in global thermodynamic equilibrium, but in which there is local thermodynamic equilibrium.

Thus, when local thermodynamic equilibrium prevails in a body, the temperature can be regarded as a spatially varying local property in that body, and this is because the temperature is an intensive variable.

Basic theory

Temperature is a measure of a quality of a state of a material.[38] The quality may be regarded as a more abstract entity than any particular temperature scale that measures it, and is called hotness by some writers.[39][40][41] The quality of hotness refers to the state of material only in a particular locality, and in general, apart from bodies held in a steady state of thermodynamic equilibrium, hotness varies from place to place. It is not necessarily the case that a material in a particular place is in a state that is steady and nearly homogeneous enough to allow it to have a well-defined hotness or temperature. Hotness may be represented abstractly as a one-dimensional manifold. Every valid temperature scale has its own one-to-one map into the hotness manifold.[42][43]

When two systems in thermal contact are at the same temperature no heat transfers between them. When a temperature difference does exist heat flows spontaneously from the warmer system to the colder system until they are in thermal equilibrium. Such heat transfer occurs by conduction or by thermal radiation.[44][45][46][47][48][49][50][51]

Experimental physicists, for example Galileo and Newton,[52] found that there are indefinitely many empirical temperature scales. Nevertheless, the zeroth law of thermodynamics says that they all measure the same quality. This means that for a body in its own state of internal thermodynamic equilibrium, every correctly calibrated thermometer, of whatever kind, that measures the temperature of the body, records one and the same temperature. For a body that is not in its own state of internal thermodynamic equilibrium, different thermometers can record different temperatures, depending respectively on the mechanisms of operation of the thermometers.

Bodies in thermodynamic equilibrium

For experimental physics, hotness means that, when comparing any two given bodies in their respective separate thermodynamic equilibria, any two suitably given empirical thermometers with numerical scale readings will agree as to which is the hotter of the two given bodies, or that they have the same temperature.[53] This does not require the two thermometers to have a linear relation between their numerical scale readings, but it does require that the relation between their numerical readings shall be strictly monotonic.[54][55] A definite sense of greater hotness can be had, independently of calorimetry, of thermodynamics, and of properties of particular materials, from Wien's displacement law of thermal radiation: the temperature of a bath of thermal radiation is proportional, by a universal constant, to the frequency of the maximum of its frequency spectrum; this frequency is always positive, but can have values that tend to zero. Thermal radiation is initially defined for a cavity in thermodynamic equilibrium. These physical facts justify a mathematical statement that hotness exists on an ordered one-dimensional manifold. This is a fundamental character of temperature and thermometers for bodies in their own thermodynamic equilibrium.[7][42][43][56][57]

Except for a system undergoing a first-order phase change such as the melting of ice, as a closed system receives heat, without a change in its volume and without a change in external force fields acting on it, its temperature rises. For a system undergoing such a phase change so slowly that departure from thermodynamic equilibrium can be neglected, its temperature remains constant as the system is supplied with latent heat. Conversely, a loss of heat from a closed system, without phase change, without change of volume, and without a change in external force fields acting on it, decreases its temperature.[58]

Bodies in a steady state but not in thermodynamic equilibrium

While for bodies in their own thermodynamic equilibrium states, the notion of temperature requires that all empirical thermometers must agree as to which of two bodies is the hotter or that they are at the same temperature, this requirement is not safe for bodies that are in steady states though not in thermodynamic equilibrium. It can then well be that different empirical thermometers disagree about which is hotter, and if this is so, then at least one of the bodies does not have a well-defined absolute thermodynamic temperature. Nevertheless, anyone has given body and any one suitable empirical thermometer can still support notions of empirical, non-absolute, hotness, and temperature, for a suitable range of processes. This is a matter for study in non-equilibrium thermodynamics.[citation needed]

Bodies not in a steady state

When a body is not in a steady-state, then the notion of temperature becomes even less safe than for a body in a steady state not in thermodynamic equilibrium. This is also a matter for study in non-equilibrium thermodynamics.

Thermodynamic equilibrium axiomatics

For the axiomatic treatment of thermodynamic equilibrium, since the 1930s, it has become customary to refer to a zeroth law of thermodynamics. The customarily stated minimalist version of such a law postulates only that all bodies, which when thermally connected would be in thermal equilibrium, should be said to have the same temperature by definition, but by itself does not establish temperature as a quantity expressed as a real number on a scale. A more physically informative version of such a law views empirical temperature as a chart on a hotness manifold.[42][57][59] While the zeroth law permits the definitions of many different empirical scales of temperature, the second law of thermodynamics selects the definition of a single preferred, absolute temperature, unique up to an arbitrary scale factor, whence called the thermodynamic temperature.[7][42][60][61][62][63] If internal energy is considered as a function of the volume and entropy of a homogeneous system in thermodynamic equilibrium, thermodynamic absolute temperature appears as the partial derivative of internal energy with respect the entropy at constant volume. Its natural, intrinsic origin or null point is absolute zero at which the entropy of any system is at a minimum. Although this is the lowest absolute temperature described by the model, the third law of thermodynamics postulates that absolute zero cannot be attained by any physical system.

Heat capacity

When an energy transfer to or from a body is only as heat, the state of the body changes. Depending on the surroundings and the walls separating them from the body, various changes are possible in the body. They include chemical reactions, increase of pressure, increase of temperature and phase change. For each kind of change under specified conditions, the heat capacity is the ratio of the quantity of heat transferred to the magnitude of the change.[64]

For example, if the change is an increase in temperature at constant volume, with no phase change and no chemical change, then the temperature of the body rises and its pressure increases. The quantity of heat transferred, ΔQ, divided by the observed temperature change, ΔT, is the body's heat capacity at constant volume:

 

If heat capacity is measured for a well-defined amount of substance, the specific heat is the measure of the heat required to increase the temperature of such a unit quantity by one unit of temperature. For example, raising the temperature of water by one kelvin (equal to one degree Celsius) requires 4186 joules per kilogram (J/kg).

Measurement

 
A typical Celsius thermometer measures a winter day temperature of −17 °C

Temperature measurement using modern scientific thermometers and temperature scales goes back at least as far as the early 18th century, when Daniel Gabriel Fahrenheit adapted a thermometer (switching to mercury) and a scale both developed by Ole Christensen Rømer. Fahrenheit's scale is still in use in the United States for non-scientific applications.

Temperature is measured with thermometers that may be calibrated to a variety of temperature scales. In most of the world (except for Belize, Myanmar, Liberia and the United States), the Celsius scale is used for most temperature measuring purposes. Most scientists measure temperature using the Celsius scale and thermodynamic temperature using the Kelvin scale, which is the Celsius scale offset so that its null point is 0 K = −273.15 °C, or absolute zero. Many engineering fields in the US, notably high-tech and US federal specifications (civil and military), also use the Kelvin and Celsius scales. Other engineering fields in the US also rely upon the Rankine scale (a shifted Fahrenheit scale) when working in thermodynamic-related disciplines such as combustion.

Units

The basic unit of temperature in the International System of Units (SI) is the kelvin. It has the symbol K.

For everyday applications, it is often convenient to use the Celsius scale, in which 0 °C corresponds very closely to the freezing point of water and 100 °C is its boiling point at sea level. Because liquid droplets commonly exist in clouds at sub-zero temperatures, 0 °C is better defined as the melting point of ice. In this scale, a temperature difference of 1 degree Celsius is the same as a 1kelvin increment, but the scale is offset by the temperature at which ice melts (273.15 K).

By international agreement,[65] until May 2019, the Kelvin and Celsius scales were defined by two fixing points: absolute zero and the triple point of Vienna Standard Mean Ocean Water, which is water specially prepared with a specified blend of hydrogen and oxygen isotopes. Absolute zero was defined as precisely 0 K and −273.15 °C. It is the temperature at which all classical translational motion of the particles comprising matter ceases and they are at complete rest in the classical model. Quantum-mechanically, however, zero-point motion remains and has an associated energy, the zero-point energy. Matter is in its ground state,[66] and contains no thermal energy. The temperatures 273.16 K and 0.01 °C were defined as those of the triple point of water. This definition served the following purposes: it fixed the magnitude of the kelvin as being precisely 1 part in 273.16 parts of the difference between absolute zero and the triple point of water; it established that one kelvin has precisely the same magnitude as one degree on the Celsius scale; and it established the difference between the null points of these scales as being 273.15 K (0 K = −273.15 °C and 273.16 K = 0.01 °C). Since 2019, there has been a new definition based on the Boltzmann constant,[67] but the scales are scarcely changed.

In the United States, the Fahrenheit scale is the most widely used. On this scale the freezing point of water corresponds to 32 °F and the boiling point to 212 °F. The Rankine scale, still used in fields of chemical engineering in the US, is an absolute scale based on the Fahrenheit increment.

Historical scales

The following temperature scales are in use or have historically been used for measuring temperature:

Plasma physics

The field of plasma physics deals with phenomena of electromagnetic nature that involve very high temperatures. It is customary to express temperature as energy in a unit related to the electronvolt or kiloelectronvolt (eV/kB or keV/kB). The corresponding energy, which is dimensionally distinct from temperature, is then calculated as the product of the Boltzmann constant and temperature,  . Then, 1 eV/kB is 11605 K. In the study of QCD matter one routinely encounters temperatures of the order of a few hundred MeV/kB, equivalent to about 1012 K.

Theoretical foundation

Historically, there are several scientific approaches to the explanation of temperature: the classical thermodynamic description based on macroscopic empirical variables that can be measured in a laboratory; the kinetic theory of gases which relates the macroscopic description to the probability distribution of the energy of motion of gas particles; and a microscopic explanation based on statistical physics and quantum mechanics. In addition, rigorous and purely mathematical treatments have provided an axiomatic approach to classical thermodynamics and temperature.[68] Statistical physics provides a deeper understanding by describing the atomic behavior of matter and derives macroscopic properties from statistical averages of microscopic states, including both classical and quantum states. In the fundamental physical description, the temperature may be measured directly in units of energy. However, in the practical systems of measurement for science, technology, and commerce, such as the modern metric system of units, the macroscopic and the microscopic descriptions are interrelated by the Boltzmann constant, a proportionality factor that scales temperature to the microscopic mean kinetic energy.

The microscopic description in statistical mechanics is based on a model that analyzes a system into its fundamental particles of matter or into a set of classical or quantum-mechanical oscillators and considers the system as a statistical ensemble of microstates. As a collection of classical material particles, the temperature is a measure of the mean energy of motion, called translational kinetic energy, of the particles, whether in solids, liquids, gases, or plasmas. The kinetic energy, a concept of classical mechanics, is half the mass of a particle times its speed squared. In this mechanical interpretation of thermal motion, the kinetic energies of material particles may reside in the velocity of the particles of their translational or vibrational motion or in the inertia of their rotational modes. In monatomic perfect gases and, approximately, in most gas and in simple metals, the temperature is a measure of the mean particle translational kinetic energy, 3/2 kBT. It also determines the probability distribution function of energy. In condensed matter, and particularly in solids, this purely mechanical description is often less useful and the oscillator model provides a better description to account for quantum mechanical phenomena. Temperature determines the statistical occupation of the microstates of the ensemble. The microscopic definition of temperature is only meaningful in the thermodynamic limit, meaning for large ensembles of states or particles, to fulfill the requirements of the statistical model.

Kinetic energy is also considered as a component of thermal energy. The thermal energy may be partitioned into independent components attributed to the degrees of freedom of the particles or to the modes of oscillators in a thermodynamic system. In general, the number of these degrees of freedom that are available for the equipartitioning of energy depends on the temperature, i.e. the energy region of the interactions under consideration. For solids, the thermal energy is associated primarily with the vibrations of its atoms or molecules about their equilibrium position. In an ideal monatomic gas, the kinetic energy is found exclusively in the purely translational motions of the particles. In other systems, vibrational and rotational motions also contribute degrees of freedom.

Kinetic theory of gases

 
A theoretical understanding of temperature in a hard-sphere model of a gas can be obtained from the Kinetic theory.

Maxwell and Boltzmann developed a kinetic theory that yields a fundamental understanding of temperature in gases.[69] This theory also explains the ideal gas law and the observed heat capacity of monatomic (or 'noble') gases.[70][71][72]

 
Plots of pressure vs temperature for three different gas samples extrapolated to absolute zero

The ideal gas law is based on observed empirical relationships between pressure (p), volume (V), and temperature (T), and was recognized long before the kinetic theory of gases was developed (see Boyle's and Charles's laws). The ideal gas law states:[73]

 

where n is the number of moles of gas and R = 8.314462618... J⋅mol−1⋅K−1[74] is the gas constant.

This relationship gives us our first hint that there is an absolute zero on the temperature scale, because it only holds if the temperature is measured on an absolute scale such as Kelvin's. The ideal gas law allows one to measure temperature on this absolute scale using the gas thermometer. The temperature in kelvins can be defined as the pressure in pascals of one mole of gas in a container of one cubic meter, divided by the gas constant.

Although it is not a particularly convenient device, the gas thermometer provides an essential theoretical basis by which all thermometers can be calibrated. As a practical matter, it is not possible to use a gas thermometer to measure absolute zero temperature since the gases condense into a liquid long before the temperature reaches zero. It is possible, however, to extrapolate to absolute zero by using the ideal gas law, as shown in the figure.

The kinetic theory assumes that pressure is caused by the force associated with individual atoms striking the walls, and that all energy is translational kinetic energy. Using a sophisticated symmetry argument,[75] Boltzmann deduced what is now called the Maxwell–Boltzmann probability distribution function for the velocity of particles in an ideal gas. From that probability distribution function, the average kinetic energy (per particle) of a monatomic ideal gas is[71][76]

 

where the Boltzmann constant kB is the ideal gas constant divided by the Avogadro number, and   is the root-mean-square speed.[77] This direct proportionality between temperature and mean molecular kinetic energy is a special case of the equipartition theorem, and holds only in the classical limit of a perfect gas. It does not hold exactly for most substances.

Zeroth law of thermodynamics

When two otherwise isolated bodies are connected together by a rigid physical path impermeable to matter, there is the spontaneous transfer of energy as heat from the hotter to the colder of them. Eventually, they reach a state of mutual thermal equilibrium, in which heat transfer has ceased, and the bodies' respective state variables have settled to become unchanging.[78][79][80]

One statement of the zeroth law of thermodynamics is that if two systems are each in thermal equilibrium with a third system, then they are also in thermal equilibrium with each other.[81][82][83]

This statement helps to define temperature but it does not, by itself, complete the definition. An empirical temperature is a numerical scale for the hotness of a thermodynamic system. Such hotness may be defined as existing on a one-dimensional manifold, stretching between hot and cold. Sometimes the zeroth law is stated to include the existence of a unique universal hotness manifold, and of numerical scales on it, so as to provide a complete definition of empirical temperature.[59] To be suitable for empirical thermometry, a material must have a monotonic relation between hotness and some easily measured state variable, such as pressure or volume, when all other relevant coordinates are fixed. An exceptionally suitable system is the ideal gas, which can provide a temperature scale that matches the absolute Kelvin scale. The Kelvin scale is defined on the basis of the second law of thermodynamics.

