Fact-checked by Grok 2 weeks ago

Boltzmann constant

The Boltzmann constant, denoted as k or kB, is a that relates the average of particles in a gas to the of the gas, serving as a bridge between microscopic statistical descriptions and macroscopic thermodynamic properties. In the (SI), it is defined exactly as 1.380649 × 10−23 joules per (J/K). Named after the Austrian physicist (1844–1906), the constant emerged from his pioneering work in during the late 19th century, where it quantified the connection between and the number of microscopic configurations of a system. Boltzmann developed the relation between and probability in his 1877 work, expressing it as S \propto \ln W, where W is the multiplicity of microstates. The constant k was introduced by in 1900, who named it in Boltzmann's honor, providing a probabilistic foundation for the second law of thermodynamics. The constant's significance extends across physics: in the (PV = N k T, where N is the number of particles and T is ), it links , , and at the molecular level; in , it scales the energy distribution in systems like the Maxwell-Boltzmann distribution; and in quantum statistics, it appears in Fermi-Dirac and Bose-Einstein formulations. Its exact value was fixed in the 2019 SI redefinition, which redefined the in terms of k rather than the of , enabling more precise and universal measurements independent of material artifacts. This redefinition, based on advanced techniques like Johnson noise thermometry, underscores k's role as an invariant cornerstone of modern .

Definition and Value

Physical Significance

The Boltzmann constant, denoted as k or k_B, is the fundamental proportionality factor that relates the average of particles in a to the , the conversion between scales (such as [kT](/page/KT)) and mechanical or units. This scaling allows physicists to quantify the energy associated with in microscopic processes, where T is the in kelvins. As a universal bridge between the macroscopic laws of phenomenological and the microscopic probabilities of , the Boltzmann constant links observable bulk properties—like and in gases—to the random motions and configurations of individual particles in . Its applicability spans diverse systems, from ideal gases to complex materials, emphasizing the shared statistical foundations of thermal phenomena across physics. The dimensions of the Boltzmann constant are those of energy per unit temperature, expressed as joules per (J/) or, in base SI units, ··s⁻²·⁻¹. The symbol k_B, with the subscript B, distinguishes it from other constants denoted by k, such as the Coulomb constant k_e = 1/(4\pi\epsilon_0). Named after Austrian physicist (1844–1906) for his foundational work in , the constant underscores the probabilistic interpretation of thermodynamic quantities.

Numerical Value

The Boltzmann constant k is defined exactly as k = 1.380649 \times 10^{-23} J/K following the 2019 revision of the (SI), which fixed its numerical value to define the kelvin independently of experimental measurements. This exactness eliminates any uncertainty in k itself, thereby shifting metrological efforts toward refining measurements of other defining constants, such as the h. Prior to the 2019 redefinition, the Committee on Data for Science and Technology (CODATA) recommended a value of k = 1.38064852(79) \times 10^{-23} J/K based on the 2014 adjustment, with a relative standard uncertainty of $5.7 \times 10^{-7}. The exact value of k facilitates conversions between and scales in various units, essential for interdisciplinary applications. The table below lists selected common conversions derived from the fixed value.
UnitValue of k
J/K (SI)$1.380649 \times 10^{-23}
/K$8.617333262145 \times 10^{-5}
/K$3.299158069393 \times 10^{-24}
These conversions, computed using exact relations between SI base units, support precise thermodynamic modeling. In , where thermal energies approach the millikelvin regime, and in high-temperature physics involving plasmas or astrophysical environments, the exact k ensures uncertainty-free computation of thermal energy scales like [kT](/page/KT), for instance, yielding approximately 25 meV at (300 K). This precision enhances the reliability of simulations and experimental validations in extreme conditions.

