Relativistic quantum chemistry

Last updated

Relativistic quantum chemistry combines relativistic mechanics with quantum chemistry to calculate elemental properties and structure, especially for the heavier elements of the periodic table. A prominent example is an explanation for the color of gold: due to relativistic effects, it is not silvery like most other metals. [1]

Contents

The term relativistic effects was developed in light of the history of quantum mechanics. Initially, quantum mechanics was developed without considering the theory of relativity. [2] Relativistic effects are those discrepancies between values calculated by models that consider relativity and those that do not. [3] Relativistic effects are important for heavier elements with high atomic numbers, such as lanthanides and actinides. [4]

Relativistic effects in chemistry can be considered to be perturbations, or small corrections, to the non-relativistic theory of chemistry, which is developed from the solutions of the Schrödinger equation. These corrections affect the electrons differently depending on the electron speed compared with the speed of light. Relativistic effects are more prominent in heavy elements because only in these elements do electrons attain sufficient speeds for the elements to have properties that differ from what non-relativistic chemistry predicts. [5]

History

Beginning in 1935, Bertha Swirles described a relativistic treatment of a many-electron system, [6] despite Paul Dirac's 1929 assertion that the only imperfections remaining in quantum mechanics "give rise to difficulties only when high-speed particles are involved and are therefore of no importance in the consideration of the atomic and molecular structure and ordinary chemical reactions in which it is, indeed, usually sufficiently accurate if one neglects relativity variation of mass and velocity and assumes only Coulomb forces between the various electrons and atomic nuclei". [7]

Theoretical chemists by and large agreed with Dirac's sentiment until the 1970s, when relativistic effects were observed in heavy elements. [8] The Schrödinger equation had been developed without considering relativity in Schrödinger's 1926 article. [9] Relativistic corrections were made to the Schrödinger equation (see Klein–Gordon equation) to describe the fine structure of atomic spectra, but this development and others did not immediately trickle into the chemical community. Since atomic spectral lines were largely in the realm of physics and not in that of chemistry, most chemists were unfamiliar with relativistic quantum mechanics, and their attention was on lighter elements typical for the organic chemistry focus of the time. [10]

Dirac's opinion on the role relativistic quantum mechanics would play for chemical systems is wrong for two reasons. First, electrons in s and p atomic orbitals travel at a significant fraction of the speed of light. Second, relativistic effects give rise to indirect consequences that are especially evident for d and f atomic orbitals. [8]

Qualitative treatment

Relativistic g as a function of velocity. For a small velocity, the
E
rel
{\displaystyle E_{\text{rel}}}
(ordinate) is equal to
E
0
=
m
c
2
,
{\displaystyle E_{0}=mc^{2},}
but as
v
e
-
c
{\displaystyle v_{\text{e}}\to c}
, the
E
rel
{\displaystyle E_{\text{rel}}}
goes to infinity. Lorentz factor.svg
Relativistic γ as a function of velocity. For a small velocity, the (ordinate) is equal to but as , the goes to infinity.

One of the most important and familiar results of relativity is that the relativistic mass of the electron increases as

where are the electron rest mass, velocity of the electron, and speed of light respectively. The figure at the right illustrates this relativistic effect as a function of velocity.

This has an immediate implication on the Bohr radius (), which is given by

where is the reduced Planck constant, and α is the fine-structure constant (a relativistic correction for the Bohr model).

Bohr calculated that a 1s orbital electron of a hydrogen atom orbiting at the Bohr radius of 0.0529 nm travels at nearly 1/137 the speed of light. [11] One can extend this to a larger element with an atomic number Z by using the expression for a 1s electron, where v is its radial velocity, i.e., its instantaneous speed tangent to the radius of the atom. For gold with Z = 79, v  0.58c, so the 1s electron will be moving at 58% of the speed of light. Plugging this in for v/c in the equation for the relativistic mass, one finds that mrel = 1.22me, and in turn putting this in for the Bohr radius above one finds that the radius shrinks by 22%.