Second law of thermodynamics

As an alternative to considering or defining the zeroth law of thermodynamics, it was the historical development in thermodynamics to define temperature in terms of the second law of thermodynamics which deals with entropy.[citation needed] The second law states that any process will result in either no change or a net increase in the entropy of the universe. This can be understood in terms of probability.

For example, in a series of coin tosses, a perfectly ordered system would be one in which either every toss comes up heads or every toss comes up tails. This means the outcome is always 100% the same result. In contrast, many mixed (disordered) outcomes are possible, and their number increases with each toss. Eventually, the combinations of ~50% heads and ~50% tails dominate, and obtaining an outcome significantly different from 50/50 becomes increasingly unlikely. Thus the system naturally progresses to a state of maximum disorder or entropy.

As temperature governs the transfer of heat between two systems and the universe tends to progress toward a maximum of entropy, it is expected that there is some relationship between temperature and entropy. A heat engine is a device for converting thermal energy into mechanical energy, resulting in the performance of work. An analysis of the Carnot heat engine provides the necessary relationships. According to energy conservation and energy being a state function that does not change over a full cycle, the work from a heat engine over a full cycle is equal to the net heat, i.e. the sum of the heat put into the system at high temperature, qH > 0, and the waste heat given off at the low temperature, qC < 0.[84]

The efficiency is the work divided by the heat input:

 

 

 

 

 

(4)

where wcy is the work done per cycle. The efficiency depends only on |qC|/qH. Because qC and qH correspond to heat transfer at the temperatures TC and TH, respectively, |qC|/qH should be some function of these temperatures:

 

 

 

 

 

(5)

Carnot's theorem states that all reversible engines operating between the same heat reservoirs are equally efficient.[citation needed] Thus, a heat engine operating between T1 and T3 must have the same efficiency as one consisting of two cycles, one between T1 and T2, and the second between T2 and T3. This can only be the case if

 

which implies

 

Since the first function is independent of T2, this temperature must cancel on the right side, meaning f(T1, T3) is of the form g(T1)/g(T3) (i.e. f(T1, T3) = f(T1, T2)f(T2, T3) = g(T1)/g(T2) · g(T2)/g(T3) = g(T1)/g(T3)), where g is a function of a single temperature. A temperature scale can now be chosen with the property that

 

 

 

 

 

(6)

Substituting (6) back into (4) gives a relationship for the efficiency in terms of temperature:

 

 

 

 

 

(7)

For TC = 0 K the efficiency is 100% and that efficiency becomes greater than 100% below 0 K. Since an efficiency greater than 100% violates the first law of thermodynamics, this implies that 0 K is the minimum possible temperature. In fact, the lowest temperature ever obtained in a macroscopic system was 20 nK, which was achieved in 1995 at NIST. Subtracting the right hand side of (5) from the middle portion and rearranging gives[21][84]

 

where the negative sign indicates heat ejected from the system. This relationship suggests the existence of a state function, S, whose change characteristically vanishes for a complete cycle if it is defined by

 

 

 

 

 

(8)

where the subscript indicates a reversible process. This function corresponds to the entropy of the system, which was described previously. Rearranging (8) gives a formula for temperature in terms of fictive infinitesimal quasi-reversible elements of entropy and heat:

 

 

 

 

 

(9)

For a constant-volume system where entropy S(E) is a function of its energy E, dE = dqrev and (9) gives

 

 

 

 

 

(10)

i.e. the reciprocal of the temperature is the rate of increase of entropy with respect to energy at constant volume.

Definition from statistical mechanics

Statistical mechanics defines temperature based on a system's fundamental degrees of freedom. Eq.(10) is the defining relation of temperature, where the entropy   is defined (up to a constant) by the logarithm of the number of microstates of the system in the given macrostate (as specified in the microcanonical ensemble):

 

where   is the Boltzmann constant and W is the number of microstates with the energy E of the system (degeneracy).

When two systems with different temperatures are put into purely thermal connection, heat will flow from the higher temperature system to the lower temperature one; thermodynamically this is understood by the second law of thermodynamics: The total change in entropy following a transfer of energy   from system 1 to system 2 is:

 

and is thus positive if  

From the point of view of statistical mechanics, the total number of microstates in the combined system 1 + system 2 is  , the logarithm of which (times the Boltzmann constant) is the sum of their entropies; thus a flow of heat from high to low temperature, which brings an increase in total entropy, is more likely than any other scenario (normally it is much more likely), as there are more microstates in the resulting macrostate.

Generalized temperature from single-particle statistics

It is possible to extend the definition of temperature even to systems of few particles, like in a quantum dot. The generalized temperature is obtained by considering time ensembles instead of configuration-space ensembles given in statistical mechanics in the case of thermal and particle exchange between a small system of fermions (N even less than 10) with a single/double-occupancy system. The finite quantum grand canonical ensemble,[85] obtained under the hypothesis of ergodicity and orthodicity,[86] allows expressing the generalized temperature from the ratio of the average time of occupation   and   of the single/double-occupancy system:[87]

 

where EF is the Fermi energy. This generalized temperature tends to the ordinary temperature when N goes to infinity.

Negative temperature

On the empirical temperature scales that are not referenced to absolute zero, a negative temperature is one below the zero-point of the scale used. For example, dry ice has a sublimation temperature of −78.5 °C which is equivalent to −109.3 °F.[88] On the absolute Kelvin scale this temperature is 194.6 K. No body can be brought to exactly 0 K (the temperature of the ideally coldest possible body) by any finite practicable process; this is a consequence of the third law of thermodynamics.[89][90][91]

The international kinetic theory temperature of a body cannot take negative values. The thermodynamic temperature scale, however, is not so constrained.

For a body of matter, there can sometimes be conceptually defined, in terms of microscopic degrees of freedom, namely particle spins, a subsystem, with a temperature other than that of the whole body. When the body is in its own state of internal thermodynamic equilibrium, the temperatures of the whole body and of the subsystem must be the same. The two temperatures can differ when, by work through externally imposed force fields, energy can be transferred to and from the subsystem, separately from the rest of the body; then the whole body is not in its own state of internal thermodynamic equilibrium. There is an upper limit of energy such a spin subsystem can attain.

Considering the subsystem to be in a temporary state of virtual thermodynamic equilibrium, it is possible to obtain a negative temperature on the thermodynamic scale. Thermodynamic temperature is the inverse of the derivative of the subsystem's entropy with respect to its internal energy. As the subsystem's internal energy increases, the entropy increases for some range, but eventually attains a maximum value and then begins to decrease as the highest energy states begin to fill. At the point of maximum entropy, the temperature function shows the behavior of a singularity, because the slope of the entropy as a function of energy decreases to zero and then turns negative. As the subsystem's entropy reaches its maximum, its thermodynamic temperature goes to positive infinity, switching to negative infinity as the slope turns negative. Such negative temperatures are hotter than any positive temperature. Over time, when the subsystem is exposed to the rest of the body, which has a positive temperature, energy is transferred as heat from the negative temperature subsystem to the positive temperature system.[92] The kinetic theory temperature is not defined for such subsystems.

Examples

Comparisons of temperatures in various scales
Temperature Peak emittance wavelength[93]
of black-body radiation
Kelvin Celsius
Absolute zero
(precisely by definition)
0 K −273.15 °C Infinity
Blackbody temperature of the black hole at
the centre of our galaxy, Sagittarius A*[94]
15 fK −273.149999999999985 °C 2.5×108 km (1.7 AU)
Lowest temperature
achieved[95]
100 pK −273.149999999900 °C 29000 km
Coldest
Bose–Einstein condensate[96]
450 pK −273.14999999955 °C 6400 km
One millikelvin
(precisely by definition)
0.001 K −273.149 °C 2.89777 m
(radio, FM band)[97]
Cosmic microwave background
(2013 measurement)
2.7260 K −270.424 °C 0.00106301 m
(millimeter-wavelength microwave)
Water triple point
(precisely by definition)
273.16 K 0.01 °C 10608.3 nm
(long-wavelength IR)
Water boiling point[A] 373.1339 K 99.9839 °C 7766.03 nm
(mid-wavelength IR)
Iron melting point 1811 K 1538 °C 1600 nm
(far infrared)
Incandescent lamp[B] 2500 K 2200 °C 1160 nm
(near infrared)[C]
Sun's visible surface[D][98] 5778 K 5505 °C 501.5 nm
(green-blue light)
Lightning bolt
channel[E]
28 kK 28000 °C 100 nm
(far ultraviolet light)
Sun's core[E] 16 MK 16 million °C 0.18 nm (X-rays)
Thermonuclear weapon
(peak temperature)[E][99]
350 MK 350 million °C 8.3×10−3 nm
(gamma rays)
Sandia National Labs'
Z machine[E][100]
2 GK 2 billion °C 1.4×10−3 nm
(gamma rays)[F]
Core of a high-mass
star on its last day
[E][101]
3 GK 3 billion °C 1×10−3 nm
(gamma rays)
Merging binary neutron
star
system[E][102]
350 GK 350 billion °C 8×10−6 nm
(gamma rays)
Relativistic Heavy
Ion Collider
[E][103]
1 TK 1 trillion °C 3×10−6 nm
(gamma rays)
CERN's proton vs
nucleus collisions[E][104]
10 TK 10 trillion °C 3×10−7 nm
(gamma rays)
  • A For Vienna Standard Mean Ocean Water at one standard atmosphere (101.325 kPa) when calibrated strictly per the two-point definition of thermodynamic temperature.
  • B The 2500 K value is approximate. The 273.15 K difference between K and °C is rounded to 300 K to avoid false precision in the Celsius value.
  • C For a true black-body (which tungsten filaments are not). Tungsten filament emissivity is greater at shorter wavelengths, which makes them appear whiter.
  • D Effective photosphere temperature. The 273.15 K difference between K and °C is rounded to 273 K to avoid false precision in the Celsius value.
  • E The 273.15 K difference between K and °C is within the precision of these values.
  • F For a true black-body (which the plasma was not). The Z machine's dominant emission originated from 40 MK electrons (soft x-ray emissions) within the plasma.