Roles in Statistical Mechanics

Equipartition of Energy

The states that, in a classical at , each quadratic degree of freedom contributes an average energy of \frac{1}{2} k [T](/page/Temperature) to the total energy, where k is the Boltzmann constant and T is the absolute . This theorem provides a fundamental link between microscopic energy distribution and macroscopic in . For a system possessing f independent quadratic degrees of freedom, the total average energy per particle is given by \langle E \rangle = \frac{f}{2} k T. For instance, the translational motion of a molecule has three quadratic (one for each Cartesian component), yielding \langle E \rangle = \frac{3}{2} k T for the average translational . The theorem derives from Maxwell-Boltzmann statistics, where the average value of an term is obtained by integrating over with the Boltzmann weight e^{-E / k T}. Consider a contribution \epsilon(p_i) = b p_i^2 depending on a single coordinate p_i, separable from the rest of the system's E'. The average \overline{\epsilon_i} is then \overline{\epsilon_i} = \frac{\int_{-\infty}^{\infty} \epsilon(p_i) \, e^{-\beta \epsilon(p_i)} \, dp_i}{\int_{-\infty}^{\infty} e^{-\beta \epsilon(p_i)} \, dp_i}, with \beta = 1 / k T. Evaluating the Gaussian integrals yields \overline{\epsilon_i} = \frac{1}{2} k T, establishing the \frac{1}{2} k T contribution per term. This result extends to position-dependent terms by analogy. In the kinetic theory of ideal gases, the three translational degrees of freedom lead to an average kinetic energy of \frac{3}{2} k T per molecule, independent of molecular mass or interactions in the dilute limit. For a classical harmonic oscillator, the energy includes both kinetic (\frac{1}{2} m v^2) and potential (\frac{1}{2} k x^2) quadratic terms, resulting in a total average energy of k T per mode. The applies strictly in the classical regime, where the k T greatly exceeds the quantum energy level spacing, allowing continuous sampling. It fails at quantum scales, particularly at low temperatures, where modes like vibrational oscillators in solids retain and do not achieve full classical excitation, leading to deficits.

Boltzmann Factors

In statistical mechanics, the Boltzmann factor quantifies the relative likelihood of a system occupying different energy states in within the . For two states with energies E_i and E_j, the ratio of their probabilities is given by \frac{P_i}{P_j} = \exp\left( -\frac{E_i - E_j}{kT} \right), where k is the Boltzmann constant and T is the absolute temperature. This exponential form arises because higher-energy states are less probable, with the factor kT determining the characteristic energy scale set by . For a with levels, the full incorporates degeneracy, the number of microstates g(E) associated with E. The probability P(E) of finding the at E is then P(E) \propto g(E) \exp\left( -\frac{E}{kT} \right), normalized such that the over all states . This distribution emerges from of maximum , where the most probable configuration maximizes the Shannon entropy subject to constraints on average and particle number, yielding the weighting as the unique solution. Alternatively, it can be derived by projecting the of the —valid for an of fixed —onto a smaller subsystem in contact with a large ; the probability then becomes proportional to the 's at the complementary , approximating the form in the . The Boltzmann factor finds direct application in determining the population ratios of excited states in atoms and molecules at . For instance, in a two-level system like the ground and first excited electronic states of an atom, the ratio of populations N_{\rm excited}/N_{\rm ground} = (g_{\rm excited}/g_{\rm ground}) \exp(-\Delta E / kT), where \Delta E is the difference; at , states separated by several kT are sparsely populated, explaining the dominance of ground states in dilute gases. Similarly, in molecular , vibrational or rotational excited states follow this distribution, enabling temperature measurements from intensities. In the context of ideal gases, the Boltzmann factor underpins the Maxwell-Boltzmann speed distribution. For non-interacting particles, the probability density f(v) for speeds v is obtained by considering the phase space volume and applying the factor to the kinetic energy E = \frac{1}{2} m v^2, yielding an integral form f(v) \, dv \propto 4\pi v^2 \exp\left( -\frac{m v^2}{2 kT} \right) \, dv, which describes the distribution of molecular speeds without requiring explicit solution of the normalization. This probabilistic weighting highlights how the Boltzmann constant establishes the thermal energy scale: at T = 300 K, kT \approx 4.14 \times 10^{-21} J, comparable to molecular binding energies but much smaller than electronic transitions, thus dictating the extent of thermal excitation. The averages derived from this distribution, such as mean kinetic energies, connect to broader principles like equipartition in quadratic systems.

