Fact-checked by Grok 2 weeks ago

Electronvolt

The electronvolt (symbol: eV) is a unit of commonly used in physics, defined as the amount of gained or lost by a single unbound when accelerated through an electrostatic potential difference of one volt in . This unit equals exactly 1.602176634 × 10-19 joules. Due to its scale aligning with atomic, nuclear, and subatomic processes, the electronvolt is particularly convenient for expressing energies in , , and , where values in joules would be impractically small. For larger energies, multiples such as kiloelectronvolt (keV; 103 eV), megaelectronvolt (MeV; 106 eV), gigaelectronvolt (GeV; 109 eV), and teraelectronvolt (TeV; 1012 eV) are employed, with the latter relevant to high-energy particle accelerators like those at . The electronvolt also facilitates descriptions of energies in and binding energies in atomic and molecular systems.

Core Concepts

Definition

The electronvolt (eV) is a unit of energy defined as the amount of kinetic energy gained or lost by a single unbound electron when accelerated through an electric potential difference of one volt, in vacuum. Under the 2019 redefinition of the SI base units, this definition establishes the electronvolt as an exactly defined quantity tied to the fixed value of the elementary charge. The symbol for the unit is eV, with common multiples in scientific notation including keV ($10^3 eV), MeV ($10^6 eV), GeV ($10^9 eV), and TeV ($10^{12} eV); these are routinely employed in high-energy physics to denote energy scales spanning many orders of magnitude. This unit derives from the fundamental relation for electrostatic potential energy, E = qV, where q is the charge of the electron (the elementary charge e) and V is the potential difference. With e fixed at the exact value $1.602176634 \times 10^{-19} C and V = 1 V (where 1 V = 1 J/C), it follows that $1eV= e \times 1J/C= 1.602176634 \times 10^{-19}$ J exactly. The electronvolt provides a convenient scale for energies in , , and , where typical values range from a few eV (e.g., atomic transitions) to TeV (e.g., collider experiments), far below the joule thresholds relevant to everyday macroscopic phenomena.

History

The unit of the electronvolt emerged in the early 20th century, shortly after J. J. Thomson's discovery of the in 1897, as researchers sought convenient ways to express the energies of in atomic and photoelectric experiments. Robert A. Millikan played a key role in its introduction during his investigations of the , where he measured the energy required to eject from metals in terms of the electron's charge times a potential difference; his 1911 paper on the topic laid foundational measurements that implicitly relied on this energy scale, though the explicit term appeared in his subsequent work. The oil-drop experiment by Millikan, conducted between 1909 and 1913, further refined the value of the e, providing the numerical basis for quantifying electron energies in volts, which influenced the unit's development in . The term "electron-volt" was first explicitly used in scientific literature in 1925 by in a discussing electron emission from metals, where he equated one electron-volt to approximately 1.59 × 10^{-12} erg. Initially written as "electron-volt" with a , reflecting its composite nature as the energy gained by an accelerated through one volt, the naming evolved over the mid-20th century to the unhyphenated "electronvolt" in standard usage, particularly as experiments proliferated. This evolution was driven by practical needs in quantifying binding energies and ionization potentials, building on Millikan's precise determination of e from oil-drop experiments. The electronvolt gained widespread adoption in following the invention of the by Ernest O. Lawrence in 1929–1930, which accelerated particles to energies expressed in electronvolts, such as protons reaching 80,000 eV in early models. By , as accelerators like the enabled higher-energy collisions, the unit became indispensable for describing particle kinematics and interaction energies, supplanting less convenient units like joules in high-energy contexts. The International Bureau of Weights and Measures (BIPM) formalized the electronvolt as an accepted non- unit compatible with the () following the SI's establishment at the 11th General Conference on Weights and Measures in 1960, with it first appearing in the inaugural edition of the SI Brochure in 1970. A significant milestone occurred with the 2019 revision of the , effective May 20, 2019, which fixed the e exactly at 1.602176634 × 10^{-19} C, rendering the electronvolt exactly 1.602176634 × 10^{-19} J without experimental uncertainty and linking it precisely to the SI base units. This redefinition, adopted by the 26th General Conference on Weights and Measures, eliminated variability in the unit's value arising from measurements of e, enhancing its stability for applications in and beyond.