If one substitutes the "relativistic mass" into the equation for the Bohr radius it can be written

Ratio of relativistic and nonrelativistic Bohr radii, as a function of electron velocity Bohrradiusfunctionofelectronvelocity.png
Ratio of relativistic and nonrelativistic Bohr radii, as a function of electron velocity

It follows that

At right, the above ratio of the relativistic and nonrelativistic Bohr radii has been plotted as a function of the electron velocity. Notice how the relativistic model shows the radius decreases with increasing velocity.

When the Bohr treatment is extended to hydrogenic atoms, the Bohr radius becomes

where is the principal quantum number, and Z is an integer for the atomic number. In the Bohr model, the angular momentum is given as . Substituting into the equation above and solving for gives

From this point, atomic units can be used to simplify the expression into

Substituting this into the expression for the Bohr ratio mentioned above gives

At this point one can see that a low value of and a high value of results in . This fits with intuition: electrons with lower principal quantum numbers will have a higher probability density of being nearer to the nucleus. A nucleus with a large charge will cause an electron to have a high velocity. A higher electron velocity means an increased electron relativistic mass, and as a result the electrons will be near the nucleus more of the time and thereby contract the radius for small principal quantum numbers. [12]

Periodic table deviations

The periodic table was constructed by scientists who noticed periodic trends in known elements of the time. Indeed, the patterns found in it are what give the periodic table its power. Many of the chemical and physical differences between the 5th period (RbXe) and the 6th period (CsRn) arise from the larger relativistic effects of the latter. These relativistic effects are particularly large for gold and its neighbors  platinum and mercury. An important quantum relativistic effect is the van der Waals force.[ citation needed ]

Mercury

Mercury (Hg) is a liquid down to approximately −39  °C, its melting point. Bonding forces are weaker for Hg–Hg bonds than for their immediate neighbors such as cadmium (m.p. 321 °C) and gold (m.p. 1064 °C). The lanthanide contraction only partially accounts for this anomaly. [11] Mercury in the gas phase is alone among metals in that it is quite typically found in a monomeric form as Hg(g).[ citation needed ] Hg22+(g) also forms, and it is a stable species due to the relativistic shortening of the bond.[ citation needed ]

Hg2(g) does not form because the 6s2 orbital is contracted by relativistic effects and may therefore only weakly contribute to any bonding; in fact, Hg–Hg bonding must be mostly the result of van der Waals forces, which explains why the bonding for Hg–Hg is weak enough to allow for Hg to be a liquid at room temperature. [11] [13] [14] [15]

Au2(g) and Hg(g) are analogous with H2(g) and He(g) with regard to having the same nature of difference. The relativistic contraction of the 6s2 orbital leads to gaseous mercury sometimes being referred to as a pseudo noble gas. [11]

Color of gold and caesium

Spectral reflectance curves for aluminum (Al), silver (Ag), and gold (Au) metal mirrors Image-Metal-reflectance.png
Spectral reflectance curves for aluminum (Al), silver (Ag), and gold (Au) metal mirrors
Alkali-metal coloration: rubidium (silvery) versus caesium (golden) Rb&Cs crystals.jpg
Alkali-metal coloration: rubidium (silvery) versus caesium (golden)

The reflectivity of aluminium (Al), silver (Ag), and gold (Au) is shown in the graph to the right. The human eye sees electromagnetic radiation with a wavelength near 600 nm as yellow. Gold absorbs blue light more than it absorbs other visible wavelengths of light; the reflected light reaching the eye is therefore lacking in blue compared with the incident light. Since yellow is complementary to blue, this makes a piece of gold under white light appear yellow to human eyes.