Statistical Definition of Entropy

In , the Boltzmann constant serves as the proportionality factor that connects the microscopic multiplicity of states to the macroscopic , quantifying the degree of disorder or uncertainty in a . The seminal expression for this is the Boltzmann entropy formula,
S = k \ln W,
where S denotes the , k is the Boltzmann constant, and W is the number of accessible microstates corresponding to a given macrostate. This formulation establishes as a logarithmic measure of probabilistic possibilities, with k ensuring the result has dimensions of energy per temperature (joules per kelvin).
A broader derivation emerges from the in equilibrium , yielding the Gibbs formula
S = -[k](/page/K) \sum_i p_i \ln p_i,
where p_i is the probability of of i. When the system has W equally likely microstates, each with p_i = 1/W, the sum simplifies to S = [k](/page/K) \ln W, recovering the original form.
This probabilistic expression closely resembles the Shannon entropy from ,
H = -\sum_i p_i \log_2 p_i,
which measures uncertainty in bits; the statistical mechanical version employs the natural logarithm and scales by k to confer thermodynamic units, thereby linking abstract information content to physical energy scales.
For a monatomic , the Sackur-Tetrode equation illustrates k's scaling role:
S \approx Nk \left[ \ln \left( \frac{V}{N} \right) + \frac{3}{2} \ln T + c \right],
where N is the particle number, V the volume, T the , and c a constant incorporating mass and quantum effects; here, k multiplies the logarithmic multiplicity terms to produce an extensive proportional to system size.
In a two-state paramagnet with N spins, each able to align up or down in a magnetic field, the multiplicity for a configuration with N_+ up-spins is W = \binom{N}{N_+}, so S = k \ln W, maximizing at S = Nk \ln 2 for equal populations and underscoring k's conversion of combinatorial growth to thermal disorder. The presence of k guarantees entropy additivity for composite systems of independent subsystems, as W_\text{total} = W_1 W_2 implies S_\text{total} = S_1 + S_2, aligning statistical predictions with the extensive nature of thermodynamic entropy. It also resolves the Gibbs paradox in gas mixing, where treating particles as indistinguishable avoids unphysical entropy jumps, with k preserving the correct scaling for identical versus distinct components.

Applications in Physics and Engineering

Ideal Gas Law

The Boltzmann constant k plays a central role in the microscopic interpretation of the , bridging the kinetic behavior of individual particles to macroscopic thermodynamic properties. In kinetic theory, the P exerted by an arises from the momentum flux of particles colliding with the container walls. Considering a gas of N particles, each with m, the pressure is derived as P = \frac{1}{3} \frac{N m \langle v^2 \rangle}{V}, where V is the volume and \langle v^2 \rangle is the mean square speed. The assigns an average translational of \frac{3}{2} kT per particle, leading to \frac{1}{2} m \langle v^2 \rangle = \frac{3}{2} kT. Substituting this relation yields the microscopic form of the : PV = NkT. This microscopic equation connects directly to the empirical molar form PV = nRT, where n is the number of moles and R is the molar gas constant. Here, N = n N_A, with N_A being Avogadro's number, so R = N_A k. Historically, the value of k emerged from dividing the measured R by N_A, providing a per-particle energy scale for temperature. Following the 2019 redefinition of the SI units, k was fixed exactly at $1.380649 \times 10^{-23} J/K, rendering both N_A and R exact as well. The ideal gas law with k underpins classical gas behaviors, such as (PV = constant at fixed T), which follows from P \propto 1/V at constant NkT, and (V \propto T at fixed P), reflecting tied to increasing kinetic energies \propto kT. For real gases, the (P + \frac{a n^2}{V^2})(V - n b) = n R T introduces corrections for intermolecular forces (a) and molecular volume (b), but reduces to the ideal form PV = n R T (or PV = N k T) in the limit of low density where these effects vanish. Experimental verifications of this framework link macroscopic observables back to kT energy scales. For instance, the in an ideal is v = \sqrt{\frac{\gamma k T}{m}}, where \gamma = \frac{5}{3} is the adiabatic index, directly incorporating the thermal per particle. Similarly, the specific heat at volume for a , C_V = \frac{3}{2} R = \frac{3}{2} N_A k, confirms the three translational each contributing \frac{1}{2} kT per particle.