Relation to SI Units

The electronvolt (eV) is precisely related to the joule (J), the SI unit of energy, by the exact conversion factor established following the 2019 redefinition of the SI base units: $1 \, \mathrm{eV} = 1.602176634 \times 10^{-19} \, \mathrm{J} This relation is exact because the e is now defined as exactly $1.602176634 \times 10^{-19} \, \mathrm{C}, and the volt (V) is defined through fixed values of other fundamental constants such as the h and the c, thereby eliminating any in the product e \times 1 \, \mathrm{V}. The inverse conversion is likewise exact: $1 \, \mathrm{J} = \frac{1}{1.602176634 \times 10^{-19}} \, \mathrm{eV} = 6.241509074 \times 10^{18} \, \mathrm{eV} In microscopic physical systems, such as those involving atomic or subatomic particles, the electronvolt provides a convenient scale because corresponding energies expressed in joules are impractically small. For instance, the of the is 13.6 , equivalent to approximately $2.18 \times 10^{-18} \, \mathrm{J}, avoiding cumbersome numerical representations in calculations. The electronvolt is recognized as a non-SI unit accepted for use with the () by the International Bureau of Weights and Measures (BIPM). It is routinely employed in international standards, such as the CODATA recommended values for fundamental physical constants, where many quantities like particle rest masses and transition energies are tabulated in electronvolts for precision and practicality.

Connections to Physical Properties

Mass Equivalence

In special relativity, the rest energy E of a particle is related to its rest mass m by Einstein's equation E = mc^2, where c is the speed of light in vacuum, exactly $299792458 m/s. Thus, the equivalent rest mass corresponding to an energy in electronvolts can be expressed as m = E / c^2, with E in eV yielding m in eV/c^2. In particle physics, this convention is standard, where particle masses are routinely tabulated in units of eV/c^2 to directly reflect their rest energies in eV when multiplied by c^2. For the electron, the rest energy is precisely $0.51099895069(16) MeV, so its rest mass is m_e = 0.51099895069(16) MeV/c^2. To convert to SI units, note that $1 /c^2 = 1.782661921 \times 10^{-36} kg, derived from the and via $1 eV = 1.602176634 \times 10^{-19} J and c^2. Thus, the electron mass is approximately $9.1093837015 \times 10^{-31} kg. The proton provides another key example, with a rest energy of $938.27208816(29) MeV, corresponding to a rest mass of $938.27208816(29) MeV/c^2 \approx 1.67262192369(51) \times 10^{-27} . Particle data tables, such as those from the Particle Data Group, quote masses for a wide range of fundamental particles and composites in eV/c^2, facilitating comparisons of their intrinsic . While the electronvolt equivalence primarily addresses rest mass for particles where the rest energy dominates, in the non-relativistic limit—where \ll mc^2—total energy approximates rest energy plus small kinetic contributions, but the focus here remains on the intrinsic rest mass scale.

Momentum

In the non-relativistic regime, applicable when the kinetic energy E is much smaller than the particle's rest energy m c^2, the magnitude of the p is given by the formula p = \sqrt{2 m E}, derived from the classical kinetic energy expression E = \frac{1}{2} m v^2 and the definition p = m v. For an electron (m = 9.1093837015 \times 10^{-31} kg) with E = 1 eV ($1.602176634 \times 10^{-19} J), this yields p \approx 5.40 \times 10^{-25} kg·m/s./09%3A_Relativity/9.04%3A_Relativistic_Momentum) In relativistic contexts, particularly in particle physics, momentum is commonly expressed in units of eV/c, where $1 eV/c \approx 5.344 \times 10^{-28} kg·m/s, obtained by dividing the energy equivalent of 1 eV by the speed of light c = 299792458 m/s. For higher energies where relativistic effects are significant, the momentum is calculated using the relation p = \frac{1}{c} \sqrt{E^2 - (m c^2)^2}, with E as the total (kinetic plus rest m c^2). This formula reduces to the non-relativistic form when E \approx m c^2 (i.e., low speeds, v \ll c), yielding p \approx \sqrt{2 m (E - m c^2)}; in the ultra-relativistic limit where E \gg m c^2, it approximates to p \approx E / c. For an with 1 MeV kinetic (rest m c^2 = 0.511 MeV, total E = 1.511 MeV), the relativistic is p \approx 1.422 MeV/c, whereas the non-relativistic gives p \approx 1.011 MeV/c, demonstrating how increases the for a given kinetic by about 41%.