The electronic transition from the 5d orbital to the 6s orbital is responsible for this absorption. An analogous transition occurs in silver, but the relativistic effects are smaller than in gold. While silver's 4d orbital experiences some relativistic expansion and the 5s orbital contraction, the 4d–5s distance in silver is much greater than the 5d–6s distance in gold. The relativistic effects increase the 5d orbital's distance from the atom's nucleus and decrease the 6s orbital's distance. Due to the decreased 6s orbital distance, the electronic transition primarily absorbs in the violet/blue region of the visible spectrum, as opposed to the UV region. [16]

Caesium, the heaviest of the alkali metals that can be collected in quantities sufficient for viewing, has a golden hue, whereas the other alkali metals are silver-white. However, relativistic effects are not very significant at Z = 55 for caesium (not far from Z = 47 for silver). The golden color of caesium comes from the decreasing frequency of light required to excite electrons of the alkali metals as the group is descended. For lithium through rubidium, this frequency is in the ultraviolet, but for caesium it reaches the blue-violet end of the visible spectrum; in other words, the plasmonic frequency of the alkali metals becomes lower from lithium to caesium. Thus caesium transmits and partially absorbs violet light preferentially, while other colors (having lower frequency) are reflected; hence it appears yellowish. [17]

Lead–acid battery

Without relativity, lead (Z = 82) would be expected to behave much like tin (Z = 50), so tin–acid batteries should work just as well as the lead–acid batteries commonly used in cars. However, calculations show that about 10 V of the 12 V produced by a 6-cell lead–acid battery arises purely from relativistic effects, explaining why tin–acid batteries do not work. [18]

Inert-pair effect

In Tl(I) (thallium), Pb(II) (lead), and Bi(III) (bismuth) complexes a 6s2 electron pair exists. The inert pair effect is the tendency of this pair of electrons to resist oxidation due to a relativistic contraction of the 6s orbital. [8]

Other effects

Additional phenomena commonly caused by relativistic effects are the following:

See also

Related Research Articles

<span class="mw-page-title-main">Atomic orbital</span> Function describing an electron in an atom

In quantum mechanics, an atomic orbital is a function describing the location and wave-like behavior of an electron in an atom. This function describes the electron's charge distribution around the atom's nucleus, and can be used to calculate the probability of finding an electron in a specific region around the nucleus.

<span class="mw-page-title-main">Bohr model</span> Atomic model introduced by Niels Bohr in 1913

In atomic physics, the Bohr model or Rutherford–Bohr model of the atom, presented by Niels Bohr and Ernest Rutherford in 1913, consists of a small, dense nucleus surrounded by orbiting electrons. It is analogous to the structure of the Solar System, but with attraction provided by electrostatic force rather than gravity, and with the electron energies quantized.

<span class="mw-page-title-main">Hydrogen atom</span> Atom of the element hydrogen

A hydrogen atom is an atom of the chemical element hydrogen. The electrically neutral atom contains a single positively charged proton and a single negatively charged electron bound to the nucleus by the Coulomb force. Atomic hydrogen constitutes about 75% of the baryonic mass of the universe.

<span class="mw-page-title-main">Ionization energy</span> Energy needed to remove an electron

In physics and chemistry, ionization energy (IE) is the minimum energy required to remove the most loosely bound electron of an isolated gaseous atom, positive ion, or molecule. The first ionization energy is quantitatively expressed as

The Bohr radius is a physical constant, approximately equal to the most probable distance between the nucleus and the electron in a hydrogen atom in its ground state. It is named after Niels Bohr, due to its role in the Bohr model of an atom. Its value is 5.29177210903(80)×10−11 m.

The atomic units are a system of natural units of measurement that is especially convenient for calculations in atomic physics and related scientific fields, such as computational chemistry and atomic spectroscopy. They were originally suggested and named by the physicist Douglas Hartree. Atomic units are often abbreviated "a.u." or "au", not to be confused with similar abbreviations used for astronomical units, arbitrary units, and absorbance units in other contexts.