Thermal Voltage

In semiconductor physics, the thermal voltage V_T is a key parameter defined as V_T = \frac{k T}{q}, where k is the , T is the absolute temperature in , and q is the , exactly 1.602176634 × 10^{-19} C (since the 2019 SI redefinition). This quantity represents the thermal energy scale in electron volts, bridging temperature to electrical potential across junctions. At (300 ), V_T evaluates to approximately 25.85 mV, providing a characteristic voltage for dynamics in devices. The thermal voltage appears prominently in the Shockley diode equation, which describes the current through a p-n junction:
I = I_s \left( \exp\left( \frac{V}{V_T} \right) - 1 \right),
where I is the , I_s is the , and V is the applied forward bias voltage. This exponential form arises from the Boltzmann factors that dictate the concentration of minority carriers injected across the junction, with V_T setting the steepness of the current rise. In bipolar junction transistors (BJTs), V_T similarly governs the base-emitter junction, where the collector follows I_C = I_S \exp\left( \frac{V_{BE}}{V_T} \right), causing the base-emitter voltage to scale logarithmically with and linearly with . These relations enable precise modeling of p-n junctions in and transistors, essential for and switching in electronic circuits.
The temperature dependence of V_T, which increases proportionally with T, significantly impacts device performance; for instance, higher temperatures reduce the forward voltage drop across a diode at fixed current by about -2 mV/K due to the enhanced thermal generation of carriers. This effect is measured via the forward voltage drop technique, where the voltage across a biased diode or transistor junction is monitored as a proxy for temperature, often using integrated sensors in integrated circuits. In electronics, V_T also underlies thermal noise, known as Johnson-Nyquist noise, where the root-mean-square noise voltage across a resistor R in bandwidth \Delta f scales as \sqrt{4 k T R \Delta f}, limiting signal integrity in low-noise amplifiers and sensors. In degenerate semiconductors, where the Fermi level lies within the conduction or valence band, the thermal voltage V_T (or equivalently [kT](/page/KT)) establishes the energy scale relative to the E_F; when E_F \gg kT, Fermi-Dirac statistics replace classical Boltzmann approximations, altering carrier concentrations and transport properties in heavily doped materials used in high-speed devices.

Historical Development

Origins and Naming

The development of the Boltzmann constant emerged in the mid-19th century amid the rise of , which sought to explain macroscopic thermodynamic properties through the microscopic behavior of . In 1860, James Clerk Maxwell published his seminal paper "Illustrations of the dynamical theory of gases," deriving the distribution of molecular velocities in an based on assumptions of random collisions and elastic interactions. This distribution implicitly incorporated a proportionality constant relating the average per molecule to , equivalent to what would later be identified as the Boltzmann constant k, appearing as k = R / N where R is the and N is the number of molecules (or Avogadro's number N_A)./27%3A_The_Kinetic_Theory_of_Gases/27.03%3A_The_Distribution_of_Molecular_Speeds_is_Given_by_the_Maxwell-Boltzmann_Distribution) Maxwell's work laid the groundwork for statistical interpretations of , though the constant itself remained unnamed and not explicitly isolated at the time. Ludwig Boltzmann advanced this framework significantly in the 1870s through his investigations into the statistical mechanics of gases. In his 1866 paper "Über die mechanische Bedeutung des zweiten Hauptsatzes der Wärmetheorie," Boltzmann explored the equilibrium distribution of molecular energies, building on Maxwell's velocity distribution and introducing probabilistic considerations for energy sharing among particles, which foreshadowed the equipartition theorem. His most influential contribution came in 1877 with the paper "Über die Beziehung zwischen dem zweiten Hauptsatz der mechanischen Wärmetheorie und der Wahrscheinlichkeitsrechnung," where he formulated the entropy of a system in terms of its microscopic states as S = k \ln W, with W representing the number of possible microstates and k the proportionality constant linking thermodynamic entropy to probability. This expression, introduced in this work—which builds on Boltzmann's earlier H-theorem (1872) by providing a probabilistic foundation for the approach to equilibrium—explicitly introduced k as a fundamental bridge between macroscopic and microscopic scales, resolving paradoxes in classical thermodynamics like the ultraviolet catastrophe through statistical averaging. Early experimental validation of the constant's role came from studies connecting and specific heats to molecular scales. By the early , Perrin's 1908 experiments on provided indirect confirmation of k by determining Avogadro's number N_A through observations of colloidal particle displacements, yielding k = R / N_A consistent with theoretical predictions from kinetic theory. These results, detailed in Perrin's work "Mouvement brownien et réalité moléculaire," supported the atomic hypothesis and quantified k's value via sedimentation equilibrium and measurements. The constant was formally named the "Boltzmann constant" in the early in recognition of Ludwig Boltzmann's pioneering statistical formulations, with noting its common attribution to Boltzmann by in his Nobel lecture. To avoid ambiguity with other physical constants like the wave number k (in k = 2\pi / \lambda) or the Coulomb constant k_e = 1/(4\pi\epsilon_0), it is conventionally denoted as k_B or k_B.