Length Scales

In , the serves as a fundamental length scale tied to a particle's rest mass , marking the regime where quantum effects become prominent in scattering processes. For a particle of rest E_0 = m c^2, the is given by \lambda_C = \frac{h}{m c} = \frac{h c}{E_0}, while the reduced , more commonly used in relativistic , is \bar{\lambda}_C = \frac{\hbar}{m c} = \frac{\hbar c}{E_0}. For the , with E_0 \approx 0.511 MeV, the reduced is \bar{\lambda}_C \approx 3.8616 \times 10^{-13} m. This scale inversely relates to the rest in electronvolts: higher E_0 yields shorter , reflecting the particle's quantum size. For photons, which are massless, an analogous characteristic length scale emerges from their E, given by l \approx \frac{\hbar c}{E}, derived from the de Broglie relation and relativistic energy-momentum equivalence E = p c, where the full is \lambda = \frac{h c}{E} = 2\pi \frac{\hbar c}{E}; this reduced scale \frac{\hbar c}{E} sets the resolution limit in high-energy interactions. The de Broglie wavelength extends this concept to moving particles, associating a wave-like length scale with their momentum p, via \lambda = \frac{h}{p}. In the non-relativistic limit, where kinetic energy K \ll m c^2, the momentum is p \approx \sqrt{2 m K}, so \lambda \approx \frac{h}{\sqrt{2 m K}}; this directly links electronvolt-scale energies to spatial extents relevant for diffraction and interference. For an electron with K = 1 eV, the de Broglie wavelength is approximately 1.23 nm, comparable to atomic spacings and thus crucial for understanding electron microscopy and low-energy scattering. At higher energies approaching the relativistic regime, the full relation p = \sqrt{E_0^2 + 2 E_0 K + K^2}/c (with total energy E_0 + K) refines this, but the inverse scaling with \sqrt{K} persists for electronvolt orders of magnitude below the rest energy. In nuclear physics, electronvolt energies connect to femtometer-scale lengths through the natural unit \hbar c \approx 197.3 MeV fm, enabling straightforward conversions between energy and distance via the characteristic scale l \approx \frac{\hbar c}{E}. This relation arises from uncertainty principle arguments, where probing a length l requires momentum p \sim \hbar / l and thus energy E \sim p c \sim \hbar c / l for relativistic particles or photons. For instance, energies around 200 MeV—common in nuclear reactions—correspond to l \approx 1 fm, matching the typical size of atomic nuclei (e.g., proton radius \sim 0.8 fm). Such scales underpin models like the liquid drop model and shell model, where MeV binding energies imply fm-dimensional wavefunctions.

Temperature Equivalence

The temperature T in corresponding to an E in electronvolts is given by the relation T = \frac{E}{k}, where k is the , which links to in the for ideal gases and s. This equivalence arises from the definition of , where the average per degree of freedom for particles in is \frac{1}{2} kT, but in plasma and high-energy contexts, the characteristic energy scale is often expressed as kT for simplicity. The exact value of the , fixed by the 2019 SI redefinition, is k = 8.617333262 \times 10^{-5} eV/K. The conversion factor for 1 to follows directly from this : T = \frac{1 \, \mathrm{[eV](/page/EV)}}{k} \approx 11604.525 K, derived by substituting the exact value of k into the equation, providing a precise scaling between electronvolt energies and thermal scales without additional assumptions. This relation enables straightforward interconversion; for instance, energies below 1 correspond to terrestrial temperatures, while keV scales reach extreme astrophysical conditions. Representative examples illustrate the scale: at of approximately 300 K, the kT \approx 0.0259 eV, representing the typical of particles in ambient conditions. In contrast, the core of reaches about 15 million K (15 MK), equivalent to a of roughly 1.3 keV, where dominates due to these high energies overcoming electrostatic barriers. This equivalence is particularly relevant in astrophysics for modeling stellar interiors, where temperatures in millions of kelvin translate to keV energies driving fusion and radiative processes, and in plasma physics for characterizing hot, ionized gases in fusion devices or space environments. Additionally, it facilitates comparisons between thermal energies kT and atomic ionization potentials (typically 5–25 eV), determining the degree of thermal ionization in plasmas via the Saha equation, where kT approaching or exceeding these potentials leads to significant ionization fractions.