See also

Notes and references

  1. ^ Agency, International Atomic Energy (1974). Thermal discharges at nuclear power stations: their management and environmental impacts: a report prepared by a group of experts as the result of a panel meeting held in Vienna, 23–27 October 1972. International Atomic Energy Agency.
  2. ^ Watkinson, John (2001). The Art of Digital Audio. Taylor & Francis. ISBN 978-0-240-51587-8.
  3. ^ Middleton, W.E.K. (1966), pp. 89–105.
  4. ^ a b Jaynes, E.T. (1965), pp. 391–398.
  5. ^ a b Cryogenic Society (2019).
  6. ^ a b (PDF), archived from the original (PDF) on 2018-04-29, retrieved 2019-10-20
  7. ^ a b c Truesdell, C.A. (1980), Sections 11 B, 11H, pp. 306–310, 320–332.
  8. ^ Quinn, T. J. (1983).
  9. ^ Germer, L.H. (1925). 'The distribution of initial velocities among thermionic electrons', Phys. Rev., 25: 795–807. here
  10. ^ Turvey, K. (1990). 'Test of validity of Maxwellian statistics for electrons thermionically emitted from an oxide cathode', European Journal of Physics, 11(1): 51–59. here
  11. ^ Zeppenfeld, M., Englert, B.G.U., Glöckner, R., Prehn, A., Mielenz, M., Sommer, C., van Buuren, L.D., Motsch, M., Rempe, G. (2012).
  12. ^ Miller, J. (2013).
  13. ^ a b de Podesta, M., Underwood, R., Sutton, G., Morantz, P, Harris, P, Mark, D.F., Stuart, F.M., Vargha, G., Machin, M. (2013). A low-uncertainty measurement of the Boltzmann constant, Metrologia, 50 (4): S213–S216, BIPM & IOP Publishing Ltd
  14. ^ Quinn, T.J. (1983), pp. 98–107.
  15. ^ Schooley, J.F. (1986), pp. 138–143.
  16. ^ Quinn, T.J. (1983), pp. 61–83.
  17. ^ Schooley, J.F. (1986), pp. 115–138.
  18. ^ Adkins, C.J. (1968/1983), pp. 119–120.
  19. ^ Buchdahl, H.A. (1966), pp. 137–138.
  20. ^ Tschoegl, N.W. (2000), p. 88.
  21. ^ a b Fermi, E. (1956). Thermodynamics. Dover Publications (still in print). p. 48. eq.(64).
  22. ^ Thomson, W. (Lord Kelvin) (1848).
  23. ^ Thomson, W. (Lord Kelvin) (1851).
  24. ^ Partington, J.R. (1949), pp. 175–177.
  25. ^ Roberts, J.K., Miller, A.R. (1928/1960), pp. 321–322.
  26. ^ Quinn, T.J. (1983). Temperature, Academic Press, London, ISBN 0-12-569680-9, pp. 160–162.
  27. ^ Tisza, L. (1966). Generalized Thermodynamics, M.I.T. Press, Cambridge MA, pp. 47, 57.
  28. ^ a b c Münster, A. (1970), Classical Thermodynamics, translated by E.S. Halberstadt, Wiley–Interscience, London, ISBN 0-471-62430-6, pp. 49, 69.
  29. ^ a b Bailyn, M. (1994). A Survey of Thermodynamics, American Institute of Physics Press, New York, ISBN 0-88318-797-3, pp. 14–15, 214.
  30. ^ a b Callen, H.B. (1960/1985), Thermodynamics and an Introduction to Thermostatistics, (first edition 1960), second edition 1985, John Wiley & Sons, New York, ISBN 0-471-86256-8, pp. 146–148.
  31. ^ Kondepudi, D., Prigogine, I. (1998). Modern Thermodynamics. From Heat Engines to Dissipative Structures, John Wiley, Chichester, ISBN 0-471-97394-7, pp. 115–116.
  32. ^ Tisza, L. (1966). Generalized Thermodynamics, M.I.T. Press, Cambridge MA, p. 58.
  33. ^ Milne, E.A. (1929). The effect of collisions on monochromatic radiative equilibrium, Monthly Notices of the Royal Astronomical Society, 88: 493–502.
  34. ^ Gyarmati, I. (1970). Non-equilibrium Thermodynamics. Field Theory and Variational Principles, translated by E. Gyarmati and W.F. Heinz, Springer, Berlin, pp. 63–66.
  35. ^ Glansdorff, P., Prigogine, I., (1971). Thermodynamic Theory of Structure, Stability and Fluctuations, Wiley, London, ISBN 0-471-30280-5, pp. 14–16.
  36. ^ Bailyn, M. (1994). A Survey of Thermodynamics, American Institute of Physics Press, New York, ISBN 0-88318-797-3, pp. 133–135.
  37. ^ Callen, H.B. (1960/1985), Thermodynamics and an Introduction to Thermostatistics, (first edition 1960), second edition 1985, John Wiley & Sons, New York, ISBN 0-471-86256-8, pp. 309–310.
  38. ^ Bryan, G.H. (1907). Thermodynamics. An Introductory Treatise dealing mainly with First Principles and their Direct Applications, B.G. Teubner, Leipzig, p. 3. "Thermodynamics by George Hartley Bryan". from the original on 2011-11-18. Retrieved 2011-10-02.
  39. ^ Pippard, A.B. (1957/1966), p. 18.
  40. ^ Adkins,C.J. (1968/1983), p. 20.
  41. ^ Bryan, G.H. (1907). Thermodynamics. An Introductory Treatise dealing mainly with First Principles and their Direct Applications, B.G. Teubner, Leipzig, p. 5: "... when a body is spoken of as growing hotter or colder an increase of temperature is always implied, for the hotness and coldness of a body are qualitative terms which can only refer to temperature." "Thermodynamics by George Hartley Bryan". from the original on 2011-11-18. Retrieved 2011-10-02.
  42. ^ a b c d Mach, E. (1900). Die Principien der Wärmelehre. Historisch-kritisch entwickelt, Johann Ambrosius Barth, Leipzig, section 22, pp. 56–57.
  43. ^ a b Serrin, J. (1986). Chapter 1, 'An Outline of Thermodynamical Structure', pp. 3–32, especially p. 6, in New Perspectives in Thermodynamics, edited by J. Serrin, Springer, Berlin, ISBN 3-540-15931-2.
  44. ^ Maxwell, J.C. (1872). Theory of Heat, third edition, Longmans, Green, London, p. 32.
  45. ^ Tait, P.G. (1884). Heat, Macmillan, London, Chapter VII, pp. 39–40.
  46. ^ Planck, M. (1897/1903). Treatise on Thermodynamics, translated by A. Ogg, Longmans, Green, London, pp. 1–2.
  47. ^ Planck, M. (1914), The Theory of Heat Radiation 2011-11-18 at the Wayback Machine, second edition, translated into English by M. Masius, Blakiston's Son & Co., Philadelphia, reprinted by Kessinger.
  48. ^ J.S. Dugdale (1996). Entropy and its Physical Interpretation. Taylor & Francis. p. 13. ISBN 978-0-7484-0569-5.
  49. ^ F. Reif (1965). Fundamentals of Statistical and Thermal Physics. McGraw-Hill. p. 102. ISBN 9780070518001.
  50. ^ M.J. Moran; H.N. Shapiro (2006). "1.6.1". Fundamentals of Engineering Thermodynamics (5 ed.). John Wiley & Sons, Ltd. p. 14. ISBN 978-0-470-03037-0.{{cite book}}: CS1 maint: multiple names: authors list (link)
  51. ^ T.W. Leland, Jr. "Basic Principles of Classical and Statistical Thermodynamics" (PDF). p. 14. (PDF) from the original on 2011-09-28. Consequently we identify temperature as a driving force which causes something called heat to be transferred.
  52. ^ Tait, P.G. (1884). Heat, Macmillan, London, Chapter VII, pp. 42, 103–117.
  53. ^ Beattie, J.A., Oppenheim, I. (1979). Principles of Thermodynamics, Elsevier Scientific Publishing Company, Amsterdam, ISBN 978-0-444-41806-7, p. 29.
  54. ^ Landsberg, P.T. (1961). Thermodynamics with Quantum Statistical Illustrations, Interscience Publishers, New York, p. 17.
  55. ^ Thomsen, J.S. (1962). "A restatement of the zeroth law of thermodynamics". Am. J. Phys. 30 (4): 294–296. Bibcode:1962AmJPh..30..294T. doi:10.1119/1.1941991.
  56. ^ Maxwell, J.C. (1872). Theory of Heat, third edition, Longman's, Green & Co, London, p. 45.
  57. ^ a b Pitteri, M. (1984). On the axiomatic foundations of temperature, Appendix G6 on pp. 522–544 of Rational Thermodynamics, C. Truesdell, second edition, Springer, New York, ISBN 0-387-90874-9.
  58. ^ Truesdell, C., Bharatha, S. (1977). The Concepts and Logic of Classical Thermodynamics as a Theory of Heat Engines, Rigorously Constructed upon the Foundation Laid by S. Carnot and F. Reech, Springer, New York, ISBN 0-387-07971-8, p. 20.
  59. ^ a b Serrin, J. (1978). The concepts of thermodynamics, in Contemporary Developments in Continuum Mechanics and Partial Differential Equations. Proceedings of the International Symposium on Continuum Mechanics and Partial Differential Equations, Rio de Janeiro, August 1977, edited by G.M. de La Penha, L.A.J. Medeiros, North-Holland, Amsterdam, ISBN 0-444-85166-6, pp. 411–451.
  60. ^ Maxwell, J.C. (1872). Theory of Heat, third edition, Longmans, Green, London, pp. 155–158.
  61. ^ Tait, P.G. (1884). Heat, Macmillan, London, Chapter VII, Section 95, pp. 68–69.
  62. ^ Buchdahl, H.A. (1966), p. 73.
  63. ^ Kondepudi, D. (2008). Introduction to Modern Thermodynamics, Wiley, Chichester, ISBN 978-0-470-01598-8, Section 32., pp. 106–108.
  64. ^ Green, Don; Perry, Robert H. (2008). Perry's Chemical Engineers' Handbook, Eighth Edition (8th ed.). McGraw-Hill Education. p. 660. ISBN 978-0071422949.
  65. ^ The kelvin in the SI Brochure 2007-09-26 at the Wayback Machine
  66. ^ "Absolute Zero". Calphad.com. from the original on 2011-07-08. Retrieved 2010-09-16.
  67. ^ Definition agreed by the 26th General Conference on Weights and Measures (CGPM) 2020-10-09 at the Wayback Machine in November 2018, implemented 20 May 2019
  68. ^ C. Caratheodory (1909). "Untersuchungen über die Grundlagen der Thermodynamik". Mathematische Annalen. 67 (3): 355–386. doi:10.1007/BF01450409. S2CID 118230148.
  69. ^ Swendsen, Robert (March 2006). (PDF). American Journal of Physics. 74 (3): 187–190. Bibcode:2006AmJPh..74..187S. doi:10.1119/1.2174962. S2CID 59471273. Archived from the original (PDF) on 2020-02-28.
  70. ^ Balescu, R. (1975). Equilibrium and Nonequilibrium Statistical Mechanics, Wiley, New York, ISBN 0-471-04600-0, pp. 148–154.
  71. ^ a b Kittel, Charles; Kroemer, Herbert (1980). Thermal Physics (2nd ed.). W.H. Freeman Company. pp. 391–397. ISBN 978-0-7167-1088-2.
  72. ^ Kondepudi, D.K. (1987). "Microscopic aspects implied by the second law". Foundations of Physics. 17 (7): 713–722. Bibcode:1987FoPh...17..713K. doi:10.1007/BF01889544. S2CID 120576357.
  73. ^ The Feynman Lectures on Physics. 39–5 The ideal gas law
  74. ^ "2018 CODATA Value: molar gas constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. 20 May 2019. Retrieved 2019-05-20.
  75. ^ "Kinetic Theory". galileo.phys.virginia.edu. from the original on 16 July 2017. Retrieved 27 January 2018.
  76. ^ Tolman, R.C. (1938). The Principles of Statistical Mechanics, Oxford University Press, London, pp. 93, 655.
  77. ^ Peter Atkins, Julio de Paula (2006). Physical Chemistry (8 ed.). Oxford University Press. p. 9.
  78. ^ Maxwell, J.C. (1872). Theory of Heat, third edition, Longman's, Green & Co, London, p. 32.
  79. ^ Bailyn, M. (1994). A Survey of Thermodynamics, American Institute of Physics Press, New York, ISBN 0-88318-797-3, p. 23, "..., if a temperature gradient exists, ..., then a flow of heat, ..., must occur to achieve a uniform temperature."
  80. ^ Guggenheim, E.A. (1967). Thermodynamics. An Advanced Treatment for Chemists and Physicists, North-Holland Publishing Company., Amsterdam, (1st edition 1949) fifth edition 1965, p. 8, "... will gradually adjust themselves until eventually they do reach mutual equilibrium after which there will of course be no further change."
  81. ^ Bailyn, M. (1994). A Survey of Thermodynamics, American Institute of Physics Press, New York, ISBN 0-88318-797-3, p. 22.
  82. ^ Guggenheim, E.A. (1967). Thermodynamics. An Advanced Treatment for Chemists and Physicists, North-Holland Publishing Company., Amsterdam, (1st edition 1949) fifth edition 1965, p. 8: "If two systems are both in thermal equilibrium with a third system then they are in thermal equilibrium with each other."
  83. ^ Buchdahl, H.A. (1966). The Concepts of Classical Thermodynamics, Cambridge University Press, Cambridge, p. 29: "... if each of two systems is in equilibrium with a third system then they are in equilibrium with each other."
  84. ^ a b Planck, M. (1945). Treatise on Thermodynamics. Dover Publications. p. §90 & §137. eqs.(39), (40), & (65).
  85. ^ Prati, E. (2010). "The finite quantum grand canonical ensemble and temperature from single-electron statistics for a mesoscopic device". J. Stat. Mech. 1 (1): P01003. arXiv:1001.2342. Bibcode:2010JSMTE..01..003P. doi:10.1088/1742-5468/2010/01/P01003. S2CID 118339343. arxiv.org 2017-11-22 at the Wayback Machine
  86. ^ "Archived copy" (PDF). (PDF) from the original on 2014-04-13. Retrieved 2014-04-11.{{cite web}}: CS1 maint: archived copy as title (link)
  87. ^ Prati, E.; et al. (2010). "Measuring the temperature of a mesoscopic electron system by means of single electron statistics". Applied Physics Letters. 96 (11): 113109. arXiv:1002.0037. Bibcode:2010ApPhL..96k3109P. doi:10.1063/1.3365204. S2CID 119209143. Archived from the original on 2016-05-14. Retrieved 2022-03-02. arxiv.org 2017-11-22 at the Wayback Machine
  88. ^ Water Science School. "Frozen carbon dioxide (dry ice) sublimates directly into a vapor". USGS.
  89. ^ Guggenheim, E.A. (1967) [1949], Thermodynamics. An Advanced Treatment for Chemists and Physicists (fifth ed.), Amsterdam: North-Holland Publishing Company., p. 157: "It is impossible by any procedure, no matter how idealized, to reduce the temperature of any system to zero temperature in a finite number of finite operations."
  90. ^ Pippard, A.B. (1957/1966). Elements of Classical Thermodynamics for Advanced Students of Physics, original publication 1957, reprint 1966, Cambridge University Press, Cambridge, page 51: "By no finite series of processes is the absolute zero attainable."
  91. ^ Tisza, L. (1966). Generalized Thermodynamics, MIT Press, Cambridge MA, page 96: "It is impossible to reach absolute zero as a result of a finite sequence of operations."
  92. ^ Kittel, Charles; Kroemer, Herbert (1980). Thermal Physics (2nd ed.). W.H. Freeman Company. p. Appendix E. ISBN 978-0-7167-1088-2.
  93. ^ The cited emission wavelengths are for black bodies in equilibrium. CODATA 2006 recommended value of 2.8977685(51)×10−3 m K used for Wien displacement law constant b.
  94. ^ This the Hawking Radiation for a Schwarzschild black hole of mass M = 4.145×106 M. It is too faint to be observed.
  95. ^ "World record in low temperatures". from the original on 2009-06-18. Retrieved 2009-05-05.
  96. ^ A temperature of 450 ±80 pK in a Bose–Einstein condensate (BEC) of sodium atoms was achieved in 2003 by researchers at MIT. Citation: Cooling Bose–Einstein Condensates Below 500 Picokelvin, A.E. Leanhardt et al., Science 301, 12 Sept. 2003, p. 1515. It's noteworthy that this record's peak emittance black-body wavelength of 6,400 kilometers is roughly the radius of Earth.
  97. ^ The peak emittance wavelength of 2.89777 m is a frequency of 103.456 MHz
  98. ^ Measurement was made in 2002 and has an uncertainty of ±3 kelvins. A 1989 measurement 2010-02-11 at the Wayback Machine produced a value of 5,777.0±2.5 K. Citation: (Chapter 1 lecture notes on Solar Physics by Division of Theoretical Physics, Dept. of Physical Sciences, University of Helsinki).
  99. ^ The 350 MK value is the maximum peak fusion fuel temperature in a thermonuclear weapon of the Teller–Ulam configuration (commonly known as a hydrogen bomb). Peak temperatures in Gadget-style fission bomb cores (commonly known as an atomic bomb) are in the range of 50 to 100 MK. Citation: Nuclear Weapons Frequently Asked Questions, 3.2.5 Matter At High Temperatures. Link to relevant Web page. 2007-05-03 at the Wayback Machine All referenced data was compiled from publicly available sources.
  100. ^ Peak temperature for a bulk quantity of matter was achieved by a pulsed-power machine used in fusion physics experiments. The term bulk quantity draws a distinction from collisions in particle accelerators wherein high temperature applies only to the debris from two subatomic particles or nuclei at any given instant. The >2 GK temperature was achieved over a period of about ten nanoseconds during shot Z1137. In fact, the iron and manganese ions in the plasma averaged 3.58±0.41 GK (309±35 keV) for 3 ns (ns 112 through 115). Ion Viscous Heating in a Magnetohydrodynamically Unstable Z Pinch at Over 2×109 Kelvin, M.G. Haines et al., Physical Review Letters 96 (2006) 075003. Link to Sandia's news release. 2010-05-30 at the Wayback Machine
  101. ^ Core temperature of a high–mass (>8–11 solar masses) star after it leaves the main sequence on the Hertzsprung–Russell diagram and begins the alpha process (which lasts one day) of fusing silicon–28 into heavier elements in the following steps: sulfur–32 → argon–36 → calcium–40 → titanium–44 → chromium–48 → iron–52 → nickel–56. Within minutes of finishing the sequence, the star explodes as a Type II supernova. Citation: Stellar Evolution: The Life and Death of Our Luminous Neighbors (by Arthur Holland and Mark Williams of the University of Michigan). Link to Web site 2009-01-16 at the Wayback Machine. More informative links can be found here . Archived from the original on 2013-04-11. Retrieved 2016-02-08., and here . Archived from the original on 2011-08-14. Retrieved 2016-02-08., and a concise treatise on stars by NASA is here . Archived from the original on 2010-10-24. Retrieved 2010-10-12.. . Archived from the original on January 16, 2009. Retrieved 2010-10-12.{{cite web}}: CS1 maint: bot: original URL status unknown (link)
  102. ^ Based on a computer model that predicted a peak internal temperature of 30 MeV (350 GK) during the merger of a binary neutron star system (which produces a gamma–ray burst). The neutron stars in the model were 1.2 and 1.6 solar masses respectively, were roughly 20 km in diameter, and were orbiting around their barycenter (common center of mass) at about 390 Hz during the last several milliseconds before they completely merged. The 350 GK portion was a small volume located at the pair's developing common core and varied from roughly 1 to 7 km across over a time span of around 5 ms. Imagine two city-sized objects of unimaginable density orbiting each other at the same frequency as the G4 musical note (the 28th white key on a piano). It's also noteworthy that at 350 GK, the average neutron has a vibrational speed of 30% the speed of light and a relativistic mass (m) 5% greater than its rest mass (m0).  Torus Formation in Neutron Star Mergers and Well-Localized Short Gamma-Ray Bursts 2017-11-22 at the Wayback Machine, R. Oechslin et al. of Max Planck Institute for Astrophysics. 2005-04-03 at the Wayback Machine, arXiv:astro-ph/0507099 v2, 22 Feb. 2006. An html summary 2010-11-09 at the Wayback Machine.
  103. ^ Results of research by Stefan Bathe using the PHENIX 2008-11-20 at the Wayback Machine detector on the Relativistic Heavy Ion Collider 2016-03-03 at the Wayback Machine at Brookhaven National Laboratory 2012-06-24 at the Wayback Machine in Upton, New York. Bathe has studied gold-gold, deuteron-gold, and proton-proton collisions to test the theory of quantum chromodynamics, the theory of the strong force that holds atomic nuclei together. Link to news release. 2009-02-11 at the Wayback Machine
  104. ^ How do physicists study particles? 2007-10-11 at the Wayback Machine by CERN 2012-07-07 at the Wayback Machine.