Measurement and 2019 Redefinition

The experimental determination of the Boltzmann constant (k) has evolved significantly over the past century, driven by advances in precision metrology to support its role in linking thermodynamic temperature to microscopic energy scales. In the 1920s, initial estimates were derived from X-ray scattering and diffraction experiments, which enabled measurements of the Avogadro constant (N_A) through crystal lattice spacings and densities; combined with the known molar gas constant (R), these yielded early values of k = R / N_A with relative uncertainties on the order of 0.1% or larger. By the 1970s, CODATA adjustments refined k to $1.380 \times 10^{-23} J/K with a relative uncertainty of about $8 \times 10^{-5}, incorporating data from speed-of-sound measurements and electrochemical cells. Further progress in the 2010s achieved precisions to parts per billion, facilitated by linkages to other fundamental constants via watt balance experiments (for the Planck constant h) and silicon sphere volumetry (for N_A). The 2014 CODATA recommended value was k = 1.38064852(79) \times 10^{-23} J/K, with a relative standard uncertainty of $5.7 \times 10^{-7}, reflecting contributions from multiple independent methods. Key experimental approaches included acoustic gas thermometry, which measures the speed of sound in gases like argon at known pressures and volumes to relate macroscopic thermodynamic properties to k; Johnson noise thermometry, which quantifies thermal voltage fluctuations across a resistor proportional to kT (where T is temperature); and Doppler broadening spectroscopy, analyzing the thermal broadening of spectral lines in gases such as helium to infer k from line widths. These methods provided consistent results with uncertainties below 1 part per million, essential for the impending SI revision. The 2019 revision of the (SI), effective May 20, 2019, fixed k exactly at $1.380649 \times 10^{-23} J/K, alongside the (h), (e), and (N_A). This redefinition anchors the to a fundamental constant, such that the kelvin, symbol K, is the SI unit of defined by taking the fixed numerical value of k to be $1.380649 \times 10^{-23} when expressed in the unit J/K = kg m^2 s^{-2} K^{-1}, where the , meter, and second are defined in terms of h, the c, and the hyperfine transition frequency \Delta \nu_{\rm Cs}. The of , previously the basis for the , now serves as a secondary reference point near 273.16 K, ensuring continuity with prior scales. Challenges in these measurements include systematic errors from thermophysical properties (e.g., gas impurities or virial coefficients in acoustic methods) and inter-method discrepancies, which required rigorous uncertainty budgeting to achieve pre-redefinition consensus. Post-redefinition, calibration standards like the International Temperature Scale of 1990 (ITS-90) must be traceable to k via primary thermometry, potentially introducing small non-uniqueness in fixed-point realizations (e.g., up to ±1.23 mK deviations in mercury triple points), necessitating updates to practical scales. In the 2020s, ongoing experiments focus on consistency checks across methods, including refined thermometry targeting 0.1 uncertainties below 25 K via effects, and high-temperature acoustic gas thermometry up to 3000 K using carbon eutectics for traceability above 1300 K. These efforts aim to validate the fixed value and support advanced applications in quantum thermometry and .

Units and Dimensionless Quantities

SI and Conventional Units

The Boltzmann constant k is exactly defined in the (SI) as k = 1.380649 \times 10^{-23} J/K, where the joule (J) is the unit of energy and the kelvin (K) is the unit of temperature. This value establishes the scale between and temperature in SI, with the equivalent expression in base SI units being k = 1.380649 \times 10^{-23} kg m² s⁻² K⁻¹. In conventional unit systems, the Boltzmann constant takes on corresponding values to maintain dimensional consistency, where its dimensions are always energy per unit temperature ([M] [L]² [T]⁻² [Θ]⁻¹, with [M] , [L] , [T] time, and [Θ] ). In the centimeter-gram-second (CGS) system, k = 1.380649 \times 10^{-16} erg/K, since 1 J = 10⁷ erg. A value frequently used in semiconductor physics and is k = 8.617333262145 \times 10^{-5} eV/K. The Boltzmann constant relates to the gas constant R via R = N_A k, where N_A is Avogadro's constant, exactly defined as N_A = 6.02214076 \times 10^{23} mol⁻¹ following the 2019 redefinition. This yields the exact value R = 8.314462618 J mol⁻¹ K⁻¹, which scales k to quantities in and . The following table summarizes the Boltzmann constant in and spectroscopic units, derived from its value and standard conversions for precision in and molecular :
UnitValueReference
hartree/K (E_h/K)$3.1668114 \times 10^{-6} E_h K⁻¹https://physics.nist.gov/cgi-bin/cuu/Value?k, https://physics.nist.gov/cgi-bin/cuu/Value?hrj
cm⁻¹/K$0.69503476 cm⁻¹ K⁻¹https://physics.nist.gov/cgi-bin/cuu/Value?kshcminv