Wavelength

The wavelength \lambda of electromagnetic radiation corresponding to a photon of energy E is given by the relation E = \frac{h c}{\lambda}, where h is Planck's constant and c is the speed of light in vacuum. Rearranging yields \lambda = \frac{h c}{E}, with h c \approx 1.986 \times 10^{-25} J\cdotm derived from fundamental constants. In practical units for spectroscopy, where E is expressed in electronvolts (eV) and \lambda in nanometers (nm), this simplifies to \lambda \approx \frac{1239.8}{E} \ \text{nm}, with E in eV; the conversion factor 1239.8 eV\cdotnm accounts for the elementary charge and unit scaling from the SI value of h c. This photon energy-wavelength relation defines key regions of the electromagnetic spectrum. For visible light, spanning photon energies of 1.65–3.1 eV, the corresponding wavelengths range from 750 nm (red) to 400 nm (violet). In the X-ray domain, energies on the order of 1 keV produce wavelengths around 1 nm, with the full soft-to-hard X-ray band extending from approximately 0.01 nm (at ~100 keV) to 10 nm (at ~0.1 keV). The (UV) region borders visible light at higher energies, covering 3.1 (400 ) to about 124 (10 ), while the infrared (IR) region extends to lower energies from 1.77 (700 ) down to roughly 0.001 (1 ). These boundaries highlight how electronvolt energies map to domains essential for identifying material properties via interactions. The relation also underpins , where the ties emission line wavelengths to quantized energy differences in ; for , the ground-state corresponds to 13.606 .

Experimental Applications

Scattering Experiments

Scattering experiments utilize electronvolt-scale energies to probe the internal structure of atoms, nuclei, and subnuclear particles by leveraging the wave-like properties of electrons and photons, where the de Broglie wavelength determines the spatial resolution. In these experiments, the incident particle's energy, measured in electronvolts (eV), mega-electronvolts (MeV), giga-electronvolts (GeV), or tera-electronvolts (TeV), inversely relates to the probe's wavelength, approximately given by \lambda \approx \frac{\hbar c}{E}, allowing resolution of features down to femtometer (fm) scales. Early experiments, analogous to Rutherford's alpha-particle scattering but using s to avoid strong interactions, revealed charge distributions. In the 1950s, and collaborators at Stanford used electron beams with energies around 100–200 MeV to scatter off nuclei like carbon and oxygen, resolving charge radii on the order of 3–5 fm. For instance, 100 MeV electrons provide a of approximately 2 fm (\lambda \approx \frac{197 \text{ MeV·fm}}{100 \text{ MeV}}), sufficient to probe sizes on the order of a few fm, such as the ~2.5 fm charge radius of carbon, confirming the diffuse nature of the charge rather than a point-like structure. Compton scattering involves photons scattering off free or loosely bound electrons, demonstrating the particle nature of light and providing insights into electron kinematics at keV energies. Arthur Compton's 1923 experiments used X-rays from the molybdenum Kα line with an energy of about 17 keV incident on , observing a wavelength shift in the scattered photons described by \Delta \lambda = \frac{h}{m_e c} (1 - \cos \theta), where h is Planck's constant, m_e is the , c is the , and \theta is the scattering angle; the Compton wavelength \frac{h}{m_e c} \approx 2.43 \times 10^{-12} m corresponds to an energy scale of 511 keV, the electron rest mass. This shift, maximized at \theta = 180^\circ to about 4.86 pm, confirmed energy-momentum conservation in quantum terms and is foundational for understanding scattering in the keV regime. Deep inelastic scattering (DIS) at GeV energies has been pivotal in elucidating nucleon substructure, particularly the quark model. At the Stanford Linear Accelerator Center (SLAC) in the late 1960s and early 1970s, Jerome Friedman, Henry Kendall, and Richard Taylor directed electron beams up to 20 GeV onto liquid hydrogen and deuterium targets, observing inelastic events where electrons scattered at wide angles, indicating interactions with point-like constituents inside protons. These experiments, with momentum transfers Q^2 up to several GeV², revealed scaling behavior consistent with scattering off fractionally charged quarks, as predicted by deep inelastic cross-section formulas; for example, at 7–20 GeV, the structure function F_2(x) showed a rise at low Bjorken x, supporting the parton model. This work earned the 1990 Nobel Prize in Physics and paved the way for quantum chromodynamics (QCD). Modern accelerators like the Large Hadron Collider (LHC) extend DIS-like probes to TeV scales in proton-proton collisions, using eV-based energy units to map quark-gluon dynamics in high-density environments.