Bibliography of cited references

  • Adkins, C.J. (1968/1983). Equilibrium Thermodynamics, (1st edition 1968), third edition 1983, Cambridge University Press, Cambridge UK, ISBN 0-521-25445-0.
  • Buchdahl, H.A. (1966). The Concepts of Classical Thermodynamics, Cambridge University Press, Cambridge.
  • Jaynes, E.T. (1965). Gibbs vs Boltzmann entropies, American Journal of Physics, 33(5), 391–398.
  • Middleton, W.E.K. (1966). A History of the Thermometer and its Use in Metrology, Johns Hopkins Press, Baltimore.
  • Miller, J (2013). "Cooling molecules the optoelectric way". Physics Today. 66 (1): 12–14. Bibcode:2013PhT....66a..12M. doi:10.1063/pt.3.1840. Archived from the original on 2016-05-15. Retrieved 2013-07-25.
  • Partington, J.R. (1949). An Advanced Treatise on Physical Chemistry, volume 1, Fundamental Principles. The Properties of Gases, Longmans, Green & Co., London, pp. 175–177.
  • Pippard, A.B. (1957/1966). Elements of Classical Thermodynamics for Advanced Students of Physics, original publication 1957, reprint 1966, Cambridge University Press, Cambridge UK.
  • Quinn, T.J. (1983). Temperature, Academic Press, London, ISBN 0-12-569680-9.
  • Schooley, J.F. (1986). Thermometry, CRC Press, Boca Raton, ISBN 0-8493-5833-7.
  • Roberts, J.K., Miller, A.R. (1928/1960). Heat and Thermodynamics, (first edition 1928), fifth edition, Blackie & Son Limited, Glasgow.
  • Thomson, W. (Lord Kelvin) (1848). On an absolute thermometric scale founded on Carnot's theory of the motive power of heat, and calculated from Regnault's observations, Proc. Camb. Phil. Soc. (1843/1863) 1, No. 5: 66–71.
  • Thomson, W. (Lord Kelvin) (March 1851). "On the Dynamical Theory of Heat, with numerical results deduced from Mr Joule's equivalent of a Thermal Unit, and M. Regnault's Observations on Steam". Transactions of the Royal Society of Edinburgh. XX (part II): 261–268, 289–298.
  • Truesdell, C.A. (1980). The Tragicomical History of Thermodynamics, 1822–1854, Springer, New York, ISBN 0-387-90403-4.
  • Tschoegl, N.W. (2000). Fundamentals of Equilibrium and Steady-State Thermodynamics, Elsevier, Amsterdam, ISBN 0-444-50426-5.
  • Zeppenfeld, M.; Englert, B.G.U.; Glöckner, R.; Prehn, A.; Mielenz, M.; Sommer, C.; van Buuren, L.D.; Motsch, M.; Rempe, G. (2012). "Sysiphus cooling of electrically trapped polyatomic molecules". Nature. 491 (7425): 570–573. arXiv:1208.0046. Bibcode:2012Natur.491..570Z. doi:10.1038/nature11595. PMID 23151480. S2CID 4367940.

Further reading

  • Chang, Hasok (2004). Inventing Temperature: Measurement and Scientific Progress. Oxford: Oxford University Press. ISBN 978-0-19-517127-3.
  • Zemansky, Mark Waldo (1964). Temperatures Very Low and Very High. Princeton, NJ: Van Nostrand.