In spectroscopy, the Rydberg constant, symbol for heavy atoms or for hydrogen, named after the Swedish physicist Johannes Rydberg, is a physical constant relating to the electromagnetic spectra of an atom. The constant first arose as an empirical fitting parameter in the Rydberg formula for the hydrogen spectral series, but Niels Bohr later showed that its value could be calculated from more fundamental constants according to his model of the atom.

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. At all scales where measurements have been practical, 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">Rydberg formula</span> Formula for spectral line wavelengths in alkali metals

In atomic physics, the Rydberg formula calculates the wavelengths of a spectral line in many chemical elements. The formula was primarily presented as a generalization of the Balmer series for all atomic electron transitions of hydrogen. It was first empirically stated in 1888 by the Swedish physicist Johannes Rydberg, then theoretically by Niels Bohr in 1913, who used a primitive form of quantum mechanics. The formula directly generalizes the equations used to calculate the wavelengths of the hydrogen spectral series.

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.

In quantum mechanics, the principal quantum number is one of four quantum numbers assigned to each electron in an atom to describe that electron's state. Its values are natural numbers making it a discrete variable.

<span class="mw-page-title-main">Fine structure</span> Details in the emission spectrum of an atom

In atomic physics, the fine structure describes the splitting of the spectral lines of atoms due to electron spin and relativistic corrections to the non-relativistic Schrödinger equation. It was first measured precisely for the hydrogen atom by Albert A. Michelson and Edward W. Morley in 1887, laying the basis for the theoretical treatment by Arnold Sommerfeld, introducing the fine-structure constant.

The Compton wavelength is a quantum mechanical property of a particle, defined as the wavelength of a photon the energy of which is the same as the rest energy of that particle. It was introduced by Arthur Compton in 1923 in his explanation of the scattering of photons by electrons.

<span class="mw-page-title-main">Aufbau principle</span> Principle of atomic physics

In atomic physics and quantum chemistry, the Aufbau principle, also called the Aufbau rule, states that in the ground state of an atom or ion, electrons first fill subshells of the lowest available energy, then fill subshells of higher energy. For example, the 1s subshell is filled before the 2s subshell is occupied. In this way, the electrons of an atom or ion form the most stable electron configuration possible. An example is the configuration 1s2 2s2 2p6 3s2 3p3 for the phosphorus atom, meaning that the 1s subshell has 2 electrons, and so on.

In atomic physics, the electron magnetic moment, or more specifically the electron magnetic dipole moment, is the magnetic moment of an electron resulting from its intrinsic properties of spin and electric charge. The value of the electron magnetic moment is −9.2847647043(28)×10−24 J⋅T−1. In units of the Bohr magneton (μB), it is −1.00115965218059(13) μB, a value that was measured with a relative accuracy of 1.3×10−13.

<span class="mw-page-title-main">Orbital motion (quantum)</span> Quantum mechanical property

Quantum orbital motion involves the quantum mechanical motion of rigid particles about some other mass, or about themselves. In classical mechanics, an object's orbital motion is characterized by its orbital angular momentum and spin angular momentum, which is the object's angular momentum about its own center of mass. In quantum mechanics there are analogous orbital and spin angular momenta which describe the orbital motion of a particle, represented as quantum mechanical operators instead of vectors.

In quantum physics, the spin–orbit interaction is a relativistic interaction of a particle's spin with its motion inside a potential. A key example of this phenomenon is the spin–orbit interaction leading to shifts in an electron's atomic energy levels, due to electromagnetic interaction between the electron's magnetic dipole, its orbital motion, and the electrostatic field of the positively charged nucleus. This phenomenon is detectable as a splitting of spectral lines, which can be thought of as a Zeeman effect product of two relativistic effects: the apparent magnetic field seen from the electron perspective and the magnetic moment of the electron associated with its intrinsic spin. A similar effect, due to the relationship between angular momentum and the strong nuclear force, occurs for protons and neutrons moving inside the nucleus, leading to a shift in their energy levels in the nucleus shell model. In the field of spintronics, spin–orbit effects for electrons in semiconductors and other materials are explored for technological applications. The spin–orbit interaction is at the origin of magnetocrystalline anisotropy and the spin Hall effect.