Natural Units

In natural unit systems employed in theoretical physics, the Boltzmann constant k is conventionally set to 1, thereby endowing temperature T with the dimensions of energy, such as electronvolts (eV) or gigaelectronvolts (GeV). This choice eliminates the explicit appearance of k in equations, treating thermal energy kT equivalently to T. Several common natural unit systems incorporate this convention. In Planck units, k is unified with the reduced Planck constant \hbar, the speed of light c, and the gravitational constant G, all set to 1, yielding a temperature scale where T aligns with Planck energy units. Particle physics natural units typically set \hbar = c = 1, expressing thermal energies kT in units like MeV, supported by conversion factors such as \hbar c \approx 197 MeV·fm for length-energy relations. In atomic units, thermal energies kT are quantified in hartrees, the base energy unit derived from electron mass, charge, and \hbar. This framework offers significant advantages by streamlining key expressions in . For instance, the Boltzmann factor simplifies from \exp(-E / kT) to \exp(-E / T), reducing algebraic complexity in partition functions and distribution calculations. In , it facilitates direct energy-temperature comparisons, as seen with the temperature T = 2.725 K equivalent to approximately $2.35 \times 10^{-4} eV. In relativistic where c = 1, thermal energies kT are immediately comparable to particle rest energies mc^2, though explicit kT / c^2 formulations occasionally appear when restoring c for mass-energy equivalence checks. However, setting k = 1 diminishes the overt connection between theoretical energy scales and empirical temperature measurements, necessitating conversions to SI units for validation against laboratory data.