External links

  • Current map of global surface temperatures

temperature, this, article, about, physical, quantity, other, uses, disambiguation, physical, quantity, that, expresses, quantitatively, perceptions, hotness, coldness, measured, with, thermometer, thermal, vibration, segment, protein, alpha, helix, amplitude,. This article is about the physical quantity For other uses see Temperature disambiguation Temperature is a physical quantity that expresses quantitatively the perceptions of hotness and coldness Temperature is measured with a thermometer TemperatureThermal vibration of a segment of protein alpha helix Its amplitude increases with temperatureCommon symbolsTSI unitKOther units C F R Ro Re N D L WIntensive YesDerivations fromother quantitiesp V n R displaystyle frac pV nR d q rev d S displaystyle frac dq text rev dS Dimension8 displaystyle mathsf Theta Thermometers are calibrated in various temperature scales that historically have relied on various reference points and thermometric substances for definition The most common scales are the Celsius scale with the unit symbol C formerly called centigrade the Fahrenheit scale F and the Kelvin scale K the latter being used predominantly for scientific purposes The kelvin is one of the seven base units in the International System of Units SI Absolute zero i e zero kelvin or 273 15 C is the lowest point in the thermodynamic temperature scale Experimentally it can be approached very closely but not actually reached as recognized in the third law of thermodynamics It would be impossible to extract energy as heat from a body at that temperature Temperature is important in all fields of natural science including physics chemistry Earth science astronomy medicine biology ecology material science metallurgy mechanical engineering and geography as well as most aspects of daily life Contents 1 Effects 2 Scales 2 1 Absolute zero 2 2 Absolute scales 2 3 Kelvin scale 2 4 Statistical mechanical versus thermodynamic temperature scales 3 Classification of scales 3 1 Empirical scales 3 2 Theoretical scales 3 2 1 Microscopic statistical mechanical scale 3 2 2 Macroscopic thermodynamic scale 3 2 3 Ideal gas 4 Kinetic theory approach 5 Thermodynamic approach 5 1 Intensive variability 5 2 Local thermodynamic equilibrium 6 Basic theory 6 1 Bodies in thermodynamic equilibrium 6 2 Bodies in a steady state but not in thermodynamic equilibrium 6 3 Bodies not in a steady state 6 4 Thermodynamic equilibrium axiomatics 7 Heat capacity 8 Measurement 8 1 Units 8 1 1 Historical scales 8 1 2 Plasma physics 9 Theoretical foundation 9 1 Kinetic theory of gases 9 2 Zeroth law of thermodynamics 9 3 Second law of thermodynamics 9 4 Definition from statistical mechanics 9 5 Generalized temperature from single particle statistics 9 6 Negative temperature 10 Examples 11 See also 12 Notes and references 12 1 Bibliography of cited references 13 Further reading 14 External linksEffects Edit Average daily variation in human body temperatureMany physical processes are related to temperature some of them are given below the physical properties of materials including the phase solid liquid gaseous or plasma density solubility vapor pressure electrical conductivity hardness wear resistance thermal conductivity corrosion resistance strength the rate and extent to which chemical reactions occur 1 the amount and properties of thermal radiation emitted from the surface of an object air temperature affects all living organisms the speed of sound which in a gas is proportional to the square root of the absolute temperature 2 Scales EditMain article Scale of temperature Two thermometers showing temperature in Celsius and Fahrenheit This section needs additional citations for verification Please help improve this article by adding citations to reliable sources Unsourced material may be challenged and removed January 2021 Learn how and when to remove this template message Temperature scales need two values for definition the point chosen as zero degrees and the magnitudes of the incremental unit of temperature The Celsius scale C is used for common temperature measurements in most of the world It is an empirical scale that developed historically which led to its zero point 0 C being defined as the freezing point of water and 100 C as the boiling point of water both at atmospheric pressure at sea level It was called a centigrade scale because of the 100 degree interval 3 Since the standardization of the kelvin in the International System of Units it has subsequently been redefined in terms of the equivalent fixing points on the Kelvin scale so that a temperature increment of one degree Celsius is the same as an increment of one kelvin though numerically the scales differ by an exact offset of 273 15 The Fahrenheit scale is in common use in the United States Water freezes at 32 F and boils at 212 F at sea level atmospheric pressure Absolute zero Edit At the absolute zero of temperature no energy can be removed from matter as heat a fact expressed in the third law of thermodynamics At this temperature matter contains no macroscopic thermal energy but still has quantum mechanical zero point energy as predicted by the uncertainty principle although this does not enter into the definition of absolute temperature Experimentally absolute zero can be approached only very closely it can never be reached the lowest temperature attained by experiment is 100 pK citation needed Theoretically in a body at a temperature of absolute zero all classical motion of its particles has ceased and they are at complete rest in this classical sense The absolute zero defined as 0 K is exactly equal to 273 15 C or 459 67 F Absolute scales Edit Referring to the Boltzmann constant to the Maxwell Boltzmann distribution and to the Boltzmann statistical mechanical definition of entropy as distinct from the Gibbs definition 4 for independently moving microscopic particles disregarding interparticle potential energy by international agreement a temperature scale is defined and said to be absolute because it is independent of the characteristics of particular thermometric substances and thermometer mechanisms Apart from the absolute zero it does not have a reference temperature It is known as the Kelvin scale widely used in science and technology The kelvin the unit name is spelled with a lower case k is the unit of temperature in the International System of Units SI The temperature of a body in a state of thermodynamic equilibrium is always positive relative to the absolute zero Besides the internationally agreed Kelvin scale there is also a thermodynamic temperature scale invented by Lord Kelvin also with its numerical zero at the absolute zero of temperature but directly relating to purely macroscopic thermodynamic concepts including the macroscopic entropy though microscopically referable to the Gibbs statistical mechanical definition of entropy for the canonical ensemble that takes interparticle potential energy into account as well as independent particle motion so that it can account for measurements of temperatures near absolute zero 4 This scale has a reference temperature at the triple point of water the numerical value of which is defined by measurements using the aforementioned internationally agreed Kelvin scale Kelvin scale Edit Many scientific measurements use the Kelvin temperature scale unit symbol K named in honor of the physicist who first defined it It is an absolute scale Its numerical zero point 0 K is at the absolute zero of temperature Since May 2019 the kelvin has been defined through particle kinetic theory and statistical mechanics In the International System of Units SI the magnitude of the kelvin is defined in terms of the Boltzmann constant the value of which is defined as fixed by international convention 5 6 Statistical mechanical versus thermodynamic temperature scales Edit Since May 2019 the magnitude of the kelvin is defined in relation to microscopic phenomena characterized in terms of statistical mechanics Previously but since 1954 the International System of Units defined a scale and unit for the kelvin as a thermodynamic temperature by using the reliably reproducible temperature of the triple point of water as a second reference point the first reference point being 0 K at absolute zero citation needed Historically the temperature of the triple point of water was defined as exactly 273 16 K Today it is an empirically measured quantity The freezing point of water at sea level atmospheric pressure occurs at very close to 273 15 K 0 C Classification of scales EditThere are various kinds of temperature scale It may be convenient to classify them as empirically and theoretically based Empirical temperature scales are historically older while theoretically based scales arose in the middle of the nineteenth century 7 8 Empirical scales Edit Empirically based temperature scales rely directly on measurements of simple macroscopic physical properties of materials For example the length of a column of mercury confined in a glass walled capillary tube is dependent largely on temperature and is the basis of the very useful mercury in glass thermometer Such scales are valid only within convenient ranges of temperature For example above the boiling point of mercury a mercury in glass thermometer is impracticable Most materials expand with temperature increase but some materials such as water contract with temperature increase over some specific range and then they are hardly useful as thermometric materials A material is of no use as a thermometer near one of its phase change temperatures for example its boiling point In spite of these limitations most generally used practical thermometers are of the empirically based kind Especially it was used for calorimetry which contributed greatly to the discovery of thermodynamics Nevertheless empirical thermometry has serious drawbacks when judged as a basis for theoretical physics Empirically based thermometers beyond their base as simple direct measurements of ordinary physical properties of thermometric materials can be re calibrated by use of theoretical physical reasoning and this can extend their range of adequacy Theoretical scales Edit Theoretically based temperature scales are based directly on theoretical arguments especially those of kinetic theory and thermodynamics They are more or less ideally realized in practically feasible physical devices and materials Theoretically based temperature scales are used to provide calibrating standards for practical empirically based thermometers Microscopic statistical mechanical scale Edit In physics the internationally agreed conventional temperature scale is called the Kelvin scale It is calibrated through the internationally agreed and prescribed value of the Boltzmann constant 5 6 referring to motions of microscopic particles such as atoms molecules and electrons constituent in the body whose temperature is to be measured In contrast with the thermodynamic temperature scale invented by Kelvin the presently conventional Kelvin temperature is not defined through comparison with the temperature of a reference state of a standard body nor in terms of macroscopic thermodynamics Apart from the absolute zero of temperature the Kelvin temperature of a body in a state of internal thermodynamic equilibrium is defined by measurements of suitably chosen of its physical properties such as have precisely known theoretical explanations in terms of the Boltzmann constant citation needed That constant refers to chosen kinds of motion of microscopic particles in the constitution of the body In those kinds of motion the particles move individually without mutual interaction Such motions are typically interrupted by inter particle collisions but for temperature measurement the motions are chosen so that between collisions the non interactive segments of their trajectories are known to be accessible to accurate measurement For this purpose interparticle potential energy is disregarded In an ideal gas and in other theoretically understood bodies the Kelvin temperature is defined to be proportional to the average kinetic energy of non interactively moving microscopic particles which can be measured by suitable techniques The proportionality constant is a simple multiple of the Boltzmann constant If molecules atoms or electrons 9 10 are emitted from material and their velocities are measured the spectrum of their velocities often nearly obeys a theoretical law called the Maxwell Boltzmann distribution which gives a well founded measurement of temperatures for which the law holds 11 There have not yet been successful experiments of this same kind that directly use the Fermi Dirac distribution for thermometry but perhaps that will be achieved in the future 12 The speed of sound in a gas can be calculated theoretically from the molecular character of the gas from its temperature and pressure and from the value of the Boltzmann constant For a gas of known molecular character and pressure this provides a relation between temperature and the Boltzmann constant Those quantities can be known or measured more precisely than can the thermodynamic variables that define the state of a sample of water at its triple point Consequently taking the value of the Boltzmann constant as a primarily defined reference of exactly defined value a measurement of the speed of sound can provide a more precise measurement of the temperature of the gas 13 Measurement of the spectrum of electromagnetic radiation from an ideal three dimensional black body can provide an accurate temperature measurement because the frequency of maximum spectral radiance of black body radiation is directly proportional to the temperature of the black body this is known as Wien s displacement law and has a theoretical explanation in Planck s law and the Bose Einstein law Measurement of the spectrum of noise power produced by an electrical resistor can also provide accurate temperature measurement The resistor has two terminals and is in effect a one dimensional body The Bose Einstein law for this case indicates that the noise power is directly proportional to the temperature of the resistor and to the value of its resistance and to the noise bandwidth In a given frequency band the noise power has equal contributions from every frequency and is called Johnson noise If the value of the resistance is known then the temperature can be found 14 15 Macroscopic thermodynamic scale Edit Historically till May 2019 the definition of the Kelvin scale was that invented by Kelvin based on a ratio of quantities of energy in processes in an ideal Carnot engine entirely in terms of macroscopic thermodynamics citation needed That Carnot engine was to work between two temperatures that of the body whose temperature was to be measured and a reference that of a body at the temperature of the triple point of water Then the reference temperature that of the triple point was defined to be exactly 273 16 K Since May 2019 that value has not been fixed by definition but is to be measured through microscopic phenomena involving the Boltzmann constant as described above The microscopic statistical mechanical definition does not have a reference temperature Ideal gas Edit A material on which a macroscopically defined temperature scale may be based is the ideal gas The pressure exerted by a fixed volume and mass of an ideal gas is directly proportional to its temperature Some natural gases show so nearly ideal properties over suitable temperature range that they can be used for thermometry this was important during the development of thermodynamics and is still of practical importance today 16 17 The ideal gas thermometer is however not theoretically perfect for thermodynamics This is because the entropy of an ideal gas at its absolute zero of temperature is not a positive semi definite quantity which puts the gas in violation of the third law of thermodynamics In contrast to real materials the ideal gas does not liquefy or solidify no matter how cold it is Alternatively thinking the ideal gas law refers to the limit of infinitely high temperature and zero pressure these conditions guarantee non interactive motions of the constituent molecules 18 19 20 Kinetic theory approach EditThe magnitude of the kelvin is now defined in terms of kinetic theory derived from the value of the Boltzmann constant Kinetic theory provides a microscopic account of temperature for some bodies of material especially gases based on macroscopic systems being composed of many microscopic particles such as molecules and ions of various species the particles of a species being all alike It explains macroscopic phenomena through the classical mechanics of the microscopic particles The equipartition theorem of kinetic theory asserts that each classical degree of freedom of a freely moving particle has an average kinetic energy of kBT 2 where kB denotes the Boltzmann constant citation needed The translational motion of the particle has three degrees of freedom so that except at very low temperatures where quantum effects predominate the average translational kinetic energy of a freely moving particle in a system with temperature T will be 3kBT 2 Molecules such as oxygen O2 have more degrees of freedom than single spherical atoms they undergo rotational and vibrational motions as well as translations Heating results in an increase of temperature due to an increase in the average translational kinetic energy of the molecules Heating will also cause through equipartitioning the energy associated with vibrational and rotational modes to increase Thus a diatomic gas will require more energy input to increase its temperature by a certain amount i e it will have a greater heat capacity than a monatomic gas As noted above the speed of sound in a gas can be calculated from the molecular character of the gas from its temperature and pressure and from the value of the Boltzmann constant Taking the value of the Boltzmann constant as a primarily defined reference of exactly defined value a measurement of the speed of sound can provide a more precise measurement of the temperature of the gas 13 It is possible to measure the average kinetic energy of constituent microscopic particles if they are allowed to escape from the bulk of the system through a small hole in the containing wall The spectrum of velocities has to be measured and the average calculated from that It is not necessarily the case that the particles that escape and are measured have the same velocity distribution as the particles that remain in the bulk of the system but sometimes a good sample is possible Thermodynamic approach EditTemperature is one of the principal quantities in the study of thermodynamics Formerly the magnitude of the kelvin was defined in thermodynamic terms but nowadays as mentioned above it is defined in terms of kinetic theory The thermodynamic temperature is said to be absolute for two reasons One is that its formal character is independent of the properties of particular materials The other reason is that its zero is in a sense absolute in that it indicates absence of microscopic classical motion of the constituent particles of matter so that they have a limiting specific heat of zero for zero temperature according to the third law of thermodynamics Nevertheless a thermodynamic temperature does in fact have a definite numerical value that has been arbitrarily chosen by tradition and is dependent on the property of particular materials it is simply less arbitrary than relative degrees scales such as Celsius and Fahrenheit Being an absolute scale with one fixed point zero there is only one degree of freedom left to arbitrary choice rather than two as in relative scales For the Kelvin scale since May 2019 by international convention the choice has been made to use knowledge