Core electrons are the electrons in an atom that are not valence electrons and do not participate in chemical bonding. The nucleus and the core electrons of an atom form the atomic core. Core electrons are tightly bound to the nucleus. Therefore, unlike valence electrons, core electrons play a secondary role in chemical bonding and reactions by screening the positive charge of the atomic nucleus from the valence electrons.

A normalized 1s Slater-type function is a function which is used in the descriptions of atoms and in a broader way in the description of atoms in molecules. It is particularly important as the accurate quantum theory description of the smallest free atom, hydrogen. It has the form

A hydrogen-like atom (or hydrogenic atom) is any atom or ion with a single valence electron. These atoms are isoelectronic with hydrogen. Examples of hydrogen-like atoms include, but are not limited to, hydrogen itself, all alkali metals such as Rb and Cs, singly ionized alkaline earth metals such as Ca+ and Sr+ and other ions such as He+, Li2+, and Be3+ and isotopes of any of the above. A hydrogen-like atom includes a positively charged core consisting of the atomic nucleus and any core electrons as well as a single valence electron. Because helium is common in the universe, the spectroscopy of singly ionized helium is important in EUV astronomy, for example, of DO white dwarf stars.

References

  1. Pekka Pyykkö (January 2012). "Relativistic Effects in Chemistry: More Common Than You Thought". Annual Review of Physical Chemistry. 63 (1): 45–64. Bibcode:2012ARPC...63...45P. doi:10.1146/annurev-physchem-032511-143755. PMID   22404585.
  2. Kleppner, Daniel (1999). "A short history of atomic physics in the twentieth century" (PDF). Reviews of Modern Physics. 71 (2): S78–S84. Bibcode:1999RvMPS..71...78K. doi:10.1103/RevModPhys.71.S78. Archived from the original (PDF) on 2016-03-03. Retrieved 2012-07-17.
  3. Kaldor, U.; Wilson, Stephen (2003). Theoretical Chemistry and Physics of Heavy and Superheavy Elements. Dordrecht, Netherlands: Kluwer Academic Publishers. p. 4. ISBN   978-1-4020-1371-3.
  4. Kaldor, U.; Wilson, Stephen (2003). Theoretical Chemistry and Physics of Heavy and Superheavy Elements. Dordrecht, Netherlands: Kluwer Academic Publishers. p. 2. ISBN   978-1-4020-1371-3.
  5. Gupta, V. P. (22 October 2015). Principles and Applications of Quantum Chemistry. Elsevier Science. ISBN   978-0-12-803478-1 . Retrieved 2024-01-07.
  6. Swirles, B. (1935). "The Relativistic Self-Consistent Field". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 152 (877): 625–649. Bibcode:1935RSPSA.152..625S. doi: 10.1098/rspa.1935.0211 .
  7. Dirac, P. A. M. (1929). "Quantum Mechanics of Many-Electron Systems". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 123 (792): 714–733. Bibcode:1929RSPSA.123..714D. doi: 10.1098/rspa.1929.0094 . JSTOR   95222.
  8. 1 2 3 Pyykkö, Pekka (1988). "Relativistic effects in structural chemistry". Chemical Reviews. 88 (3): 563–594. doi:10.1021/cr00085a006.
  9. Erwin Schrödinger (1926). "Über das Verhältnis der Heisenberg-Born-Jordanschen Quantenmechanik zu der meinem" (PDF). Annalen der Physik (in German). 384 (8). Leipzig: 734–756. Bibcode:1926AnP...384..734S. doi:10.1002/andp.19263840804. Archived from the original (PDF) on 2008-12-17.
