Tests of relativistic energy and momentum

Last updated
Kinetic energy in special relativity and Newtonian mechanics. Relativistic kinetic energy increases to infinity when approaching the speed of light, thus no massive body can reach this speed. Rel-Newton-Kinetic.svg
Kinetic energy in special relativity and Newtonian mechanics. Relativistic kinetic energy increases to infinity when approaching the speed of light, thus no massive body can reach this speed.

Tests of relativistic energy and momentum are aimed at measuring the relativistic expressions for energy, momentum, and mass. According to special relativity, the properties of particles moving approximately at the speed of light significantly deviate from the predictions of Newtonian mechanics. For instance, the speed of light cannot be reached by massive particles.

Contents

Today, those relativistic expressions for particles close to the speed of light are routinely confirmed in undergraduate laboratories, and necessary in the design and theoretical evaluation of collision experiments in particle accelerators. [1] [2] See also Tests of special relativity for a general overview.

Overview

Similar to kinetic energy, relativistic momentum increases to infinity when approaching the speed of light. Rel-Newton-Momentum.svg
Similar to kinetic energy, relativistic momentum increases to infinity when approaching the speed of light.

In classical mechanics, kinetic energy and momentum are expressed as

On the other hand, special relativity predicts that the speed of light is constant in all inertial frames of references. The relativistic energy–momentum relation reads:

,

from which the relations for rest energy , relativistic energy (rest + kinetic) , kinetic energy , and momentum of massive particles follow:

,

where . So relativistic energy and momentum significantly increase with speed, thus the speed of light cannot be reached by massive particles. In some relativity textbooks, the so-called "relativistic mass" is used as well. However, this concept is considered disadvantageous by many authors, instead the expressions of relativistic energy and momentum should be used to express the velocity dependence in relativity, which provide the same experimental predictions.

Early experiments

First experiments capable of detecting such relations were conducted by Walter Kaufmann, Alfred Bucherer and others between 1901 and 1915. These experiments were aimed at measuring the deflection of beta rays within a magnetic field so as to determine the mass-to-charge ratio of electrons. Since the charge was known to be velocity independent, any variation had to be attributed to alterations in the electron's momentum or mass (formerly known as transverse electromagnetic mass equivalent to the "relativistic mass" as indicated above). Since relativistic mass is not often used anymore in modern textbooks, those tests can be described of measurements of relativistic momentum or energy, because the following relation applies:

Electrons traveling between 0.25–0.75c indicated an increase of momentum in agreement with the relativistic predictions, and were considered as clear confirmations of special relativity. However, it was later pointed out that although the experiments were in agreement with relativity, the precision wasn't sufficient to rule out competing models of the electron, such as the one of Max Abraham. [3] [4]

Already in 1915, however, Arnold Sommerfeld was able to derive the Fine structure of hydrogen-like spectra by using the relativistic expressions for momentum and energy (in the context of the Bohr–Sommerfeld theory). Subsequently, Karl Glitscher simply substituted the relativistic expression's for Abraham's, demonstrating that Abraham's theory is in conflict with experimental data and is therefore refuted, while relativity is in agreement with the data. [5]

Precision measurements

Three data points of Rogers et al., in agreement with special relativity RogersExpGraph2.svg
Three data points of Rogers et al., in agreement with special relativity

In 1940, Rogers et al. performed the first electron deflection test sufficiently precise to definitely rule out competing models. As in the Bucherer-Neumann experiments, the velocity and the charge-mass-ratio of beta particles of velocities up to 0.75c was measured. However, they made many improvements, including the employment of a Geiger counter. The accuracy of the experiment by which relativity was confirmed was within 1%. [6]

An even more precise electron deflection test was conducted by Meyer et al. (1963). They tested electrons traveling at velocities from 0.987 to 0.99c, which were deflected in a static homogenous magnetic field by which p was measured, and a static cylindrical electric field by which was measured. They confirmed relativity with an upper limit for deviations of ~0.00037. [7]