of modes of operation of various thermometric devices relying on microscopic kinetic theories about molecular motion The numerical scale is settled by a conventional definition of the value of the Boltzmann constant which relates macroscopic temperature to average microscopic kinetic energy of particles such as molecules Its numerical value is arbitrary and an alternate less widely used absolute temperature scale exists called the Rankine scale made to be aligned with the Fahrenheit scale as Kelvin is with Celsius The thermodynamic definition of temperature is due to Kelvin It is framed in terms of an idealized device called a Carnot engine imagined to run in a fictive continuous cycle of successive processes that traverse a cycle of states of its working body The engine takes in a quantity of heat Q1 from a hot reservoir and passes out a lesser quantity of waste heat Q2 lt 0 to a cold reservoir The net heat energy absorbed by the working body is passed as thermodynamic work to a work reservoir and is considered to be the output of the engine The cycle is imagined to run so slowly that at each point of the cycle the working body is in a state of thermodynamic equilibrium The successive processes of the cycle are thus imagined to run reversibly with no entropy production Then the quantity of entropy taken in from the hot reservoir when the working body is heated is equal to that passed to the cold reservoir when the working body is cooled Then the absolute or thermodynamic temperatures T1 and T2 of the reservoirs are defined such that 21 T 1 T 2 Q 1 Q 2 displaystyle frac T 1 T 2 frac Q 1 Q 2 1 The zeroth law of thermodynamics allows this definition to be used to measure the absolute or thermodynamic temperature of an arbitrary body of interest by making the other heat reservoir have the same temperature as the body of interest Kelvin s original work postulating absolute temperature was published in 1848 It was based on the work of Carnot before the formulation of the first law of thermodynamics Carnot had no sound understanding of heat and no specific concept of entropy He wrote of caloric and said that all the caloric that passed from the hot reservoir was passed into the cold reservoir Kelvin wrote in his 1848 paper that his scale was absolute in the sense that it was defined independently of the properties of any particular kind of matter His definitive publication which sets out the definition just stated was printed in 1853 a paper read in 1851 22 23 24 25 Numerical details were formerly settled by making one of the heat reservoirs a cell at the triple point of water which was defined to have an absolute temperature of 273 16 K 26 Nowadays the numerical value is instead obtained from measurement through the microscopic statistical mechanical international definition as above Intensive variability Edit In thermodynamic terms temperature is an intensive variable because it is equal to a differential coefficient of one extensive variable with respect to another for a given body It thus has the dimensions of a ratio of two extensive variables In thermodynamics two bodies are often considered as connected by contact with a common wall which has some specific permeability properties Such specific permeability can be referred to a specific intensive variable An example is a diathermic wall that is permeable only to heat the intensive variable for this case is temperature When the two bodies have been connected through the specifically permeable wall for a very long time and have settled to a permanent steady state the relevant intensive variables are equal in the two bodies for a diathermal wall this statement is sometimes called the zeroth law of thermodynamics 27 28 29 In particular when the body is described by stating its internal energy U an extensive variable as a function of its entropy S also an extensive variable and other state variables V N with U U S V N then the temperature is equal to the partial derivative of the internal energy with respect to the entropy 28 29 30 T U S V N displaystyle T left frac partial U partial S right V N 2 Likewise when the body is described by stating its entropy S as a function of its internal energy U and other state variables V N with S S U V N then the reciprocal of the temperature is equal to the partial derivative of the entropy with respect to the internal energy 28 30 31 1 T S U V N displaystyle frac 1 T left frac partial S partial U right V N 3 The above definition equation 1 of the absolute temperature is due to Kelvin It refers to systems closed to the transfer of matter and has a special emphasis on directly experimental procedures A presentation of thermodynamics by Gibbs starts at a more abstract level and deals with systems open to the transfer of matter in this development of thermodynamics the equations 2 and 3 above are actually alternative definitions of temperature 32 Local thermodynamic equilibrium Edit Real world bodies are often not in thermodynamic equilibrium and not homogeneous For the study by methods of classical irreversible thermodynamics a body is usually spatially and temporally divided conceptually into cells of small size If classical thermodynamic equilibrium conditions for matter are fulfilled to good approximation in such a cell then it is homogeneous and a temperature exists for it If this is so for every cell of the body then local thermodynamic equilibrium is said to prevail throughout the body 33 34 35 36 37 It makes good sense for example to say of the extensive variable U or of the extensive variable S that it has a density per unit volume or a quantity per unit mass of the system but it makes no sense to speak of the density of temperature per unit volume or quantity of temperature per unit mass of the system On the other hand it makes no sense to speak of the internal energy at a point while when local thermodynamic equilibrium prevails it makes good sense to speak of the temperature at a point Consequently the temperature can vary from point to point in a medium that is not in global thermodynamic equilibrium but in which there is local thermodynamic equilibrium Thus when local thermodynamic equilibrium prevails in a body the temperature can be regarded as a spatially varying local property in that body and this is because the temperature is an intensive variable Basic theory EditTemperature is a measure of a quality of a state of a material 38 The quality may be regarded as a more abstract entity than any particular temperature scale that measures it and is called hotness by some writers 39 40 41 The quality of hotness refers to the state of material only in a particular locality and in general apart from bodies held in a steady state of thermodynamic equilibrium hotness varies from place to place It is not necessarily the case that a material in a particular place is in a state that is steady and nearly homogeneous enough to allow it to have a well defined hotness or temperature Hotness may be represented abstractly as a one dimensional manifold Every valid temperature scale has its own one to one map into the hotness manifold 42 43 When two systems in thermal contact are at the same temperature no heat transfers between them When a temperature difference does exist heat flows spontaneously from the warmer system to the colder system until they are in thermal equilibrium Such heat transfer occurs by conduction or by thermal radiation 44 45 46 47 48 49 50 51 Experimental physicists for example Galileo and Newton 52 found that there are indefinitely many empirical temperature scales Nevertheless the zeroth law of thermodynamics says that they all measure the same quality This means that for a body in its own state of internal thermodynamic equilibrium every correctly calibrated thermometer of whatever kind that measures the temperature of the body records one and the same temperature For a body that is not in its own state of internal thermodynamic equilibrium different thermometers can record different temperatures depending respectively on the mechanisms of operation of the thermometers Bodies in thermodynamic equilibrium Edit For experimental physics hotness means that when comparing any two given bodies in their respective separate thermodynamic equilibria any two suitably given empirical thermometers with numerical scale readings will agree as to which is the hotter of the two given bodies or that they have the same temperature 53 This does not require the two thermometers to have a linear relation between their numerical scale readings but it does require that the relation between their numerical readings shall be strictly monotonic 54 55 A definite sense of greater hotness can be had independently of calorimetry of thermodynamics and of properties of particular materials from Wien s displacement law of thermal radiation the temperature of a bath of thermal radiation is proportional by a universal constant to the frequency of the maximum of its frequency spectrum this frequency is always positive but can have values that tend to zero Thermal radiation is initially defined for a cavity in thermodynamic equilibrium These physical facts justify a mathematical statement that hotness exists on an ordered one dimensional manifold This is a fundamental character of temperature and thermometers for bodies in their own thermodynamic equilibrium 7 42 43 56 57 Except for a system undergoing a first order phase change such as the melting of ice as a closed system receives heat without a change in its volume and without a change in external force fields acting on it its temperature rises For a system undergoing such a phase change so slowly that departure from thermodynamic equilibrium can be neglected its temperature remains constant as the system is supplied with latent heat Conversely a loss of heat from a closed system without phase change without change of volume and without a change in external force fields acting on it decreases its temperature 58 Bodies in a steady state but not in thermodynamic equilibrium Edit While for bodies in their own thermodynamic equilibrium states the notion of temperature requires that all empirical thermometers must agree as to which of two bodies is the hotter or that they are at the same temperature this requirement is not safe for bodies that are in steady states though not in thermodynamic equilibrium It can then well be that different empirical thermometers disagree about which is hotter and if this is so then at least one of the bodies does not have a well defined absolute thermodynamic temperature Nevertheless anyone has given body and any one suitable empirical thermometer can still support notions of empirical non absolute hotness and temperature for a suitable range of processes This is a matter for study in non equilibrium thermodynamics citation needed Bodies not in a steady state Edit When a body is not in a steady state then the notion of temperature becomes even less safe than for a body in a steady state not in thermodynamic equilibrium This is also a matter for study in non equilibrium thermodynamics Thermodynamic equilibrium axiomatics Edit For the axiomatic treatment of thermodynamic equilibrium since the 1930s it has become customary to refer to a zeroth law of thermodynamics The customarily stated minimalist version of such a law postulates only that all bodies which when thermally connected would be in thermal equilibrium should be said to have the same temperature by definition but by itself does not establish temperature as a quantity expressed as a real number on a scale A more physically informative version of such a law views empirical temperature as a chart on a hotness manifold 42 57 59 While the zeroth law permits the definitions of many different empirical scales of temperature the second law of thermodynamics selects the definition of a single preferred absolute temperature unique up to an arbitrary scale factor whence called the thermodynamic temperature 7 42 60 61 62 63 If internal energy is considered as a function of the volume and entropy of a homogeneous system in thermodynamic equilibrium thermodynamic absolute temperature appears as the partial derivative of internal energy with respect the entropy at constant volume Its natural intrinsic origin or null point is absolute zero at which the entropy of any system is at a minimum Although this is the lowest absolute temperature described by the model the third law of thermodynamics postulates that absolute zero cannot be attained by any physical system Heat capacity EditSee also Heat capacity and Calorimetry When an energy transfer to or from a body is only as heat the state of the body changes Depending on the surroundings and the walls separating them from the body various changes are possible in the body They include chemical reactions increase of pressure increase of temperature and phase change For each kind of change under specified conditions the heat capacity is the ratio of the quantity of heat transferred to the magnitude of the change 64 For example if the change is an increase in temperature at constant volume with no phase change and no chemical change then the temperature of the body rises and its pressure increases The quantity of heat transferred DQ divided by the observed temperature change DT is the body s heat capacity at constant volume C V D Q D T displaystyle C V frac Delta Q Delta T If heat capacity is measured for a well defined amount of substance the specific heat is the measure of the heat required to increase the temperature of such a unit quantity by one unit of temperature For example raising the temperature of water by one kelvin equal to one degree Celsius requires 4186 joules per kilogram J kg Measurement Edit A typical Celsius thermometer measures a winter day temperature of 17 C See also Timeline of temperature and pressure measurement technology International Temperature Scale of 1990 and Comparison of temperature scales Temperature measurement using modern scientific thermometers and temperature scales goes back at least as far as the early 18th century when Daniel Gabriel Fahrenheit adapted a thermometer switching to mercury and a scale both developed by Ole Christensen Romer Fahrenheit s scale is still in use in the United States for non scientific applications Temperature is measured with thermometers that may be calibrated to a variety of temperature scales In most of the world except for Belize Myanmar Liberia and the United States the Celsius scale is used for most temperature measuring purposes Most scientists measure temperature using the Celsius scale and thermodynamic temperature using the Kelvin scale which is the Celsius scale offset so that its null point is 0 K 273 15 C or absolute zero Many engineering fields in the US notably high tech and US federal specifications civil and military also use the Kelvin and Celsius scales Other engineering fields in the US also rely upon the Rankine scale a shifted Fahrenheit scale when working in thermodynamic related disciplines such as combustion Units Edit The basic unit of temperature in the International System of Units SI is the kelvin It has the symbol K For everyday applications it is often convenient to use the Celsius scale in which 0 C corresponds very closely to the freezing point of water and 100 C is its boiling point at sea level Because liquid droplets commonly exist in clouds at sub zero temperatures 0 C is better defined as the melting point of ice In this scale a temperature difference of 1 degree Celsius is the same as a 1kelvin increment but the scale is offset by the temperature at which ice melts 273 15 K By international agreement 65 until May 2019 the Kelvin and Celsius scales were defined by two fixing points absolute zero and the triple point of Vienna Standard Mean Ocean Water which is water specially prepared with a specified blend of hydrogen and oxygen isotopes Absolute zero was defined as precisely 0 K and 273 15 C It is the temperature at which all classical translational motion of the particles comprising matter ceases and they are at complete rest in the classical model Quantum mechanically however zero point motion remains and has an associated energy the zero point energy Matter is in its ground state 66 and contains no thermal energy The temperatures 273 16 K and 0 01 C were defined as those of the triple point of water This definition served the following purposes it fixed the magnitude of the kelvin as being precisely 1 part in 273 16 parts of the difference between absolute zero and the triple point of water it established that one kelvin has precisely the same magnitude as one degree on the Celsius scale and it established the difference between the null points of these scales as being 273 15 K 0 K 273 15 C and 273 16 K 0 01 C Since 2019 there has been a new definition based on the Boltzmann constant 67 but the scales are scarcely changed In the United States the Fahrenheit scale is the most widely used On this scale the freezing point of water corresponds to 32 F and the boiling point to 212 F The Rankine scale still used in fields of chemical engineering in the US is an absolute scale based on the Fahrenheit increment Historical scales Edit See also Conversion of scales of temperature The following temperature scales are in use or have historically been used for measuring temperature Kelvin scale Celsius scale Fahrenheit scale Rankine scale Delisle scale Newton scale Reaumur scale Romer scalePlasma physics Edit The field of plasma physics deals with phenomena of electromagnetic nature that involve very high temperatures It is customary to express temperature as energy in a unit related to the electronvolt or kiloelectronvolt eV kB or keV kB The corresponding energy which is dimensionally distinct from temperature is then calculated as the product of the Boltzmann constant and temperature E k B T displaystyle E k text B T Then 1 eV kB is 11605 K In the study of QCD matter one routinely encounters temperatures of the order of a few hundred MeV kB equivalent to about 1012 K Theoretical foundation EditSee also Thermodynamic temperature Historically there are several scientific approaches to the explanation of temperature the classical thermodynamic description based on macroscopic empirical variables that can be measured in a laboratory the kinetic theory of gases which relates the macroscopic description to the probability distribution of the energy of motion of gas particles and a microscopic explanation based on statistical physics and quantum mechanics In addition rigorous and purely mathematical treatments have provided an axiomatic approach to classical thermodynamics and temperature 68 Statistical physics provides a deeper understanding by describing the atomic behavior of matter and derives macroscopic properties from statistical averages of microscopic states including both classical and quantum states In the fundamental physical description the temperature may be measured directly in units of energy However in the practical systems of measurement for science technology and commerce such as the modern metric system of units the macroscopic and the microscopic descriptions are interrelated by the Boltzmann constant a proportionality factor that scales temperature to the microscopic mean kinetic energy The microscopic description in statistical mechanics is based on a model that analyzes a system into its fundamental particles of matter or into a set of classical or quantum mechanical oscillators and considers the system as a statistical ensemble of microstates As a collection of classical material particles the temperature is a measure of the mean energy of motion called translational kinetic energy of the particles whether in solids liquids gases or plasmas The kinetic energy a concept of classical mechanics is half the mass of a particle times its speed squared In this