Spectroscopy and Binding Energies

In , electronvolt units are commonly used to quantify the differences between states in atoms, which correspond to transitions observed in emission or absorption spectra. For the , the involves transitions from excited states (n ≥ 2) to the (n=1), with differences ranging from approximately 10.2 eV for the n=2 to n=1 to 13.6 eV approaching the limit. These values, derived from precise measurements of , highlight how eV-scale govern and de-excitation processes, enabling the mapping of diagrams. In multi-electron atoms, binding —representing the required to remove an outer-shell electron—typically fall in the 5–15 eV range, varying with ; for example, sodium's first is 5.14 eV, while chlorine's is 12.97 eV. Photoelectron spectroscopy employs electronvolts to measure ionization potentials and binding energies by ejecting electrons with photons or electron beams and analyzing their kinetic energies. In ultraviolet photoelectron spectroscopy (UPS), valence electron binding energies are probed, often revealing molecular orbital structures in the eV range. For metals, the work function—the minimum energy to extract an electron from the surface—ranges from about 4 to 5 eV, as seen in aluminum (4.08 eV) and (4.7 eV), which determines thresholds for photoelectric emission. This technique provides direct insight into surface electronic properties, with binding energies referenced to the vacuum level. In molecular , electronvolts describe finer scales for vibrational and rotational transitions superimposed on electronic levels, as well as overall energies. Vibrational and rotational energies in diatomic molecules typically span 0.01–0.1 eV, corresponding to and spectral regions; for instance, the vibrational spacing in HCl is about 0.37 eV, while rotational levels are much smaller. Molecular energies, the to break bonds into neutral fragments, range from 1–10 eV, such as 4.52 eV for the H-H bond in H₂ and 9.92 eV for N≡N in N₂, influencing stability and reactivity. These eV-scale measurements via techniques like absorption spectroscopy elucidate intramolecular dynamics. X-ray photoelectron spectroscopy (XPS) extends eV measurements to core-level binding energies, using high-energy X-ray sources (hundreds to thousands of eV) to probe inner-shell electrons. Core binding energies range from ~100 eV for light elements (e.g., carbon 1s at 284 eV) to over 1000 eV for heavier atoms (e.g., copper 1s at 8979 eV, though typically analyzed in lower shells), revealing chemical shifts due to local environments. Developed by and colleagues, XPS has become essential for surface analysis, with binding energies calibrated against standards like adventitious carbon at 284.8 eV.

