Triatomic hydrogen

Last updated
Triatomic hydrogen
Trihydrogen.png
Identifiers
3D model (JSmol)
  • InChI=1S/H3/c1-2-3-1
    Key: FVJJEQURJXHDEY-UHFFFAOYSA-N
  • [H]1[H][H]1
Properties
H3
Molar mass 3.024 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Triatomic hydrogen or H3 is an unstable triatomic molecule containing only hydrogen. Since this molecule contains only three atoms of hydrogen it is the simplest triatomic molecule [1] and it is relatively simple to numerically solve the quantum mechanics description of the particles. Being unstable the molecule breaks up in under a millionth of a second. Its fleeting lifetime makes it rare, but it is quite commonly formed and destroyed in the universe thanks to the commonness of the trihydrogen cation. The infrared spectrum of H3 due to vibration and rotation is very similar to that of the ion, H+
3
. In the early universe this ability to emit infrared light allowed the primordial hydrogen and helium gas to cool down so as to form stars.

Contents

Formation

The neutral molecule can be formed in a low pressure gas discharge tube. [2]

A neutral beam of H3 can be formed from a beam of H+
3
ions passing through gaseous potassium, which donates an electron to the ion, forming K+. [3] Other gaseous alkali metals, such as caesium, can also be used to donate electrons. [4] H+
3
ions can be made in a duoplasmatron where an electric discharge passed through low pressure molecular hydrogen. This causes some H2 to become H+
2
. Then H2 + H+
2
H+
3
+ H. The reaction is exothermic with an energy of 1.7 eV, so the ions produced are hot with much vibrational energy. These can cool down via collisions with cooler gas if the pressure is high enough. This is significant because strongly vibrating ions produce strongly vibrating neutral molecules when neutralised according to the Franck–Condon principle. [3]

Breakup

H3 can break up in the following ways:

[5]

Properties

The molecule can only exist in an excited state. The different excited electronic states are represented by symbols for the outer electron nLΓ with n the principal quantum number, L is the electronic angular momentum, and Γ is the electronic symmetry selected from the D3h group. Extra bracketed symbols can be attached showing vibration in the core: {s,dl} with s representing symmetrical stretch, d degenerate mode, and l vibrational angular momentum. Yet another term can be inserted to indicate molecular rotation: (N,G) with N angular momentum apart from electrons as projected on the molecular axis, and G the Hougen's convenient quantum number determined by G=l+λ-K. This is often (1,0), as the rotational states are restricted by the constituent particles all being fermions. Examples of these states are: [5] 2sA1' 3sA1' 2pA2" 3dE' 3DE" 3dA1' 3pE' 3pA2". The 2p2A2" state has a lifetime of 700 ns. If the molecule attempts to lose energy and go to the repulsive ground state, it spontaneously breaks up. The lowest energy metastable state, 2sA1' has an energy -3.777 eV below the H+
3
and e state but decays in around 1  ps. [5] The unstable ground state designated 2p2E' spontaneously breaks up into a H2 molecule and an H atom. [1] Rotationless states have a longer life time than rotating molecules. [1]

The electronic state for a trihydrogen cation with an electron delocalized around it is a Rydberg state. [6]

The outer electron can be boosted to high Rydberg state, and can ionise if the energy gets to 29562.6 cm−1 above the 2pA2" state, in which case H+
3
forms. [7]

Shape

The shape of the molecule is predicted to be an equilateral triangle. [1] Vibrations can occur in the molecule in two ways, firstly the molecule can expand and contract retaining the equilateral triangle shape (breathing), or one atom can move relative to the others distorting the triangle (bending). The bending vibration has a dipole moment and thus couples to infrared radiation. [1]

Spectrum

Gerhard Herzberg was the first to find spectroscopic lines of neutral H3 when he was 75 years old in 1979. Later he announced that this observation was one of his favourite discoveries. [8] The lines came about from a cathode discharge tube. [8] The reason that earlier observers could not see any H3 spectral lines, was due to them being swamped by the spectrum of the much more abundant H2. The important advance was to separate out H3 so it could be observed alone. Separation uses mass spectroscopy separation of the positive ions, so that H3 with mass 3 can be separated from H2 with mass 2. However there is still some contamination from HD, which also has mass 3. [3] The spectrum of H3 is mainly due to transitions to the longer lived state of 2p2A2". The spectrum can be measured via a two step photo-ionization method. [1]

Transitions dropping to the lower 2s2A1' state are affected by its very short lifetime in what is called predissociation. The spectral lines involved are broadened. [3] In the spectrum there are bands due to rotation with P Q and R branches. The R branch is very weak in H3 isotopomer but strong with D3 (trideuterium). [3]

lower stateupper electronic statebreathing vibrationbending vibrationangular momentumG=λ+l2-Kwavenumber cm−1 [1] wavelength Åfrequency THzenergy eV
2p2A2"3s2A1'00166955990500.52.069
3d2A"00172975781518.62.1446
3d2A1'00177425636531.92.1997
3p2E'11185215399555.22.2963
3p2A2"01194515141.1583.12.4116
3d2E'01195425117585.852.4229
3s2A1'10199075023.39596.82.46818
3p2E'03199945001.58599.482.47898
3d2E"10204654886.4613.5242.5373
2s2A1'3p2E'140847100422.21.746
3p2A2"band1785756005352.2
3p2A2" Q branchall superimposedband1778756225332.205