  10. Kaldor, U.; Wilson, Stephen, eds. (2003). Theoretical Chemistry and Physics of Heavy and Superheavy Elements. Dordrecht, Netherlands: Kluwer Academic Publishers. p. 17. ISBN   978-1-4020-1371-3.
  11. 1 2 3 4 Norrby, Lars J. (1991). "Why is mercury liquid? Or, why do relativistic effects not get into chemistry textbooks?". Journal of Chemical Education. 68 (2): 110. Bibcode:1991JChEd..68..110N. doi:10.1021/ed068p110.
  12. Pitzer, Kenneth S. (1979). "Relativistic effects on chemical properties" (PDF). Accounts of Chemical Research. 12 (8): 271–276. doi:10.1021/ar50140a001. S2CID   95601322.
  13. Calvo, Florent; Pahl, Elke; Wormit, Michael; Schwerdtfeger, Peter (2013). "Evidence for Low-Temperature Melting of Mercury owing to Relativity". Angewandte Chemie International Edition. 52 (29): 7583–7585. doi:10.1002/anie.201302742. ISSN   1521-3773. PMID   23780699.
  14. Howes2013-06-21T00:00:00+01:00, Laura. "Relativity behind mercury's liquidity". Chemistry World. Retrieved 2022-01-21.{{cite web}}: CS1 maint: numeric names: authors list (link)
  15. Jogalekar, Ashutosh. "What does mercury being liquid at room temperature have to do with Einstein's theory of relativity?". Scientific American Blog Network. Retrieved 2022-01-21.
  16. Pyykkö, Pekka; Desclaux, Jean Paul (1979). "Relativity and the periodic system of elements". Accounts of Chemical Research. 12 (8): 276. doi:10.1021/ar50140a002.
  17. Addison, C. C. (1984). The chemistry of the liquid alkali metals. Wiley. p. 7. ISBN   9780471905080.
  18. Ahuja, Rajeev; Blomqvist, Anders; Larsson, Peter; Pyykkö, Pekka; Zaleski-Ejgierd, Patryk (2011). "Relativity and the Lead-Acid Battery". Physical Review Letters. 106 (1): 018301. arXiv: 1008.4872 . Bibcode:2011PhRvL.106a8301A. doi:10.1103/PhysRevLett.106.018301. PMID   21231773. S2CID   39265906.
  19. Runeberg, Nino; Schütz, Martin; Werner, Hans-Joachim (1999). "The aurophilic attraction as interpreted by local correlation methods". J. Chem. Phys. 110 (15): 7210–7215. Bibcode:1999JChPh.110.7210R. doi:10.1063/1.478665.
  20. Wan, Qingyun; Yang, Jun; To, Wai-Pong; Che, Chi-Ming (2021-01-05). "Strong metal–metal Pauli repulsion leads to repulsive metallophilicity in closed-shell d 8 and d 10 organometallic complexes". Proceedings of the National Academy of Sciences. 118 (1): e2019265118. Bibcode:2021PNAS..11819265W. doi: 10.1073/pnas.2019265118 . ISSN   0027-8424. PMC   7817198 . PMID   33372160.
  21. Jansen, Martin (December 2005). "Effects of relativistic motion of electrons on the chemistry of gold and platinum". Solid State Sciences. 7 (12): 1464–1474. Bibcode:2005SSSci...7.1464J. doi: 10.1016/j.solidstatesciences.2005.06.015 .
  22. "IYPT 2019 Elements 087: Francium: Not the most reactive group 1 element". Explorations of everyday chemical compounds. November 6, 2019.
  23. Andreev, S. V.; Letokhov, V. S.; Mishin, V. I. (21 September 1987). "Laser resonance photoionization spectroscopy of Rydberg levels in Fr". Physical Review Letters. 59 (12): 1274–1276. Bibcode:1987PhRvL..59.1274A. doi:10.1103/PhysRevLett.59.1274. PMID   10035190.
  24. Pyykko, Pekka (1 May 1988). "Relativistic effects in structural chemistry". Chemical Reviews. 88 (3): 563–594. doi:10.1021/cr00085a006.

Further reading