Also measurements of the charge-to-mass ratio and thus momentum of protons have been conducted. Grove and Fox (1953) measured 385-MeV protons moving at ~0.7c. Determination of the angular frequencies and of the magnetic field provided the charge-to-mass ratio. This, together with measuring the magnetic center, allowed to confirm the relativistic expression for the charge-to-mass ratio with a precision of ~0.0006. [8]

However, Zrelov et al. (1958) criticized the scant information given by Grove and Fox, emphasizing the difficulty of such measurements due to the complex motion of the protons. Therefore, they conducted a more extensive measurement, in which protons of 660 MeV with mean velocity of 0.8112c were employed. The proton's momentum was measured using a Litz wire, and the velocity was determined by evaluation of Cherenkov radiation. They confirmed relativity with an upper limit for deviations of ~0.0041. [9]

Bertozzi experiment

Data of the Bertozzi experiment show close agreement with special relativity. Kinetic energy of five electron runs: 0.5, 1, 1.5, 4.5, 15 MeV (or 1, 2, 3, 9, 30 in mc2). Speed: 0.752, 0.828, 0.922, 0.974, 1.0 in c2 (or 0.867, 0.910, 0.960, 0.987, 1 in c). BertozziExp.svg
Data of the Bertozzi experiment show close agreement with special relativity. Kinetic energy of five electron runs: 0.5, 1, 1.5, 4.5, 15 MeV (or 1, 2, 3, 9, 30 in mc²). Speed: 0.752, 0.828, 0.922, 0.974, 1.0 in (or 0.867, 0.910, 0.960, 0.987, 1 in c).

Since the 1930s, relativity was needed in the construction of particle accelerators, and the precision measurements mentioned above clearly confirmed the theory as well. But those tests demonstrate the relativistic expressions in an indirect way, since many other effects have to be considered in order to evaluate the deflection curve, velocity, and momentum. So an experiment specifically aimed at demonstrating the relativistic effects in a very direct way was conducted by William Bertozzi (1962, 1964). [10] [11]

He employed the electron accelerator facility at MIT in order to initiate five electron runs, with electrons of kinetic energies between 0.5 and 15 MeV. These electrons were produced by a Van de Graaff generator and traveled a distance of 8.4 m, until they hit an aluminium disc. First, the time of flight of the electrons was measured in all five runs – the velocity data obtained were in close agreement with the relativistic expectation. However, at this stage the kinetic energy was only indirectly determined by the accelerating fields. Therefore, the heat produced by some electrons hitting the aluminium disc was measured by calorimetry in order to directly obtain their kinetic energy - those results agreed with the expected energy within 10% error margin.

Undergraduate experiments

Various experiments have been performed which, due to their simplicity, are still used as undergraduate experiments. Mass, velocity, momentum, and energy of electrons have been measured in different ways in those experiments, all of them confirming relativity. [12] They include experiments involving beta particles, Compton scattering in which electrons exhibit highly relativistic properties and positron annihilation.

Beta particles
Marvel et al. [12] 2011
Lund et al. [13] 2009
Luetzelschwab [14] 2003
Couch et al. [15] 1982
Geller et al. [16] 1972
Parker [17] 1972
Bartlett et al. [18] 1965
Compton recoil electrons
Jolivette et al. [19] 1994
Hoffman [20] 1989
Egelstaff et al. [21] 1981
Higbie [22] 1974
Positron annihilation
Dryzek et al. [23] 2006

Particle accelerators

In modern particle accelerators at high energies, the predictions of special relativity are routinely confirmed, and are necessary for the design and theoretical evaluation of collision experiments, especially in the ultrarelativistic limit. [2] For instance, time dilation must be taken into account to understand the dynamics of particle decay, and the relativistic velocity addition theorem explains the distribution of synchrotron radiation. Regarding the relativistic energy-momentum relations, a series of high precision velocity and energy-momentum experiments have been conducted, in which the energies employed were necessarily much higher than the experiments mentioned above. [24]