The symmetric stretch vibration mode has a wave number of 3213.1 cm−1 for the 3s2A1' level and 3168 cm−1 for 3d2E" and 3254 cm−1 for 2p2A2". [1] The bending vibrational frequencies are also quite similar to those for H+
3
. [1]

Levels

electronic statenotewavenumber cm−1 [1] frequency THzenergy eVlife ns [ citation needed ]
3d2A1'18511554.952.295112.9
3d2E"18409551.892.282411.9
3d2E'18037540.732.23639.4
3p2A2"17789533.302.205541.3 4.1
3s2A1'17600527.6382.182158.1
3p2E'13961418.541.730922.6
2p2A2"longest life99329.760.1231169700
2p2A2"predissociation00021.8
2p2E'dissociation−16674−499.87−2.06730

Cation

The related H+
3
ion
is the most prevalent molecular ion in interstellar space. It is believed to have played a crucial role in the cooling of early stars in the history of the Universe through its ability readily to absorb and emit photons. [9] One of the most important chemical reactions in interstellar space is H+
3
+ e H3 and then H2 + H. [6]

Calculations

Since the molecule is relatively simple, researchers have attempted to calculate the properties of the molecule ab-initio from quantum theory. The Hartree–Fock equations have been used. [10]

Natural occurrence

Triatomic hydrogen will be formed during the neutralization of H+
3
. This ion will be neutralised in the presence of gasses other than He or H2, as it can abstract an electron. Thus H3 is formed in the aurora in the ionosphere of Jupiter and Saturn. [11]

History

Stark's 1913 model of triatomic hydrogen Stark model of triatomic hydrogen.png
Stark's 1913 model of triatomic hydrogen

J. J. Thomson observed H+
3
while experimenting with positive rays. He believed that it was an ionised form of H3 from about 1911. He believed that H3 was a stable molecule and wrote and lectured about it. He stated that the easiest way to make it was to target potassium hydroxide with cathode rays. [8] In 1913 Johannes Stark proposed that three hydrogen nuclei and electrons could form a stable ring shape. In 1919 Niels Bohr proposed a structure with three nuclei in a straight line, with three electrons orbiting in a circle around the central nucleus. He believed that H+
3
would be unstable, but that reacting H
2
with H+ could yield neutral H3. Stanley Allen's structure was in the shape of a hexagon with alternating electrons and nuclei. [8]

In 1916 Arthur Dempster showed that H3 gas was unstable, but at the same time also confirmed that the cation existed. In 1917 Gerald Wendt and William Duane discovered that hydrogen gas subjected to alpha particles shrank in volume and thought that diatomic hydrogen was converted to triatomic. [8] After this researchers thought that active hydrogen could be the triatomic form. [8] Joseph Lévine went so far as to postulate that low pressure systems on the Earth happened due to triatomic hydrogen high in the atmosphere. [8] In 1920 Wendt and Landauer named the substance "Hyzone" in analogy to ozone and its extra reactivity over normal hydrogen. [12] Earlier Gottfried Wilhelm Osann believed he had discovered a form of hydrogen analogous to ozone which he called "Ozonwasserstoff". It was made by electrolysis of dilute sulfuric acid. In those days no one knew that ozone was triatomic so he did not announce triatomic hydrogen. [13] This was later shown to be a mixture with sulfur dioxide, and not a new form of hydrogen. [12]

In the 1930s active hydrogen was found to be hydrogen with hydrogen sulfide contamination, and scientists stopped believing in triatomic hydrogen. [8] Quantum mechanical calculations showed that neutral H3 was unstable but that ionised H+
3
could exist. [8] When the concept of isotopes came along, people such as Bohr then thought there may be an eka-hydrogen with atomic weight 3. This idea was later proven with the existence of tritium, but that was not the explanation of why molecular weight 3 was observed in mass spectrometers. [8] J. J. Thomson later believed that the molecular weight 3 molecule he observed was Hydrogen deuteride. [13] In the Orion nebula lines were observed that were attributed to nebulium which could have been the new element eka-hydrogen, especially when its atomic weight was calculated as near 3. Later this was shown to be ionised nitrogen and oxygen. [8]

Gerhard Herzberg was the first to actually observe the spectrum of neutral H3, and this triatomic molecule was the first to have a Rydberg spectrum measured where its own ground state was unstable. [1]

See also

Related Research Articles

<span class="mw-page-title-main">Hydrogen</span> Chemical element, symbol H and atomic number 1

Hydrogen is the chemical element with the symbol H and atomic number 1. Hydrogen is the lightest element. At standard conditions hydrogen is a gas of diatomic molecules having the formula H2. It is colorless, odorless, tasteless, non-toxic, and highly combustible. Hydrogen is the most abundant chemical substance in the universe, constituting roughly 75% of all normal matter. Stars such as the Sun are mainly composed of hydrogen in the plasma state. Most of the hydrogen on Earth exists in molecular forms such as water and organic compounds. For the most common isotope of hydrogen each atom has one proton, one electron, and no neutrons.