References

  1. [1]
    Kelvin: Boltzmann Constant | NIST
    May 15, 2018 · The Boltzmann constant (k B ) relates temperature to energy. It is an indispensable tool in thermodynamics, the study of heat and its relationship to other ...
  2. [2]
    Boltzmann's Work in Statistical Physics
    Nov 17, 2004 · Ludwig Boltzmann (1844–1906) is generally acknowledged as one of the most important physicists of the nineteenth century.
  3. [3]
    Boltzmann constant - CODATA Value
    Concise form, 1.380 649 x 10-23 J K ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  4. [4]
    - kelvin - BIPM
    It is defined by taking the fixed numerical value of the Boltzmann constant k to be 1.380 649 x 10–23 when expressed in the unit J K–1, which is equal to kg ...
  5. [5]
    The International System of Units (SI): Defining constants - BIPM
    The SI is defined in terms of a set of seven defining constants. The complete system of units can be derived from the fixed values of these defining constants.
  6. [6]
    [PDF] CODATA recommended values of the fundamental physical constants
    Sep 26, 2016 · Relative atomic mass of the electron Ar(e). The relative standard uncertainty of the 2014 recommended value of Ar(e) is 2.9 × 10−11. , nearly ...
  7. [7]
    Boltzmann constant in eV/K - CODATA Value
    Concise form, 8.617 333 262... x 10-5 eV K ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  8. [8]
    The equipartition theorem - Richard Fitzpatrick
    This is the famous equipartition theorem of classical physics. It states that the mean value of every independent quadratic term in the energy is equal to $(1/ ...
  9. [9]
    [PDF] The Equipartition Theorem - FSU High Energy Physics
    A result from classical statistical mechanics is the equipartition theorem: When a substance is in equilibrium, there is an average energy of kT/2 per molecule ...
  10. [10]
    [PDF] Lecture 4: Temperature
    The ideal gas constant is a mole of Boltzmann's constants: R =kBNA =8.3. J mol K . 4 Equipartition theorem. Recall that for a monatomic ideal gas the energy ...
  11. [11]
    [PDF] A Derivation of the Equipartition Theorem - Physics Courses
    The general derivation of the equipartition theorem involves statistical mechanics beyond the scope of our discussions, so we will do a special case using a ...
  12. [12]
    Lecture 11: Equipartition of energy - ESM Intranet Site
    Limitations of the equipartition theorem​​ The equipartition theorem is based on the fact that an integral such as the one shown above can be performed with a ...
  13. [13]
    [PDF] Failure of Classical Equipartition
    The limits are. • T → ∞ (β → 0), c → k,. • T → 0 (β → ∞), c → 0, exponentially fast. The Einstein theory of specific heats consisted in multiplying this by 3N, ...
  14. [14]
    [PDF] Lecture 22: 12.02.05 The Boltzmann Factor and Partition Function
    Aug 14, 2006 · • The Boltzmann factor and partition function are the two most important quantities for making statistical mechanical calculations. If we have ...
  15. [15]
    [DOC] Derivation of Boltzmann factor - BYU Physics and Astronomy
    ... Boltzmann's constant” and is given the symbol kB: Definition of entropy (2). Boltzmann's constant has the experimentally measured value of 1.38e-23 J/K. Back ...
  16. [16]
    [PDF] Notes on the Boltzmann distribution - Casegroup
    What we have shown is that this Boltzmann distribution maximizes the entropy of a system in thermal equilibrium with other systems kept at a temperature T.
  17. [17]
    [PDF] Probability distributions and maximum entropy
    In thermodynamics, the distribution arising in Theorem 5.11 is the Maxwell–Boltzmann energy distribution of a system of non-interacting particles in ...
  18. [18]
    [PDF] Lecture 7: Ensembles
    This is the Boltzmann distribution. Within the context of the microcanonical ensemble, we also derived the Boltzmann distribution using the principle of maximum ...
  19. [19]
    [PDF] 4. Blackbody Radiation, Boltzmann Statistics, Temperature, and ...
    4.2 Boltzmann statistics. Consider a simplified set of energy states an atom or molecule may occupy (Figure 4.1). There are a number of discrete energy states.
  20. [20]
    [PDF] Lecture 23: Maxwell Distribution, Partition Functions and Free Energy
    The first factor should be just the Boltzmann factor. D(v) ∝ eE/kT = e. -mv2/2kT. (23.2). This only accounts for an ideal gas, where the transnational motion ...<|control11|><|separator|>
  21. [21]
    [PDF] Boltzmann Distribution and Partition Function - MIT OpenCourseWare
    From the Boltzmann distribution, we relate the Helmholtz free energy to the spectrum of probabilities to be in various microstates.
  22. [22]
    Translation of Ludwig Boltzmann's Paper “On the Relationship ...
    Confusingly, Planck later chose to write Boltzmann's equation for entropy as S=klnW+constant [6]. Crucially, Ω is an extensive quantity, which leads to ...<|separator|>
  23. [23]
    [PDF] Elementray Principles in Statistical Mechanics. - Project Gutenberg
    Jan 22, 2016 · The laws of thermodynamics may be easily obtained from the principles of statistical mechanics, of which they are the incomplete expression, ...
  24. [24]
    [PDF] A Mathematical Theory of Communication
    The maximum entropy source is then the first approximation to English and its entropy determines the required channel capacity. As a simple example of some of ...
  25. [25]
    [PDF] Information Theory and Statistical Mechanics
    4, 620-630, May 15, 1957. Printed in U. S. A.. Information Theory and Statistical Mechanics. E. T. JAYNES. Department of Physics, Stanford University, Stanford ...
  26. [26]
    [PDF] On the 100th anniversary of the Sackur–Tetrode equation - arXiv
    Jan 23, 2013 · Sackur and Tetrode determined the volume of phase space cells as hn where h is Planck's constant and n is the number of degrees of freedom.