Velocity

Time of flight measurements have been conducted to measure differences in the velocities of electrons and light at the SLAC National Accelerator Laboratory. For instance, Brown et al. (1973) found no difference in the time of flight of 11-GeV electrons and visible light, setting an upper limit of velocity differences of . [25] Another SLAC experiment conducted by Guiragossián et al. (1974) accelerated electrons up to energies of 15 to 20.5 GeV. They used a radio frequency separator (RFS) to measure time-of-flight differences and thus velocity differences between those electrons and 15-GeV gamma rays on a path length of 1015 m. They found no difference, increasing the upper limit to . [26]

Already before, Alväger et al. (1964) at the CERN Proton Synchrotron executed a time of flight measurement to test the Newtonian momentum relations for light, being valid in the so-called emission theory. In this experiment, gamma rays were produced in the decay of 6-GeV pions traveling at 0.99975c. If Newtonian momentum were valid, those gamma rays should have traveled at superluminal speeds. However, they found no difference and gave an upper limit of . [27]

Energy and Calorimetry

The intrusion of particles into particle detectors is connected with electron–positron annihilation, Compton scattering, Cherenkov radiation etc., so that a cascade of effects is leading to the production of new particles (photons, electrons, neutrinos, etc.). The energy of such particle showers corresponds to the relativistic kinetic energy and rest energy of the initial particles. This energy can be measured by calorimeters in an electrical, optical, thermal, or acoustical way. [28]

Thermal measurements in order to estimate the relativistic kinetic energy were already carried out by Bertozzi as mentioned above. Additional measurements at SLAC followed, in which the heat produced by 20-GeV electrons was measured in 1982. A beam dump of water-cooled aluminium was employed as calorimeter. The results were in agreement with special relativity, even though the accuracy was only 30%. [29] However, the experimentalists alluded to the fact, that calorimetric tests with 10-GeV electrons were executed already in 1969. There, copper was used as beam dump, and an accuracy of 1% was achieved. [30]

In modern calorimeters called electromagnetic or hadronic depending on the interaction, the energy of the particle showers is often measured by the ionization caused by them. Also excitations can arise in scintillators (see scintillation), whereby light is emitted and then measured by a scintillation counter. Cherenkov radiation is measured as well. In all of those methods, the measured energy is proportional to the initial particle energy. [28]

Annihilation and pair production

Relativistic energy and momentum can also be measured by studying processes such as annihilation and pair production. [1] For instance, the rest energy of electrons and positrons is 0.51 MeV respectively. When a photon interacts with an atomic nucleus, electron-positron pairs can be generated in case the energy of the photon matches the required threshold energy, which is the combined electron-positron rest energy of 1.02 MeV. However, if the photon energy is even higher, then the exceeding energy is converted into kinetic energy of the particles. The reverse process occurs in electron-positron annihilation at low energies, in which process photons are created having the same energy as the electron-positron pair. These are direct examples of (mass–energy equivalence).

There are also many examples of conversion of relativistic kinetic energy into rest energy. In 1974, SLAC National Accelerator Laboratory accelerated electrons and positrons up to relativistic velocities, so that their relativistic energy (i.e. the sum of their rest energy and kinetic energy) is significantly increased to about 1500 MeV each. When those particles collide, other particles such as the J/ψ meson of rest energy of about 3000 MeV were produced. [31] Much higher energies were employed at the Large Electron–Positron Collider in 1989, where electrons and positrons were accelerated up to 45 GeV each, in order to produce W and Z bosons of rest energies between 80 and 91 GeV. Later, the energies were considerably increased to 200 GeV to generate pairs of W bosons. [32] Such bosons were also measured using proton-antiproton annihilation. The combined rest energy of those particles amounts to approximately 0.938 GeV each. The Super Proton Synchrotron accelerated those particle up to relativistic velocities and energies of approximately 270 GeV each, so that the center of mass energy at the collision reaches 540 GeV. Thereby, quarks and antiquarks gained the necessary energy and momentum to annihilate into W and Z bosons. [33]

Many other experiments involving the creation of a considerable amount of different particles at relativistic velocities have been (and still are) conducted in hadron colliders such as Tevatron (up to 1 TeV), the Relativistic Heavy Ion Collider (up to 200 GeV), and most recently the Large Hadron Collider (up to 7 TeV) in the course of searching for the Higgs boson.