Energy Comparisons

Comparisons to Other Energies

The electronvolt (eV) is a unit particularly suited to describing energies at the atomic and subatomic scales, far smaller than those encountered in everyday macroscopic phenomena. For instance, typical energies range from about 1 to 10 eV per bond. The dissociation energy of the hydrogen molecule (H-H bond) is specifically 4.52 eV. In contrast, the required to remove an from a is 13.6 eV, underscoring the eV's relevance to processes like bond breaking and . At the nuclear scale, energies are orders of magnitude larger, emphasizing the eV's progression through multiples like mega-eV (MeV). of releases approximately 200 MeV per event, equivalent to the energy from splitting one heavy into lighter fragments. , another nuclear process, typically liberates around 5 MeV, manifested as the of the emitted . These values highlight how the eV scales up to capture the vastly greater binding forces within atomic compared to electron orbitals. Everyday energies dwarf the single-particle scales of the eV, illustrating its microscopic focus. A visible photon carries roughly 2 eV, corresponding to wavelengths between (~1.8 eV) and (~3.1 eV) that humans perceive. In household electricity at a typical voltage of 120 V ( standard), the energy per accelerated across the potential is 120 eV, but a macroscopic unit like 1 (kWh) equates to about 2.25 × 10^{25} eV, representing the collective from trillions of such electrons powering appliances. On cosmic scales, the eV spans from minuscule thermal backgrounds to extreme particle accelerations. Photons in the (CMB), the relic radiation from the , have an average energy of approximately 6.35 × 10^{-4} , reflecting the universe's current temperature of 2.725 K. At the opposite extreme, the highest-energy cosmic rays—protons or nuclei accelerated by astrophysical phenomena—reach up to 3 × 10^{20} , equivalent to the of a fast-pitched yet confined to a single . These comparisons reveal the eV's versatility in bridging quantum micro-energies to the universe's most violent processes.

Molar Energy Equivalents

In and , electronvolt energies are frequently scaled to molar quantities to align with thermodynamic conventions, where the energy per mole of particles is obtained by multiplying the per-particle value by Avogadro's constant. Specifically, 1 eV per equates to exactly 96.48533212 /mol, derived from the product of the and Avogadro's constant, which yields the of 96.48533212 /(V·mol). This precise conversion factor, established by the 2019 redefinition of the SI units, enables seamless integration of atomic-scale electronvolt measurements into macroscopic chemical analyses. A key application arises in bond dissociation energetics, where values are commonly reported in kJ/mol for bulk reactions but can be equivalently expressed in eV for molecular-level interpretations. For instance, the dissociation energy of a carbon-carbon is approximately 3.6 eV per , corresponding to 348 kJ/mol, highlighting how electronvolts bridge quantum mechanical calculations with experimental . In , standard electrode potentials—tabulated in volts—directly translate to per transferred, as the potential difference represents the energy change in eV for each in the half-reaction. This equivalence, where 1 V corresponds to 1 eV per , underpins the analysis of processes on scales; for example, the standard potential for the Zn²⁺/Zn couple at -0.76 V implies a 0.76 eV energy per for the oxidation, scaled to approximately -147 kJ/mol for the two-electron process. Lattice energies of ionic compounds further illustrate molar electronvolt equivalents, quantifying the electrostatic in structures. These energies typically span 10–30 eV per pair for common salts with divalent cations, equivalent to roughly 0.96–2.89 /, as seen in compounds like MgO (approximately 39 eV or 3.79 /, adjusted for the ) where stronger interactions elevate the values. Such conversions are essential for predicting and reactivity in , converting quantum defect models into practical thermodynamic data.