mechanical interpretation of thermal motion the kinetic energies of material particles may reside in the velocity of the particles of their translational or vibrational motion or in the inertia of their rotational modes In monatomic perfect gases and approximately in most gas and in simple metals the temperature is a measure of the mean particle translational kinetic energy 3 2 kBT It also determines the probability distribution function of energy In condensed matter and particularly in solids this purely mechanical description is often less useful and the oscillator model provides a better description to account for quantum mechanical phenomena Temperature determines the statistical occupation of the microstates of the ensemble The microscopic definition of temperature is only meaningful in the thermodynamic limit meaning for large ensembles of states or particles to fulfill the requirements of the statistical model Kinetic energy is also considered as a component of thermal energy The thermal energy may be partitioned into independent components attributed to the degrees of freedom of the particles or to the modes of oscillators in a thermodynamic system In general the number of these degrees of freedom that are available for the equipartitioning of energy depends on the temperature i e the energy region of the interactions under consideration For solids the thermal energy is associated primarily with the vibrations of its atoms or molecules about their equilibrium position In an ideal monatomic gas the kinetic energy is found exclusively in the purely translational motions of the particles In other systems vibrational and rotational motions also contribute degrees of freedom Kinetic theory of gases Edit A theoretical understanding of temperature in a hard sphere model of a gas can be obtained from the Kinetic theory Maxwell and Boltzmann developed a kinetic theory that yields a fundamental understanding of temperature in gases 69 This theory also explains the ideal gas law and the observed heat capacity of monatomic or noble gases 70 71 72 Plots of pressure vs temperature for three different gas samples extrapolated to absolute zero The ideal gas law is based on observed empirical relationships between pressure p volume V and temperature T and was recognized long before the kinetic theory of gases was developed see Boyle s and Charles s laws The ideal gas law states 73 p V n R T displaystyle pV nRT where n is the number of moles of gas and R 8 314462 618 J mol 1 K 1 74 is the gas constant This relationship gives us our first hint that there is an absolute zero on the temperature scale because it only holds if the temperature is measured on an absolute scale such as Kelvin s The ideal gas law allows one to measure temperature on this absolute scale using the gas thermometer The temperature in kelvins can be defined as the pressure in pascals of one mole of gas in a container of one cubic meter divided by the gas constant Although it is not a particularly convenient device the gas thermometer provides an essential theoretical basis by which all thermometers can be calibrated As a practical matter it is not possible to use a gas thermometer to measure absolute zero temperature since the gases condense into a liquid long before the temperature reaches zero It is possible however to extrapolate to absolute zero by using the ideal gas law as shown in the figure The kinetic theory assumes that pressure is caused by the force associated with individual atoms striking the walls and that all energy is translational kinetic energy Using a sophisticated symmetry argument 75 Boltzmann deduced what is now called the Maxwell Boltzmann probability distribution function for the velocity of particles in an ideal gas From that probability distribution function the average kinetic energy per particle of a monatomic ideal gas is 71 76 E k 1 2 m v rms 2 3 2 k B T displaystyle E text k frac 1 2 mv text rms 2 frac 3 2 k text B T where the Boltzmann constant kB is the ideal gas constant divided by the Avogadro number and v rms v 2 v v textstyle v text rms sqrt langle v 2 rangle sqrt langle mathbf v cdot v rangle is the root mean square speed 77 This direct proportionality between temperature and mean molecular kinetic energy is a special case of the equipartition theorem and holds only in the classical limit of a perfect gas It does not hold exactly for most substances Zeroth law of thermodynamics Edit Main article Zeroth law of thermodynamics When two otherwise isolated bodies are connected together by a rigid physical path impermeable to matter there is the spontaneous transfer of energy as heat from the hotter to the colder of them Eventually they reach a state of mutual thermal equilibrium in which heat transfer has ceased and the bodies respective state variables have settled to become unchanging 78 79 80 One statement of the zeroth law of thermodynamics is that if two systems are each in thermal equilibrium with a third system then they are also in thermal equilibrium with each other 81 82 83 This statement helps to define temperature but it does not by itself complete the definition An empirical temperature is a numerical scale for the hotness of a thermodynamic system Such hotness may be defined as existing on a one dimensional manifold stretching between hot and cold Sometimes the zeroth law is stated to include the existence of a unique universal hotness manifold and of numerical scales on it so as to provide a complete definition of empirical temperature 59 To be suitable for empirical thermometry a material must have a monotonic relation between hotness and some easily measured state variable such as pressure or volume when all other relevant coordinates are fixed An exceptionally suitable system is the ideal gas which can provide a temperature scale that matches the absolute Kelvin scale The Kelvin scale is defined on the basis of the second law of thermodynamics Second law of thermodynamics Edit Main article Second law of thermodynamics As an alternative to considering or defining the zeroth law of thermodynamics it was the historical development in thermodynamics to define temperature in terms of the second law of thermodynamics which deals with entropy citation needed The second law states that any process will result in either no change or a net increase in the entropy of the universe This can be understood in terms of probability For example in a series of coin tosses a perfectly ordered system would be one in which either every toss comes up heads or every toss comes up tails This means the outcome is always 100 the same result In contrast many mixed disordered outcomes are possible and their number increases with each toss Eventually the combinations of 50 heads and 50 tails dominate and obtaining an outcome significantly different from 50 50 becomes increasingly unlikely Thus the system naturally progresses to a state of maximum disorder or entropy As temperature governs the transfer of heat between two systems and the universe tends to progress toward a maximum of entropy it is expected that there is some relationship between temperature and entropy A heat engine is a device for converting thermal energy into mechanical energy resulting in the performance of work An analysis of the Carnot heat engine provides the necessary relationships According to energy conservation and energy being a state function that does not change over a full cycle the work from a heat engine over a full cycle is equal to the net heat i e the sum of the heat put into the system at high temperature qH gt 0 and the waste heat given off at the low temperature qC lt 0 84 The efficiency is the work divided by the heat input efficiency w cy q H q H q C q H 1 q C q H displaystyle text efficiency frac w text cy q text H frac q text H q text C q text H 1 frac q text C q text H 4 where wcy is the work done per cycle The efficiency depends only on qC qH Because qC and qH correspond to heat transfer at the temperatures TC and TH respectively qC qH should be some function of these temperatures q C q H f T H T C displaystyle frac q text C q text H f left T text H T text C right 5 Carnot s theorem states that all reversible engines operating between the same heat reservoirs are equally efficient citation needed Thus a heat engine operating between T1 and T3 must have the same efficiency as one consisting of two cycles one between T1 and T2 and the second between T2 and T3 This can only be the case if q 13 q 1 q 2 q 2 q 3 displaystyle q 13 frac q 1 q 2 q 2 q 3 which implies q 13 f T 1 T 3 f T 1 T 2 f T 2 T 3 displaystyle q 13 f left T 1 T 3 right f left T 1 T 2 right f left T 2 T 3 right Since the first function is independent of T2 this temperature must cancel on the right side meaning f T1 T3 is of the form g T1 g T3 i e f T1 T3 f T1 T2 f T2 T3 g T1 g T2 g T2 g T3 g T1 g T3 where g is a function of a single temperature A temperature scale can now be chosen with the property that q C q H T C T H displaystyle frac q text C q text H frac T text C T text H 6 Substituting 6 back into 4 gives a relationship for the efficiency in terms of temperature efficiency 1 q C q H 1 T C T H displaystyle text efficiency 1 frac q text C q text H 1 frac T text C T text H 7 For TC 0 K the efficiency is 100 and that efficiency becomes greater than 100 below 0 K Since an efficiency greater than 100 violates the first law of thermodynamics this implies that 0 K is the minimum possible temperature In fact the lowest temperature ever obtained in a macroscopic system was 20 nK which was achieved in 1995 at NIST Subtracting the right hand side of 5 from the middle portion and rearranging gives 21 84 q H T H q C T C 0 displaystyle frac q text H T text H frac q text C T text C 0 where the negative sign indicates heat ejected from the system This relationship suggests the existence of a state function S whose change characteristically vanishes for a complete cycle if it is defined by d S d q rev T displaystyle dS frac dq text rev T 8 where the subscript indicates a reversible process This function corresponds to the entropy of the system which was described previously Rearranging 8 gives a formula for temperature in terms of fictive infinitesimal quasi reversible elements of entropy and heat T d q rev d S displaystyle T frac dq text rev dS 9 For a constant volume system where entropy S E is a function of its energy E dE dqrev and 9 gives T 1 d d E S E displaystyle T 1 frac d dE S E 10 i e the reciprocal of the temperature is the rate of increase of entropy with respect to energy at constant volume Definition from statistical mechanics Edit Statistical mechanics defines temperature based on a system s fundamental degrees of freedom Eq 10 is the defining relation of temperature where the entropy S displaystyle S is defined up to a constant by the logarithm of the number of microstates of the system in the given macrostate as specified in the microcanonical ensemble S k B ln W displaystyle S k mathrm B ln W where k B displaystyle k mathrm B is the Boltzmann constant and W is the number of microstates with the energy E of the system degeneracy When two systems with different temperatures are put into purely thermal connection heat will flow from the higher temperature system to the lower temperature one thermodynamically this is understood by the second law of thermodynamics The total change in entropy following a transfer of energy D E displaystyle Delta E from system 1 to system 2 is D S d S d E 1 D E d S d E 2 D E 1 T 2 1 T 1 D E displaystyle Delta S dS dE 1 cdot Delta E dS dE 2 cdot Delta E left frac 1 T 2 frac 1 T 1 right Delta E and is thus positive if T 1 gt T 2 displaystyle T 1 gt T 2 From the point of view of statistical mechanics the total number of microstates in the combined system 1 system 2 is N 1 N 2 displaystyle N 1 cdot N 2 the logarithm of which times the Boltzmann constant is the sum of their entropies thus a flow of heat from high to low temperature which brings an increase in total entropy is more likely than any other scenario normally it is much more likely as there are more microstates in the resulting macrostate Generalized temperature from single particle statistics Edit It is possible to extend the definition of temperature even to systems of few particles like in a quantum dot The generalized temperature is obtained by considering time ensembles instead of configuration space ensembles given in statistical mechanics in the case of thermal and particle exchange between a small system of fermions N even less than 10 with a single double occupancy system The finite quantum grand canonical ensemble 85 obtained under the hypothesis of ergodicity and orthodicity 86 allows expressing the generalized temperature from the ratio of the average time of occupation t 1 displaystyle tau 1 and t 2 displaystyle tau 2 of the single double occupancy system 87 T E E F 1 3 2 N k B ln 2 t 2 t 1 displaystyle T frac E E text F left 1 frac 3 2N right k text B ln left 2 frac tau 2 tau 1 right where EF is the Fermi energy This generalized temperature tends to the ordinary temperature when N goes to infinity Negative temperature Edit Main article Negative temperature On the empirical temperature scales that are not referenced to absolute zero a negative temperature is one below the zero point of the scale used For example dry ice has a sublimation temperature of 78 5 C which is equivalent to 109 3 F 88 On the absolute Kelvin scale this temperature is 194 6 K No body can be brought to exactly 0 K the temperature of the ideally coldest possible body by any finite practicable process this is a consequence of the third law of thermodynamics 89 90 91 The international kinetic theory temperature of a body cannot take negative values The thermodynamic temperature scale however is not so constrained For a body of matter there can sometimes be conceptually defined in terms of microscopic degrees of freedom namely particle spins a subsystem with a temperature other than that of the whole body When the body is in its own state of internal thermodynamic equilibrium the temperatures of the whole body and of the subsystem must be the same The two temperatures can differ when by work through externally imposed force fields energy can be transferred to and from the subsystem separately from the rest of the body then the whole body is not in its own state of internal thermodynamic equilibrium There is an upper limit of energy such a spin subsystem can attain Considering the subsystem to be in a temporary state of virtual thermodynamic equilibrium it is possible to obtain a negative temperature on the thermodynamic scale Thermodynamic temperature is the inverse of the derivative of the subsystem s entropy with respect to its internal energy As the subsystem s internal energy increases the entropy increases for some range but eventually attains a maximum value and then begins to decrease as the highest energy states begin to fill At the point of maximum entropy the temperature function shows the behavior of a singularity because the slope of the entropy as a function of energy decreases to zero and then turns negative As the subsystem s entropy reaches its maximum its thermodynamic temperature goes to positive infinity switching to negative infinity as the slope turns negative Such negative temperatures are hotter than any positive temperature Over time when the subsystem is exposed to the rest of the body which has a positive temperature energy is transferred as heat from the negative temperature subsystem to the positive temperature system 92 The kinetic theory temperature is not defined for such subsystems Examples EditMain article Orders of magnitude temperature Comparisons of temperatures in various scales Temperature Peak emittance wavelength 93 of black body radiationKelvin CelsiusAbsolute zero precisely by definition 0 K 273 15 C InfinityBlackbody temperature of the black hole atthe centre of our galaxy Sagittarius A 94 15 fK 273 149999 999 999 985 C 2 5 108 km 1 7 AU Lowest temperatureachieved 95 100 pK 273 149999 999 900 C 29000 kmColdest Bose Einstein condensate 96 450 pK 273 149999 999 55 C 6400 kmOne millikelvin precisely by definition 0 001 K 273 149 C 2 89777 m radio FM band 97 Cosmic microwave background 2013 measurement 2 7260 K 270 424 C 0 001063 01 m millimeter wavelength microwave Water triple point precisely by definition 273 16 K 0 01 C 10608 3 nm long wavelength IR Water boiling point A 373 1339 K 99 9839 C 7766 03 nm mid wavelength IR Iron melting point 1811 K 1538 C 1600 nm far infrared Incandescent lamp B 2500 K 2200 C 1160 nm near infrared C Sun s visible surface D 98 5778 K 5505 C 501 5 nm green blue light Lightning boltchannel E 28 kK 28000 C 100 nm far ultraviolet light Sun s core E 16 MK 16 million C 0 18 nm X rays Thermonuclear weapon peak temperature E 99 350 MK 350 million C 8 3 10 3 nm gamma rays Sandia National Labs Z machine E 100 2 GK 2 billion C 1 4 10 3 nm gamma rays F Core of a high massstar on its last day E 101 3 GK 3 billion C 1 10 3 nm gamma rays Merging binary neutronstar system E 102 350 GK 350 billion C 8 10 6 nm gamma rays Relativistic HeavyIon Collider E 103 1 TK 1 trillion C 3 10 6 nm gamma rays CERN s proton vs nucleus collisions E 104 10 TK 10 trillion C 3 10 7 nm gamma rays A For Vienna Standard Mean Ocean Water at one standard atmosphere 101 325 kPa when calibrated strictly per the two point definition of thermodynamic temperature B The 2500 K value is approximate The 273 15 K difference between K and C is rounded to 300 K to avoid false precision in the Celsius value C For a true black body which tungsten filaments are not Tungsten filament emissivity is greater at shorter wavelengths which makes them appear whiter D Effective photosphere temperature The 273 15 K difference between K and C is rounded to 273 K to avoid false precision in the Celsius value E The 273 15 K difference between K and C is within the precision of these values F For a true black body which the plasma was not The Z machine s dominant emission originated from 40 MK electrons soft x ray emissions within the plasma See also EditAtmospheric temperature Physical quantity that expresses hot and cold in the atmosphere Body temperature Ability of an organism to keep its body temperature within certain boundaries thermoregulation Color temperature Property of light sources related to black body radiation Dry bulb temperature Temperature of air as measured by a thermometer shielded from radiation and moisture Thermal conduction Process by which heat is transferred within an object Convective heat transfer Instrumental temperature record In situ measurements that provide the temperature of Earth s climate system ISO 1 ISO standard temperature 20 C International Temperature Scale of 1990 ITS 90 Practical temperature scale Laser schlieren deflectometry List of cities by average temperature Maxwell s demon Thought experiment of 1867 Orders of magnitude temperature Range of temperatures from absolute zero to very high Outside air temperature Planck temperature Rankine scale Absolute temperature scale using Fahrenheit degrees Relativistic heat conduction Model compatible with special relativity Satellite temperature measurements Scale of temperature Method to measure temperature quantitatively Sea surface temperature Water temperature close to the ocean s surface Stagnation temperature Thermal radiation Electromagnetic radiation generated by the thermal motion of particles Thermoception Temperature stimulus inferred from a heat flux and converted into a molecular signal Thermodynamic absolute temperature Measure of absolute temperature Thermography Use of thermograms to study heat distribution in structures or regions Thermometer Device to measure temperature Virtual temperature Wet bulb globe temperature Type of apparent temperature used to estimate the effect of temperature and