Nuclear reactions

The relation can be tested in nuclear reactions, as the percent differences between the masses of the reactants and the products are big enough to measure; the change in total mass should account for the change in total kinetic energy. Einstein proposed such a test in the paper where he first stated the equivalence of mass and energy, mentioning the radioactive decay of radium as a possibility. [34] The first test in a nuclear reaction, however, used the absorption of an incident proton by lithium-7, which then breaks into two alpha particles. The change in mass corresponded to the change in kinetic energy to within 0.5%. [35] [36]

A particularly sensitive test was carried out in 2005 in the gamma decay of excited sulfur and silicon nuclei, in each case to the non-excited state (ground state). The masses of the excited and ground states were measured by measuring their revolution frequencies in an electromagnetic trap. The gamma rays' energies were measured by measuring their wavelengths with gamma-ray diffraction, similar to X-ray diffraction, and using the well-established relation between photon energy and wavelength. The results confirmed the predictions of relativity to a precision of 0.0000004. [37] [38]

Related Research Articles

<span class="mw-page-title-main">Beta decay</span> Type of radioactive decay

In nuclear physics, beta decay (β-decay) is a type of radioactive decay in which an atomic nucleus emits a beta particle, transforming into an isobar of that nuclide. For example, beta decay of a neutron transforms it into a proton by the emission of an electron accompanied by an antineutrino; or, conversely a proton is converted into a neutron by the emission of a positron with a neutrino in so-called positron emission. Neither the beta particle nor its associated (anti-)neutrino exist within the nucleus prior to beta decay, but are created in the decay process. By this process, unstable atoms obtain a more stable ratio of protons to neutrons. The probability of a nuclide decaying due to beta and other forms of decay is determined by its nuclear binding energy. The binding energies of all existing nuclides form what is called the nuclear band or valley of stability. For either electron or positron emission to be energetically possible, the energy release or Q value must be positive.

In physics, an electronvolt is the measure of an amount of kinetic energy gained by a single electron accelerating from rest through an electric potential difference of one volt in vacuum. When used as a unit of energy, the numerical value of 1 eV in joules is equivalent to the numerical value of the charge of an electron in coulombs. Under the 2019 redefinition of the SI base units, this sets 1 eV equal to the exact value 1.602176634×10−19 J.

<span class="mw-page-title-main">Kinetic energy</span> Energy of a moving physical body

In physics, the kinetic energy of an object is the form of energy that it possesses due to its motion.

<span class="mw-page-title-main">Positron</span> Anti-particle to the electron

The positron or antielectron is the particle with an electric charge of +1e, a spin of 1/2, and the same mass as an electron. It is the antiparticle of the electron. When a positron collides with an electron, annihilation occurs. If this collision occurs at low energies, it results in the production of two or more photons.

<span class="mw-page-title-main">Special relativity</span> Theory of interwoven space and time by Albert Einstein

In physics, the special theory of relativity, or special relativity for short, is a scientific theory of the relationship between space and time. In Albert Einstein's 1905 treatment, the theory is presented as being based on just two postulates:

  1. The laws of physics are invariant (identical) in all inertial frames of reference.
  2. The speed of light in vacuum is the same for all observers, regardless of the motion of light source or observer.
<span class="mw-page-title-main">Spacetime</span> Mathematical model combining space and time

In physics, spacetime is any mathematical model that fuses the three dimensions of space and the one dimension of time into a single four-dimensional continuum. Spacetime diagrams are useful in visualizing and understanding relativistic effects such as how different observers perceive where and when events occur.

<span class="mw-page-title-main">Compton scattering</span> Scattering of photons off charged particles

Compton scattering is the quantum theory of high frequency photons scattering following an interaction with a charged particle, usually an electron. Specifically, when the photon hits electrons, it releases loosely bound electrons from the outer valence shells of atoms or molecules.