References

  1. [1]
    electron volt - CODATA Value
    Concise form, 1.602 176 634 x 10-19 J ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  2. [2]
    Energies in electron volts - HyperPhysics
    Electron Volts. A convenient energy unit, particularly for atomic and nuclear processes, is the energy given to an electron by accelerating it through 1 volt ...
  3. [3]
    19.1 Electric Potential Energy: Potential Difference - UCF Pressbooks
    The electron volt (eV) is the most common energy unit for submicroscopic processes. This will be particularly noticeable in the chapters on modern physics.<|control11|><|separator|>
  4. [4]
    Fermilab | Science | Inquiring Minds | Questions About Physics
    Apr 28, 2014 · Let us start with eV (electron-volt). It is an energy unit used in High-energy physics. A MeV is the Mega electron-volt, e.g. million times more ...
  5. [5]
    Anatomy of an Electromagnetic Wave - NASA Science
    Aug 3, 2023 · An electron volt is the amount of kinetic energy needed to move an electron through one volt potential. Moving along the spectrum from long to ...
  6. [6]
    [PDF] SI Brochure - 9th ed./version 3.02 - BIPM
    May 20, 2019 · Thus, this redefinition marks a significant and historic step forward. The changes were agreed by the CGPM in November 2018 with effect from May ...
  7. [7]
  8. [8]
    Robert A. Millikan and the Oil Drop Experiment - AIP Publishing
    Oct 1, 2019 · Millikan was not the only scientist around this time attempting to find the charge of the electron. In his 1911 paper, Millikan compared his ...
  9. [9]
    Early Particle Accelerators - Ernest Lawrence and the Cyclotron
    Combined with a reasonable magnetic field, a potential on the electrodes of only ten thousand volts could accelerate an alpha particle to one million electron ...
  10. [10]
  11. [11]
    joule-electron volt relationship - CODATA Value
    joule-electron volt relationship 1 J. Numerical value, 6.241 509 074... x 1018 eV. Standard uncertainty, (exact). Relative standard uncertainty, (exact).
  12. [12]
  13. [13]
    [PDF] 1. Physical Constants - Particle Data Group
    Table 1.1: Revised 2024 by D. Robinson (LBNL) and P.A. Zyla (LBNL) ... proton mass mp. 938.272 088 16(29) MeV/c2 = 1.672 621 923 69(51)×10−27 kg.
  14. [14]
    electron mass energy equivalent in MeV - CODATA Value
    electron mass energy equivalent in MeV $m_{\rm e}c^2$ ; Numerical value, 0.510 998 950 69 MeV ; Standard uncertainty, 0.000 000 000 16 MeV.Missing: 2018 | Show results with:2018
  15. [15]
  16. [16]
  17. [17]
    Fundamental Physical Constants from NIST
    ### Extracted Values and Summary
  18. [18]
    Relativistic Energy | Physics - Lumen Learning
    The equation E2 = (pc)2 + (mc2)2 relates the relativistic total energy E and the relativistic momentum p. At extremely high velocities, the rest energy mc2 ...
  19. [19]
    Relativistic Momentum - HyperPhysics
    The relativistic momentum is given by which is the ordinary definition of momentum with the mass replaced by the relativistic mass.
  20. [20]
    reduced Compton wavelength - CODATA Value
    reduced Compton wavelength $\lbar_{\rm C}$ ; Numerical value, 3.861 592 6744 x 10-13 m ; Standard uncertainty, 0.000 000 0012 x 10-13 m.
  21. [21]
  22. [22]
    DeBroglie Wavelength - HyperPhysics
    For an electron with KE = 1 eV and rest mass energy 0.511 MeV, the associated DeBroglie wavelength is 1.23 nm, about a thousand times smaller than a 1 eV photon ...
  23. [23]
  24. [24]
    [PDF] CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL ...
    CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL PHYSICAL CONSTANTS: 2018. NIST SP 961 (May 2019). An extensive list of constants is available on the NIST ...Missing: rest | Show results with:rest
  25. [25]
    Our Sun: Facts - NASA Science
    The temperature in the Sun's core is about 27 million degrees Fahrenheit (15 million degrees Celsius) – hot enough to sustain nuclear fusion. This creates ...
  26. [26]
    [PDF] Lecture Notes on Introduction to Plasma Astrophysics
    An energy of 1 eV corresponds to a temperature of ∼104 K (more precisely ... Kulsrud, R. M. 2005 Plasma physics for astrophysics. Princeton University ...
  27. [27]
    Thermal Ionization - an overview | ScienceDirect Topics
    The kinetic energy of an atom at a given temperature is equal to (3/2) kT where k is the Boltzman constant (8.63 × 10−5 eV/atom°K) and T is the temperature in ...