  27. [27]
    [PDF] Statistical Physics - DAMTP
    1.2.3 An Example: The Two State System. Consider a system of N non ... the Boltzmann distribution forces the system into its ground state (i.e. the state with.
  28. [28]
    [PDF] THE GIBBS PARADOX
    In the present note we consider the \Gibbs Paradox" about entropy of mixing and the logically inseparable topics of reversibility and the extensive property of ...
  29. [29]
  30. [30]
    Updates to Fundamental Constants Reflect Redefinition of ...
    Dec 16, 2021 · This update of values for over 300 constants is the first since a major redefinition of the International System of Units (SI).<|control11|><|separator|>
  31. [31]
    Physical Constants - PVEducation.Org
    Physical Constants ; k, 1.3806488 × 10-16 erg/K · 1.3806488 × 10-23 joule/K · Boltzmann's constant ; σ, 5.67 × 10-8 J/m2s K · Stefan-Boltzmann constant ; kT/q ...<|control11|><|separator|>
  32. [32]
    The theory of p-n junctions in semiconductors and p-n ... - IEEE Xplore
    The theory of p-n junctions in semiconductors and p-n junction transistors. Abstract: In a single crystal of semiconductor the impurity concentration may vary ...
  33. [33]
    [PDF] Diode-Based Temperature Measurement - Texas Instruments
    This linear relationship between forward voltage and temperature is the reason why diodes can be used as temperature measurement devices. A diode's data sheet ...
  34. [34]
    [PDF] Johnson Noise and Shot Noise: The Determination of the Boltzmann ...
    The only atomic constant that occurs in Nyquist's theoretical expression for the Johnson noise voltage is the Boltzmann constant k. Therefore, as a result ...
  35. [35]
    [PDF] Electrons and Holes in Semiconductors
    (c) In a silicon sample at T = 300 K, the Fermi level is located at 0.26 eV (10 kT) above the intrinsic Fermi level. What are the hole and electron ...
  36. [36]
    On deriving the Maxwellian velocity distribution - AIP Publishing
    Jul 1, 2013 · Maxwell's 1860 derivation of the molecular velocity distribution does not constitute a valid approach for pedagogical use.Missing: origins | Show results with:origins
  37. [37]
    [PDF] The early phase of Boltzmann's H-theorem (1868-1877)
    In 1868 Boltzmann published a paper devoted to a thorough study of the equilib- ... ment, Boltzmann's 1877 paper does not concern the H-theorem at all.
  38. [38]
    Jean Baptiste Perrin – Nobel Lecture - NobelPrize.org
    65 x 1022 by means of the rotational Brownian movement; or, as a crude average, 64 x 1022.
  39. [39]
    [PDF] The Mole, Avogadro's Number and Albert Einstein
    Mar 24, 2021 · The X-ray crystallographic method of determining Avogadro's number also shows a precise deter- mination of Planck's constant, h, which is ...<|control11|><|separator|>
  40. [40]
    [PDF] CODATA recommended values of the fundamental physical constants
    Sep 26, 2016 · This paper gives the 2014 self-consistent set of values of the constants and conversion factors of physics and chemistry recommended by the ...
  41. [41]
    Determinations of the Boltzmann constant - ScienceDirect
    We review measurements of the Boltzmann constant, k, the value of which is soon to be fixed at exactly 1.380 649 × 10 − 23 J⋅K−1 for the future revised ...Missing: symbol origin
  42. [42]
    [PDF] SI Brochure - 9th ed./version 3.02 - BIPM
    May 20, 2019 · The Boltzmann constant k is a proportionality constant between the quantities temperature (with the unit kelvin) and energy (with the unit ...
  43. [43]
    Towards realising the redefined kelvin - ScienceDirect.com
    Sep 30, 2022 · In 2019 the kelvin was redefined in terms of the Boltzmann constant. Here we outline the implications of that redefinition for the ...<|control11|><|separator|>
  44. [44]
  45. [45]
    molar gas constant - CODATA Value
    Concise form, 8.314 462 618... J mol-1 K ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  46. [46]
    [PDF] Natural units paperFin - arXiv
    In natural units (where c=h. –. =k= 1), the Stefan-. Boltzmann constant has the value π2/60 and is exact. Furthermore, by a suitable redefinition of the kelvin ...
  47. [47]
    [PDF] Constants and Conversions
    dimension of our set of ¯h = c = 1 natural units. This dimension is commonly ... Boltzmann constant. Then from E = kBT and kB = 8.617×10−14 GeV K−1 ...
  48. [48]
    [PDF] A Bayesian theory of quantum gravity without gravitons
    We use Planck units (or sometimes referred to as natural units) throughout this article. - thus, Boltzmann constant kB = 1, speed of light c = 1, reduced ...
  49. [49]
    Essential Statistical Thermodynamics - CCCBDB
    By convention, energies from ab initio calculations are reported in hartrees, the atomic unit of energy (1 hartree = 2625.5 kJ/mol = 627.51 kcal/mol = 219474.6 ...Missing: exact | Show results with:exact
  50. [50]
    Natural units, numbers and numerical clusters - Inspire HEP
    The so called `natural' units represent the unique system of units conveniently used in the realm of High Energy Physics. The system of natural units is defined ...<|separator|>
  51. [51]
    [PDF] 29. Cosmic Microwave Background - Particle Data Group
    Dec 1, 2021 · .260 eV cm−3, and a fraction of the critical density ... Figure 29.2: CMB temperature anisotropy band-power estimates from the Planck, WMAP, ACT,.
  52. [52]
    [PDF] Basic Units and Introduction to Natural Units - UF Physics Department
    Aug 24, 2015 · In this course, we follow researchers in particle physics, nuclear physics and astrophysics in adopting “natural units”, where = 1 and c = 1 ...
  53. [53]
    The Uses of Natural Units | Ben Brubaker
    Dec 15, 2020 · Natural units are a convenient choice for questions involving light or other forms of electromagnetic radiation; more mundane units are more suitable in most ...