<span class="mw-page-title-main">Electron–positron annihilation</span> Collision causing gamma ray emission

Electron–positron annihilation occurs when an electron and a positron collide. At low energies, the result of the collision is the annihilation of the electron and positron, and the creation of energetic photons:

<span class="mw-page-title-main">Synchrotron radiation</span> Electromagnetic radiation emitted by charged particles accelerated perpendicular to their velocity

Synchrotron radiation is the electromagnetic radiation emitted when relativistically charged particles are subject to an acceleration perpendicular to their velocity. It is produced artificially in some types of particle accelerators or naturally by fast electrons moving through magnetic fields. The radiation produced in this way has a characteristic polarization, and the frequencies generated can range over a large portion of the electromagnetic spectrum.

The Greisen–Zatsepin–Kuzmin limit (GZK limit or GZK cutoff) is a theoretical upper limit on the energy of cosmic ray protons traveling from other galaxies through the intergalactic medium to our galaxy. The limit is 5×1019 eV (50 EeV), or about 8 joules (the energy of a proton travelling at ≈ 99.99999999999999999998% the speed of light). The limit is set by the slowing effect of interactions of the protons with the microwave background radiation over long distances (≈ 160 million light-years). The limit is at the same order of magnitude as the upper limit for energy at which cosmic rays have experimentally been detected, although indeed some detections appear to have exceeded the limit, as noted below. For example, one extreme-energy cosmic ray, the Oh-My-God Particle, which has been found to possess a record-breaking 3.12×1020 eV (50 joules) of energy (about the same as the kinetic energy of a 95 km/h baseball).

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. All matter exhibits wave-like behavior. For example, a beam of electrons can be diffracted just like a beam of light or a water wave.

<span class="mw-page-title-main">Trouton–Noble experiment</span>

The Trouton–Noble experiment was an attempt to detect motion of the Earth through the luminiferous aether, and was conducted in 1901–1903 by Frederick Thomas Trouton and H. R. Noble. It was based on a suggestion by George FitzGerald that a charged parallel-plate capacitor moving through the aether should orient itself perpendicular to the motion. Like the earlier Michelson–Morley experiment, Trouton and Noble obtained a null result: no motion relative to the aether could be detected. This null result was reproduced, with increasing sensitivity, by Rudolf Tomaschek, Chase and Hayden in 1994. Such experimental results are now seen, consistent with special relativity, to reflect the validity of the principle of relativity and the absence of any absolute rest frame. The experiment is a test of special relativity.

<span class="mw-page-title-main">Mass–energy equivalence</span> Relativity concept expressed as E = mc²

In physics, mass–energy equivalence is the relationship between mass and energy in a system's rest frame, where the two quantities differ only by a multiplicative constant and the units of measurement. The principle is described by the physicist Albert Einstein's formula: . In a reference frame where the system is moving, its relativistic energy and relativistic mass obey the same formula.

The word "mass" has two meanings in special relativity: invariant mass is an invariant quantity which is the same for all observers in all reference frames, while the relativistic mass is dependent on the velocity of the observer. According to the concept of mass–energy equivalence, invariant mass is equivalent to rest energy, while relativistic mass is equivalent to relativistic energy.

A collider is a type of particle accelerator that brings two opposing particle beams together such that the particles collide. Colliders may either be ring accelerators or linear accelerators.

Special relativity is a physical theory that plays a fundamental role in the description of all physical phenomena, as long as gravitation is not significant. Many experiments played an important role in its development and justification. The strength of the theory lies in its unique ability to correctly predict to high precision the outcome of an extremely diverse range of experiments. Repeats of many of those experiments are still being conducted with steadily increased precision, with modern experiments focusing on effects such as at the Planck scale and in the neutrino sector. Their results are consistent with the predictions of special relativity. Collections of various tests were given by Jakob Laub, Zhang, Mattingly, Clifford Will, and Roberts/Schleif.