  28. [28]
    Electromagnetic Spectrum
    Infrared · Frequencies: .003 - 4 x 1014 Hz. Wavelengths: 1 mm - 750 nm. Quantum energies: 0.0012 - 1.65 eV ; Visible Light · Frequencies: 4 - 7.5 x 1014 Hz.
  29. [29]
    X-ray Imaging - Medical Imaging Systems - NCBI Bookshelf - NIH
    Note that the wavelength of most X-rays lies in the range of 0.01 nm up to 10 nm. This corresponds to an energy range of 100 keV down to 100 eV.Introduction · X-ray Generation · X-ray Matter Interaction · X-ray Imaging
  30. [30]
    Rydberg constant times hc in eV - CODATA Value
    Rydberg constant times hc in eV $hc\,R_\infty$ ; Numerical value, 13.605 693 122 990 eV ; Standard uncertainty, 0.000 000 000 015 eV.
  31. [31]
    [PDF] Electron Scattering for Nuclear and Nucleon Structure
    The scattering of high-energy electrons from nuclear and nucleon targets provides a mi- croscope for examining the structure of these tiny objects.
  32. [32]
    [PDF] DEEP INELASTIC SCATTERING - Nobel Prize
    Deep inelastic scattering is large energy loss scattering of electrons from nucleons, using high energy electrons to study nucleons at small distances.Missing: electronvolt | Show results with:electronvolt
  33. [33]
    Energy Levels of Neutral Hydrogen ( H I )
    NIST Physical Measurement Laboratory, Handbook of Basic Atomic Spectroscopic ... Hydrogen Neutral Atom Energy Levels Hydrogen References · Switch to ASCII ...
  34. [34]
    NIST: Atomic Spectra Database - Ionization Energies Form
    This form provides access to NIST critically evaluated data on ground states and ionization energies of atoms and atomic ions.
  35. [35]
    Work Functions for Photoelectric Effect - HyperPhysics
    Element, Work Function(eV). Aluminum. 4.08. Beryllium. 5.0. Cadmium. 4.07. Calcium. 2.9. Carbon. 4.81. Cesium. 2.1. Cobalt. 5.0. Copper. 4.7. Gold ...
  36. [36]
    [PDF] Tables of Molecular Vibrational Frequencies
    Selected values of the fundamental vibrational frequencies are given for each molecule, together with observed infrared and Raman spectral data and citations to ...
  37. [37]
    [PDF] Table 1-1. Electron binding energies, in electron volts, for the ...
    Electron binding energies, in electron volts, for the elements in their natural forms. Element. K 1s. L1 2s. L2 2p1/2. L3 2p3/2. M1 3s. M2 3p1/2. M3 3p3/2. M4 ...
  38. [38]
    Chemical Bonding in the H 2 Molecule
    For comparision, the experimental H-H bond length is 0.74 Å and the experimental bond dissociation energy is 436 kJ/mole or 4.52 eV. H-H Chemical Bond. As ...
  39. [39]
    Ionization Energy | Periodic Table of Elements - PubChem
    Ionization energy, also called ionization potential, is the amount of energy required to remove an electron from an isolated atom or molecule.
  40. [40]
    Nuclear Fission and Nuclear Fusion - Purdue University
    The induced fission of this isotope releases an average of 200 MeV per atom, or 80 million kilojoules per gram of 235U. The attraction of nuclear fission as a ...
  41. [41]
    [PDF] Week 5 Lecture 2 – Alpha Decay - MSU chemistry
    The typical energy released (Q-‐value) for these decays is on the order of 5 to 6 MeV – this energy is sufficient to use the alpha par5cles to induce.
  42. [42]
    Visible Light - UCAR Center for Science Education
    Red photons of light carry about 1.8 electron volts (eV) of energy, while each blue photon transmits about 3.1 eV. Visible light's neighbors on the EM ...Missing: authoritative | Show results with:authoritative
  43. [43]
    Convert Kilowatt-hour to Electron-volt
    The kilowatt-hour [kW*h] to electron-volt [eV] ... How to Convert Kilowatt-hour to Electron-volt. 1 kW*h = 2.246943266806E+25 eV 1 eV = 4.4504906499998E-26 kW*h.
  44. [44]
    New physics from the CMB - D. Scott
    The Cosmic Microwave Background (CMB) is by far the dominant background. The CMB corresponds to an energy density of 0.260 eV cm -3 , or a number density of ...
  45. [45]
    Mystery of the Highest Energy Cosmic Rays/ - HEASARC
    The highest energy of a detected cosmic ray is 3x1020 eV (electron volts), equivalent to about 12 calories, more than ten-million times greater than the ...
  46. [46]
  47. [47]
  48. [48]
    17.3 Electrode and Cell Potentials - Chemistry 2e | OpenStax
    Feb 14, 2019 · Potentials are measured in the volt unit, defined as one joule of energy per one coulomb of charge, V = J/C. When measured for purposes of ...<|separator|>