other factors on humans Wet bulb temperature Temperature read by a thermometer covered in water soaked clothNotes and references Edit Agency International Atomic Energy 1974 Thermal discharges at nuclear power stations their management and environmental impacts a report prepared by a group of experts as the result of a panel meeting held in Vienna 23 27 October 1972 International Atomic Energy Agency Watkinson John 2001 The Art of Digital Audio Taylor amp Francis ISBN 978 0 240 51587 8 Middleton W E K 1966 pp 89 105 a b Jaynes E T 1965 pp 391 398 a b Cryogenic Society 2019 a b Draft Resolution A On the revision of the International System of Units SI to be submitted to the CGPM at its 26th meeting 2018 PDF archived from the original PDF on 2018 04 29 retrieved 2019 10 20 a b c Truesdell C A 1980 Sections 11 B 11H pp 306 310 320 332 Quinn T J 1983 Germer L H 1925 The distribution of initial velocities among thermionic electrons Phys Rev 25 795 807 here Turvey K 1990 Test of validity of Maxwellian statistics for electrons thermionically emitted from an oxide cathode European Journal of Physics 11 1 51 59 here Zeppenfeld M Englert B G U Glockner R Prehn A Mielenz M Sommer C van Buuren L D Motsch M Rempe G 2012 Miller J 2013 a b de Podesta M Underwood R Sutton G Morantz P Harris P Mark D F Stuart F M Vargha G Machin M 2013 A low uncertainty measurement of the Boltzmann constant Metrologia 50 4 S213 S216 BIPM amp IOP Publishing Ltd Quinn T J 1983 pp 98 107 Schooley J F 1986 pp 138 143 Quinn T J 1983 pp 61 83 Schooley J F 1986 pp 115 138 Adkins C J 1968 1983 pp 119 120 Buchdahl H A 1966 pp 137 138 Tschoegl N W 2000 p 88 a b Fermi E 1956 Thermodynamics Dover Publications still in print p 48 eq 64 Thomson W Lord Kelvin 1848 Thomson W Lord Kelvin 1851 Partington J R 1949 pp 175 177 Roberts J K Miller A R 1928 1960 pp 321 322 Quinn T J 1983 Temperature Academic Press London ISBN 0 12 569680 9 pp 160 162 Tisza L 1966 Generalized Thermodynamics M I T Press Cambridge MA pp 47 57 a b c Munster A 1970 Classical Thermodynamics translated by E S Halberstadt Wiley Interscience London ISBN 0 471 62430 6 pp 49 69 a b Bailyn M 1994 A Survey of Thermodynamics American Institute of Physics Press New York ISBN 0 88318 797 3 pp 14 15 214 a b Callen H B 1960 1985 Thermodynamics and an Introduction to Thermostatistics first edition 1960 second edition 1985 John Wiley amp Sons New York ISBN 0 471 86256 8 pp 146 148 Kondepudi D Prigogine I 1998 Modern Thermodynamics From Heat Engines to Dissipative Structures John Wiley Chichester ISBN 0 471 97394 7 pp 115 116 Tisza L 1966 Generalized Thermodynamics M I T Press Cambridge MA p 58 Milne E A 1929 The effect of collisions on monochromatic radiative equilibrium Monthly Notices of the Royal Astronomical Society 88 493 502 Gyarmati I 1970 Non equilibrium Thermodynamics Field Theory and Variational Principles translated by E Gyarmati and W F Heinz Springer Berlin pp 63 66 Glansdorff P Prigogine I 1971 Thermodynamic Theory of Structure Stability and Fluctuations Wiley London ISBN 0 471 30280 5 pp 14 16 Bailyn M 1994 A Survey of Thermodynamics American Institute of Physics Press New York ISBN 0 88318 797 3 pp 133 135 Callen H B 1960 1985 Thermodynamics and an Introduction to Thermostatistics first edition 1960 second edition 1985 John Wiley amp Sons New York ISBN 0 471 86256 8 pp 309 310 Bryan G H 1907 Thermodynamics An Introductory Treatise dealing mainly with First Principles and their Direct Applications B G Teubner Leipzig p 3 Thermodynamics by George Hartley Bryan Archived from the original on 2011 11 18 Retrieved 2011 10 02 Pippard A B 1957 1966 p 18 Adkins C J 1968 1983 p 20 Bryan G H 1907 Thermodynamics An Introductory Treatise dealing mainly with First Principles and their Direct Applications B G Teubner Leipzig p 5 when a body is spoken of as growing hotter or colder an increase of temperature is always implied for the hotness and coldness of a body are qualitative terms which can only refer to temperature Thermodynamics by George Hartley Bryan Archived from the original on 2011 11 18 Retrieved 2011 10 02 a b c d Mach E 1900 Die Principien der Warmelehre Historisch kritisch entwickelt Johann Ambrosius Barth Leipzig section 22 pp 56 57 a b Serrin J 1986 Chapter 1 An Outline of Thermodynamical Structure pp 3 32 especially p 6 in New Perspectives in Thermodynamics edited by J Serrin Springer Berlin ISBN 3 540 15931 2 Maxwell J C 1872 Theory of Heat third edition Longmans Green London p 32 Tait P G 1884 Heat Macmillan London Chapter VII pp 39 40 Planck M 1897 1903 Treatise on Thermodynamics translated by A Ogg Longmans Green London pp 1 2 Planck M 1914 The Theory of Heat Radiation Archived 2011 11 18 at the Wayback Machine second edition translated into English by M Masius Blakiston s Son amp Co Philadelphia reprinted by Kessinger J S Dugdale 1996 Entropy and its Physical Interpretation Taylor amp Francis p 13 ISBN 978 0 7484 0569 5 F Reif 1965 Fundamentals of Statistical and Thermal Physics McGraw Hill p 102 ISBN 9780070518001 M J Moran H N Shapiro 2006 1 6 1 Fundamentals of Engineering Thermodynamics 5 ed John Wiley amp Sons Ltd p 14 ISBN 978 0 470 03037 0 a href Template Cite book html title Template Cite book cite book a CS1 maint multiple names authors list link T W Leland Jr Basic Principles of Classical and Statistical Thermodynamics PDF p 14 Archived PDF from the original on 2011 09 28 Consequently we identify temperature as a driving force which causes something called heat to be transferred Tait P G 1884 Heat Macmillan London Chapter VII pp 42 103 117 Beattie J A Oppenheim I 1979 Principles of Thermodynamics Elsevier Scientific Publishing Company Amsterdam ISBN 978 0 444 41806 7 p 29 Landsberg P T 1961 Thermodynamics with Quantum Statistical Illustrations Interscience Publishers New York p 17 Thomsen J S 1962 A restatement of the zeroth law of thermodynamics Am J Phys 30 4 294 296 Bibcode 1962AmJPh 30 294T doi 10 1119 1 1941991 Maxwell J C 1872 Theory of Heat third edition Longman s Green amp Co London p 45 a b Pitteri M 1984 On the axiomatic foundations of temperature Appendix G6 on pp 522 544 of Rational Thermodynamics C Truesdell second edition Springer New York ISBN 0 387 90874 9 Truesdell C Bharatha S 1977 The Concepts and Logic of Classical Thermodynamics as a Theory of Heat Engines Rigorously Constructed upon the Foundation Laid by S Carnot and F Reech Springer New York ISBN 0 387 07971 8 p 20 a b Serrin J 1978 The concepts of thermodynamics in Contemporary Developments in Continuum Mechanics and Partial Differential Equations Proceedings of the International Symposium on Continuum Mechanics and Partial Differential Equations Rio de Janeiro August 1977 edited by G M de La Penha L A J Medeiros North Holland Amsterdam ISBN 0 444 85166 6 pp 411 451 Maxwell J C 1872 Theory of Heat third edition Longmans Green London pp 155 158 Tait P G 1884 Heat Macmillan London Chapter VII Section 95 pp 68 69 Buchdahl H A 1966 p 73 Kondepudi D 2008 Introduction to Modern Thermodynamics Wiley Chichester ISBN 978 0 470 01598 8 Section 32 pp 106 108 Green Don Perry Robert H 2008 Perry s Chemical Engineers Handbook Eighth Edition 8th ed McGraw Hill Education p 660 ISBN 978 0071422949 The kelvin in the SI Brochure Archived 2007 09 26 at the Wayback Machine Absolute Zero Calphad com Archived from the original on 2011 07 08 Retrieved 2010 09 16 Definition agreed by the 26th General Conference on Weights and Measures CGPM Archived 2020 10 09 at the Wayback Machine in November 2018 implemented 20 May 2019 C Caratheodory 1909 Untersuchungen uber die Grundlagen der Thermodynamik Mathematische Annalen 67 3 355 386 doi 10 1007 BF01450409 S2CID 118230148 Swendsen Robert March 2006 Statistical mechanics of colloids and Boltzmann s definition of entropy PDF American Journal of Physics 74 3 187 190 Bibcode 2006AmJPh 74 187S doi 10 1119 1 2174962 S2CID 59471273 Archived from the original PDF on 2020 02 28 Balescu R 1975 Equilibrium and Nonequilibrium Statistical Mechanics Wiley New York ISBN 0 471 04600 0 pp 148 154 a b Kittel Charles Kroemer Herbert 1980 Thermal Physics 2nd ed W H Freeman Company pp 391 397 ISBN 978 0 7167 1088 2 Kondepudi D K 1987 Microscopic aspects implied by the second law Foundations of Physics 17 7 713 722 Bibcode 1987FoPh 17 713K doi 10 1007 BF01889544 S2CID 120576357 The Feynman Lectures on Physics 39 5 The ideal gas law 2018 CODATA Value molar gas constant The NIST Reference on Constants Units and Uncertainty NIST 20 May 2019 Retrieved 2019 05 20 Kinetic Theory galileo phys virginia edu Archived from the original on 16 July 2017 Retrieved 27 January 2018 Tolman R C 1938 The Principles of Statistical Mechanics Oxford University Press London pp 93 655 Peter Atkins Julio de Paula 2006 Physical Chemistry 8 ed Oxford University Press p 9 Maxwell J C 1872 Theory of Heat third edition Longman s Green amp Co London p 32 Bailyn M 1994 A Survey of Thermodynamics American Institute of Physics Press New York ISBN 0 88318 797 3 p 23 if a temperature gradient exists then a flow of heat must occur to achieve a uniform temperature Guggenheim E A 1967 Thermodynamics An Advanced Treatment for Chemists and Physicists North Holland Publishing Company Amsterdam 1st edition 1949 fifth edition 1965 p 8 will gradually adjust themselves until eventually they do reach mutual equilibrium after which there will of course be no further change Bailyn M 1994 A Survey of Thermodynamics American Institute of Physics Press New York ISBN 0 88318 797 3 p 22 Guggenheim E A 1967 Thermodynamics An Advanced Treatment for Chemists and Physicists North Holland Publishing Company Amsterdam 1st edition 1949 fifth edition 1965 p 8 If two systems are both in thermal equilibrium with a third system then they are in thermal equilibrium with each other Buchdahl H A 1966 The Concepts of Classical Thermodynamics Cambridge University Press Cambridge p 29 if each of two systems is in equilibrium with a third system then they are in equilibrium with each other a b Planck M 1945 Treatise on Thermodynamics Dover Publications p 90 amp 137 eqs 39 40 amp 65 Prati E 2010 The finite quantum grand canonical ensemble and temperature from single electron statistics for a mesoscopic device J Stat Mech 1 1 P01003 arXiv 1001 2342 Bibcode 2010JSMTE 01 003P doi 10 1088 1742 5468 2010 01 P01003 S2CID 118339343 arxiv org Archived 2017 11 22 at the Wayback Machine Archived copy PDF Archived PDF from the original on 2014 04 13 Retrieved 2014 04 11 a href Template Cite web html title Template Cite web cite web a CS1 maint archived copy as title link Prati E et al 2010 Measuring the temperature of a mesoscopic electron system by means of single electron statistics Applied Physics Letters 96 11 113109 arXiv 1002 0037 Bibcode 2010ApPhL 96k3109P doi 10 1063 1 3365204 S2CID 119209143 Archived from the original on 2016 05 14 Retrieved 2022 03 02 arxiv org Archived 2017 11 22 at the Wayback Machine Water Science School Frozen carbon dioxide dry ice sublimates directly into a vapor USGS Guggenheim E A 1967 1949 Thermodynamics An Advanced Treatment for Chemists and Physicists fifth ed Amsterdam North Holland Publishing Company p 157 It is impossible by any procedure no matter how idealized to reduce the temperature of any system to zero temperature in a finite number of finite operations Pippard A B 1957 1966 Elements of Classical Thermodynamics for Advanced Students of Physics original publication 1957 reprint 1966 Cambridge University Press Cambridge page 51 By no finite series of processes is the absolute zero attainable Tisza L 1966 Generalized Thermodynamics MIT Press Cambridge MA page 96 It is impossible to reach absolute zero as a result of a finite sequence of operations Kittel Charles Kroemer Herbert 1980 Thermal Physics 2nd ed W H Freeman Company p Appendix E ISBN 978 0 7167 1088 2 The cited emission wavelengths are for black bodies in equilibrium CODATA 2006 recommended value of 2 8977685 51 10 3 m K used for Wien displacement law constant b This the Hawking Radiation for a Schwarzschild black hole of mass M 4 145 106 M It is too faint to be observed World record in low temperatures Archived from the original on 2009 06 18 Retrieved 2009 05 05 A temperature of 450 80 pK in a Bose Einstein condensate BEC of sodium atoms was achieved in 2003 by researchers at MIT Citation Cooling Bose Einstein Condensates Below 500 Picokelvin A E Leanhardt et al Science 301 12 Sept 2003 p 1515 It s noteworthy that this record s peak emittance black body wavelength of 6 400 kilometers is roughly the radius of Earth The peak emittance wavelength of 2 89777 m is a frequency of 103 456 MHz Measurement was made in 2002 and has an uncertainty of 3 kelvins A 1989 measurement Archived 2010 02 11 at the Wayback Machine produced a value of 5 777 0 2 5 K Citation Overview of the Sun Chapter 1 lecture notes on Solar Physics by Division of Theoretical Physics Dept of Physical Sciences University of Helsinki The 350 MK value is the maximum peak fusion fuel temperature in a thermonuclear weapon of the Teller Ulam configuration commonly known as a hydrogen bomb Peak temperatures in Gadget style fission bomb cores commonly known as an atomic bomb are in the range of 50 to 100 MK Citation Nuclear Weapons Frequently Asked Questions 3 2 5 Matter At High Temperatures Link to relevant Web page Archived 2007 05 03 at the Wayback Machine All referenced data was compiled from publicly available sources Peak temperature for a bulk quantity of matter was achieved by a pulsed power machine used in fusion physics experiments The term bulk quantity draws a distinction from collisions in particle accelerators wherein high temperature applies only to the debris from two subatomic particles or nuclei at any given instant The gt 2 GK temperature was achieved over a period of about ten nanoseconds during shot Z1137 In fact the iron and manganese ions in the plasma averaged 3 58 0 41 GK 309 35 keV for 3 ns ns 112 through 115 Ion Viscous Heating in a Magnetohydrodynamically Unstable Z Pinch at Over 2 109 Kelvin M G Haines et al Physical Review Letters 96 2006 075003 Link to Sandia s news release Archived 2010 05 30 at the Wayback Machine Core temperature of a high mass gt 8 11 solar masses star after it leaves the main sequence on the Hertzsprung Russell diagram and begins the alpha process which lasts one day of fusing silicon 28 into heavier elements in the following steps sulfur 32 argon 36 calcium 40 titanium 44 chromium 48 iron 52 nickel 56 Within minutes of finishing the sequence the star explodes as a Type II supernova Citation Stellar Evolution The Life and Death of Our Luminous Neighbors by Arthur Holland and Mark Williams of the University of Michigan Link to Web site Archived 2009 01 16 at the Wayback Machine More informative links can be found here Chapter 21 Stellar Explosions Archived from the original on 2013 04 11 Retrieved 2016 02 08 and here Trans Archived from the original on 2011 08 14 Retrieved 2016 02 08 and a concise treatise on stars by NASA is here NASA Star Archived from the original on 2010 10 24 Retrieved 2010 10 12 Stellar Archived from the original on January 16 2009 Retrieved 2010 10 12 a href Template Cite web html title Template Cite web cite web a CS1 maint bot original URL status unknown link Based on a computer model that predicted a peak internal temperature of 30 MeV 350 GK during the merger of a binary neutron star system which produces a gamma ray burst The neutron stars in the model were 1 2 and 1 6 solar masses respectively were roughly 20 km in diameter and were orbiting around their barycenter common center of mass at about 390 Hz during the last several milliseconds before they completely merged The 350 GK portion was a small volume located at the pair s developing common core and varied from roughly 1 to 7 km across over a time span of around 5 ms Imagine two city sized objects of unimaginable density orbiting each other at the same frequency as the G4 musical note the 28th white key on a piano It s also noteworthy that at 350 GK the average neutron has a vibrational speed of 30 the speed of light and a relativistic mass m 5 greater than its rest mass m0 Torus Formation in Neutron Star Mergers and Well Localized Short Gamma Ray Bursts Archived 2017 11 22 at the Wayback Machine R Oechslin et al of Max Planck Institute for Astrophysics Archived 2005 04 03 at the Wayback Machine arXiv astro ph 0507099 v2 22 Feb 2006 An html summary Archived 2010 11 09 at the Wayback Machine Results of research by Stefan Bathe using the PHENIX Archived 2008 11 20 at the Wayback Machine detector on the Relativistic Heavy Ion Collider Archived 2016 03 03 at the Wayback Machine at Brookhaven National Laboratory Archived 2012 06 24 at the Wayback Machine in Upton New York Bathe has studied gold gold deuteron gold and proton proton collisions to test the theory of quantum chromodynamics the theory of the strong force that holds atomic nuclei together Link to news release Archived 2009 02 11 at the Wayback Machine How do physicists study particles Archived 2007 10 11 at the Wayback Machine by CERN Archived 2012 07 07 at the Wayback Machine Bibliography of cited references Edit Adkins C J 1968 1983 Equilibrium Thermodynamics 1st edition 1968 third edition 1983 Cambridge University Press Cambridge UK ISBN 0 521 25445 0 Buchdahl H A 1966 The Concepts of Classical Thermodynamics Cambridge University Press Cambridge Jaynes E T 1965 Gibbs vs Boltzmann entropies American Journal of Physics 33 5 391 398 Middleton W E K 1966 A History of the Thermometer and its Use in Metrology Johns Hopkins Press Baltimore Miller J 2013 Cooling molecules the optoelectric way Physics Today 66 1 12 14 Bibcode 2013PhT 66a 12M doi 10 1063 pt 3 1840 Archived from the original on 2016 05 15 Retrieved 2013 07 25 Partington J R 1949 An Advanced Treatise on Physical Chemistry volume 1 Fundamental Principles The Properties of Gases Longmans Green amp Co London pp 175 177 Pippard A B 1957 1966 Elements of Classical Thermodynamics for Advanced Students of Physics original publication 1957 reprint 1966 Cambridge University Press Cambridge UK Quinn T J 1983 Temperature Academic Press London ISBN 0 12 569680 9 Schooley J F 1986 Thermometry CRC Press Boca Raton ISBN 0 8493 5833 7 Roberts J K Miller A R 1928 1960 Heat and Thermodynamics first edition 1928 fifth edition Blackie amp Son Limited Glasgow Thomson W Lord Kelvin 1848 On an absolute thermometric scale founded on Carnot s theory of the motive power of heat and calculated from Regnault s observations Proc Camb Phil Soc 1843 1863 1 No 5 66 71 Thomson W Lord Kelvin March 1851 On the Dynamical Theory of Heat with numerical results deduced from Mr Joule s equivalent of a Thermal Unit and M Regnault s Observations on Steam Transactions of the Royal Society of Edinburgh XX part II 261 268 289 298 Truesdell C A 1980 The Tragicomical History of Thermodynamics 1822 1854 Springer New York ISBN 0 387 90403 4 Tschoegl N W 2000 Fundamentals of Equilibrium and Steady State Thermodynamics Elsevier Amsterdam ISBN 0 444 50426 5 Zeppenfeld M Englert B G U Glockner R Prehn A Mielenz M Sommer C van Buuren L D Motsch M Rempe G 2012 Sysiphus cooling of electrically trapped polyatomic molecules Nature 491 7425 570 573 arXiv 1208 0046 Bibcode 2012Natur 491 570Z doi 10 1038 nature11595 PMID 23151480 S2CID 4367940 Further reading EditChang Hasok 2004 Inventing Temperature Measurement and Scientific Progress Oxford Oxford University Press ISBN 978 0 19 517127 3 Zemansky Mark Waldo 1964 Temperatures Very Low and Very High Princeton NJ Van Nostrand External links EditTemperature at Wikipedia s sister projects Definitions from Wiktionary Media from Commons Resources from Wikiversity Data from Wikidata Current map of global surface temperatures Portal Energy Retrieved from https en wikipedia org w index php title Temperature amp oldid 1129420628, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.