In physics, relativistic mechanics refers to mechanics compatible with special relativity (SR) and general relativity (GR). It provides a non-quantum mechanical description of a system of particles, or of a fluid, in cases where the velocities of moving objects are comparable to the speed of light c. As a result, classical mechanics is extended correctly to particles traveling at high velocities and energies, and provides a consistent inclusion of electromagnetism with the mechanics of particles. This was not possible in Galilean relativity, where it would be permitted for particles and light to travel at any speed, including faster than light. The foundations of relativistic mechanics are the postulates of special relativity and general relativity. The unification of SR with quantum mechanics is relativistic quantum mechanics, while attempts for that of GR is quantum gravity, an unsolved problem in physics.

<span class="mw-page-title-main">Experimental testing of time dilation</span> Tests of special relativity

Time dilation as predicted by special relativity is often verified by means of particle lifetime experiments. According to special relativity, the rate of a clock C traveling between two synchronized laboratory clocks A and B, as seen by a laboratory observer, is slowed relative to the laboratory clock rates. Since any periodic process can be considered a clock, the lifetimes of unstable particles such as muons must also be affected, so that moving muons should have a longer lifetime than resting ones. A variety of experiments confirming this effect have been performed both in the atmosphere and in particle accelerators. Another type of time dilation experiments is the group of Ives–Stilwell experiments measuring the relativistic Doppler effect.

In particle physics, a relativistic particle is an elementary particle with kinetic energy greater than or equal to its rest-mass energy given by Einstein's relation, , or specifically, of which the velocity is comparable to the speed of light .

<span class="mw-page-title-main">Proper velocity</span> Ratio in relativity

In relativity, proper velocityw of an object relative to an observer is the ratio between observer-measured displacement vector and proper time τ elapsed on the clocks of the traveling object:

References

  1. 1 2 Edwin F. Taylor; John Archibald Wheeler (1992). Spacetime Physics: Introduction to Special Relativity . New York: W. H. Freeman. ISBN   0-7167-2327-1.
  2. 1 2 Plettner, Tomas; Byer, Robert L.; Siemann, Robert H. (2005), "The impact of Einstein's theory of special relativity on particle accelerators", Journal of Physics B, 38 (9): S741–S752, Bibcode:2005JPhB...38S.741P, doi:10.1088/0953-4075/38/9/020, S2CID   53334246
  3. Zahn, C. T. & Spees, A. A. (1938), "A Critical Analysis of the Classical Experiments on the Variation of Electron Mass", Physical Review, 53 (7): 511–521, Bibcode:1938PhRv...53..511Z, doi:10.1103/PhysRev.53.511
  4. P. S. Faragó & L. Jánossy (1957), "Review of the experimental evidence for the law of variation of the electron mass with velocity", Il Nuovo Cimento, 5 (6): 379–383, Bibcode:1957NCim....5.1411F, doi:10.1007/BF02856033, S2CID   121179531
  5. Glitscher, Karl (1917). "Spektroskopischer Vergleich zwischen den Theorien des starren und des deformierbaren Elektrons". Annalen der Physik. 357 (6): 608–630. Bibcode:1917AnP...357..608G. doi:10.1002/andp.19173570603. hdl:2027/uc1.b2637473.
  6. Rogers, Marguerite M.; McReynolds, A. W.; Rogers, F. T. (1940), "A Determination of the Masses and Velocities of Three Radium B Beta-Particles: The Relativistic Mass of the Electron", Physical Review, 57 (5): 379–383, Bibcode:1940PhRv...57..379R, doi:10.1103/PhysRev.57.379, hdl:1911/18426
  7. Meyer, V.; Reichart, W.; Staub, H.H. (1963). "Experimentelle Untersuchung der Massen-Impulsrelation des Elektrons". Helvetica Physica Acta. 36: 981–992. doi:10.5169/seals-113412.
  8. Grove, D. J.; Fox, J. C. (1953). "e/m for 385-MeV protons (UA7)". Physical Review. 90 (2): 378. Bibcode:1953PhRv...90..333.. doi:10.1103/PhysRev.90.333.
  9. Zrelov, V. P.; Tiapkin, A. A.; Farago, P. S. (1958). "Measurement of the mass of 600 MeV protons". Soviet Physics JETP. 7 (3): 384–387.
  10. Bertozzi, William (1964), "Speed and Kinetic Energy of Relativistic Electrons", American Journal of Physics, 32 (7): 551–555, Bibcode:1964AmJPh..32..551B, doi:10.1119/1.1970770
  11. Bertozzi, William (1962), The Ultimate Speed - An Exploration with High Energy Electrons https://www.youtube.com/watch?v=B0BOpiMQXQA
  12. 1 2 Marvel, Robert E.; Vineyard, Michael F. (2011). "Relativistic Electron Experiment for the Undergraduate Laboratory". arXiv: 1108.5977 [physics.ed-ph].
  13. Lund, M.; Uggerhøj, U. I. (2009), "Experimental special relativity with a meter stick and a clock", American Journal of Physics, 77 (8): 757–761, Bibcode:2009AmJPh..77..757L, doi:10.1119/1.3049532
  14. Luetzelschwab, John W. (2003), "Apparatus to measure relativistic mass increase", American Journal of Physics, 71 (8): 878–884, Bibcode:2003AmJPh..71..878L, doi: 10.1119/1.1561457
  15. Couch, Jack G.; Dorries, Terry K. (1982), "Measuring relativistic electrons in the undergraduate laboratory", American Journal of Physics, 50 (10): 917–921, Bibcode:1982AmJPh..50..917C, doi:10.1119/1.12973
  16. Geller, Kenneth N.; Kollarits, Richard (1972), "Experiment to Measure the Increase in Electron Mass with Velocity", American Journal of Physics, 40 (8): 1125–1130, Bibcode:1972AmJPh..40.1125G, doi:10.1119/1.1986771
  17. Parker, Sherwood (1972), "Relativity in an Undergraduate Laboratory-Measuring the Relativistic Mass Increase", American Journal of Physics, 40 (2): 241–244, Bibcode:1972AmJPh..40..241P, doi:10.1119/1.1986498, S2CID   122167869
  18. Bartlett, A. A.; Correll, Malcolm (1965), "An Undergraduate Laboratory Apparatus for Measuring e/m as a Function of Velocity. I", American Journal of Physics, 33 (4): 327–339, Bibcode:1965AmJPh..33..327B, doi:10.1119/1.1971493
  19. Jolivette, P. L.; Rouze, N. (1994), "Compton scattering, the electron mass, and relativity: A laboratory experiment", American Journal of Physics, 62 (3): 266–271, Bibcode:1994AmJPh..62..266J, doi:10.1119/1.17611
  20. Hoffman, Matthiam J. H. (1989), "The Compton effect as an experimental approach toward relativistic mass", American Journal of Physics, 57 (9): 822–825, Bibcode:1989AmJPh..57..822H, doi:10.1119/1.15902
  21. Egelstaff, P. A.; Jackman, J. A.; Schultz, P. J.; Nickel, B. G.; MacKenzie, I. K. (1981), "Experiments in special relativity using Compton scattering of gamma rays", American Journal of Physics, 49 (1): 43–47, Bibcode:1981AmJPh..49...43E, doi:10.1119/1.12659
  22. Higbie, J. (1974), "Undergraduate Relativity Experiment", American Journal of Physics, 42 (8): 642–644, Bibcode:1974AmJPh..42..642H, doi:10.1119/1.1987800
  23. Dryzek, Jerzy; Singleton, Douglas; Suzuki, Takenori; Yu, Runsheng (2006), "An undergraduate experiment to test relativistic kinematics using in flight positron annihilation", American Journal of Physics, 74 (1): 49–53, Bibcode:2006AmJPh..74...49D, doi:10.1119/1.2142624
  24. Zhang, Yuan Zhong (1997). Special Relativity and Its Experimental Foundations . World Scientific. ISBN   978-981-02-2749-4.
  25. Brown, B. C.; Masek, G. E.; Maung, T.; Miller, E. S.; Ruderman, H.; Vernon, W. (1973), "Experimental Comparison of the Velocities of eV (Visible) and GeV Electromagnetic Radiation", Physical Review Letters, 30 (16): 763–766, Bibcode:1973PhRvL..30..763B, doi:10.1103/PhysRevLett.30.763
  26. Guiragossián, Z. G. T.; Rothbart, G. B.; Yearian, M. R.; Gearhart, R. A.; Murray, J. J. (1974), "Relative Velocity Measurements of Electrons and Gamma Rays at 15 GeV", Physical Review Letters, 34 (6): 335–338, Bibcode:1975PhRvL..34..335G, doi:10.1103/PhysRevLett.34.335, OSTI   1443188
  27. Alväger, T.; Farley, F. J. M.; Kjellman, J.; Wallin, L. (1964), "Test of the second postulate of special relativity in the GeV region", Physics Letters, 12 (3): 260–262, Bibcode:1964PhL....12..260A, doi:10.1016/0031-9163(64)91095-9.
  28. 1 2 Fabjan, Christian W.; Gianotti, Fabiola (2003). "Calorimetry for particle physics" (PDF). Reviews of Modern Physics. 75 (4): 1243–1286. Bibcode:2003RvMP...75.1243F. doi:10.1103/RevModPhys.75.1243.
  29. Walz, Dieter R.; Noyes, H. Pierre; Carezani, Ricardo L. (1984). "Calorimetric test of special relativity". Physical Review A. 29 (4): 2110–2113. Bibcode:1984PhRvA..29.2110W. doi:10.1103/PhysRevA.29.2110. OSTI   1446354.
  30. Fischer, G. E.; Murata, Y. (1970). "A beam monitor system for high-intensity photon beams in the multi-GeV range". Nuclear Instruments and Methods. 78 (1): 25–39. Bibcode:1970NucIM..78...25F. doi:10.1016/0029-554X(70)90425-8. OSTI   4752864.
  31. Burton Richter (1976). "From the Psi to Charm – The Experiments of 1975 and 1976". Nobel lecture 1976.
  32. LEP collaborations (1992), "Electroweak parameters of the Z0 resonance and the standard model", Physics Letters B, 276 (12): 247–253, Bibcode:1992PhLB..276..247., doi:10.1016/0370-2693(92)90572-L, hdl: 2066/124399
  33. Carlo Rubbia (1984). "Experimental Observation of the Intermediate Vector Bosons W+, W- and Z0". Nobel lecture 1984.
  34. Einstein, A. (1905). "Ist die Trägheit eines Körpers von seinem Energieinhalt abhängig?" [Does the Inertia of a Body Depend Upon its Energy-Content?]. Annalen der Physik (in German). 323 (13): 639–641. Bibcode:1905AnP...323..639E. doi: 10.1002/andp.19053231314 . ISSN   1521-3889.
  35. Oliphant, M. L. E.; Kinsey, B. B.; Lord Rutherford (1933). "The Transformation of Lithium by Protons and by Ions of the Heavy Isotope of Hydrogen". Proceedings of the Royal Society. 141 (845): 722–733. doi:10.1098/rspa.1935.0071.
  36. Oliphant, M. L. E.; Kempton, A. R.; Lord Rutherford (1935). "The Accurate Determination of the Energy Released in Certain Nuclear Transformations". Proceedings of the Royal Society. 149 (867): 406–416. Bibcode:1935RSPSA.149..406O. doi:10.1098/rspa.1935.0071.
  37. "Einstein Was Right (Again): Experiments Confirm that E= mc2". NIST. 2005.
  38. Rainville, S.; Thompson, J.K.; Myers, E.G.; Brown, J.M.; Dewey, M.S.; Kessler, E.G. Jr.; Deslattes, R.D.; Börner, H.G; Jentschel, M.; Mutti, P.; Pritchard, D.E. (2005-12-22). "A direct test of E = mc2". Nature. 48 (7071): 1096–1097. doi:10.1038/4381096a. PMID   16371997. S2CID   4426118.