<span class="mw-page-title-main">Molecule</span> Electrically neutral group of two or more atoms

A molecule is a group of two or more atoms held together by attractive forces known as chemical bonds; depending on context, the term may or may not include ions which satisfy this criterion. In quantum physics, organic chemistry, and biochemistry, the distinction from ions is dropped and molecule is often used when referring to polyatomic ions.

<span class="mw-page-title-main">Energy level</span> Different states of quantum systems

A quantum mechanical system or particle that is bound—that is, confined spatially—can only take on certain discrete values of energy, called energy levels. This contrasts with classical particles, which can have any amount of energy. The term is commonly used for the energy levels of the electrons in atoms, ions, or molecules, which are bound by the electric field of the nucleus, but can also refer to energy levels of nuclei or vibrational or rotational energy levels in molecules. The energy spectrum of a system with such discrete energy levels is said to be quantized.

<span class="mw-page-title-main">Rydberg atom</span> Excited atomic quantum state with high principal quantum number (n)

A Rydberg atom is an excited atom with one or more electrons that have a very high principal quantum number, n. The higher the value of n, the farther the electron is from the nucleus, on average. Rydberg atoms have a number of peculiar properties including an exaggerated response to electric and magnetic fields, long decay periods and electron wavefunctions that approximate, under some conditions, classical orbits of electrons about the nuclei. The core electrons shield the outer electron from the electric field of the nucleus such that, from a distance, the electric potential looks identical to that experienced by the electron in a hydrogen atom.

<span class="mw-page-title-main">Trihydrogen cation</span> Polyatomic ion (H₃, charge +1)

The trihydrogen cation or protonated molecular hydrogen is a cation with formula H+
3
, consisting of three hydrogen nuclei (protons) sharing two electrons.

The Rydberg states of an atom or molecule are electronically excited states with energies that follow the Rydberg formula as they converge on an ionic state with an ionization energy. Although the Rydberg formula was developed to describe atomic energy levels, it has been used to describe many other systems that have electronic structure roughly similar to atomic hydrogen. In general, at sufficiently high principal quantum numbers, an excited electron-ionic core system will have the general character of a hydrogenic system and the energy levels will follow the Rydberg formula. Rydberg states have energies converging on the energy of the ion. The ionization energy threshold is the energy required to completely liberate an electron from the ionic core of an atom or molecule. In practice, a Rydberg wave packet is created by a laser pulse on a hydrogenic atom and thus populates a superposition of Rydberg states. Modern investigations using pump-probe experiments show molecular pathways – e.g. dissociation of (NO)2 – via these special states.

<span class="mw-page-title-main">Methanium</span> Ion of carbon with five hydrogens

In chemistry, methanium is a complex positive ion with formula [CH5]+ or [CH3(H2)]+, bearing a +1 electric charge. It is a superacid and one of the onium ions, indeed the simplest carbonium ion.

A heavy Rydberg system consists of a weakly bound positive and negative ion orbiting their common centre of mass. Such systems share many properties with the conventional Rydberg atom and consequently are sometimes referred to as heavy Rydberg atoms. While such a system is a type of ionically bound molecule, it should not be confused with a molecular Rydberg state, which is simply a molecule with one or more highly excited electrons.

<span class="mw-page-title-main">Helium hydride ion</span> Chemical compound

The helium hydride ion or hydridohelium(1+) ion or helonium is a cation (positively charged ion) with chemical formula HeH+. It consists of a helium atom bonded to a hydrogen atom, with one electron removed. It can also be viewed as protonated helium. It is the lightest heteronuclear ion, and is believed to be the first compound formed in the Universe after the Big Bang.

The dihydrogen cation or hydrogen molecular ion is a cation with formula H+
2
. It consists of two hydrogen nuclei (protons) sharing a single electron. It is the simplest molecular ion.

In organic chemistry, cyanopolyynes are a family of organic compounds with the chemical formula HCnN (n = 3,5,7,…) and the structural formula H−[C≡C−]nC≡N (n = 1,2,3,…). Structurally, they are polyynes with a cyano group (−C≡N) covalently bonded to one of the terminal acetylene units (H−C≡C).

In quantum mechanics, a repulsive state is an electronic state of a molecule for which there is no minimum in the potential energy. This means that the state is unstable and unbound since the potential energy smoothly decreases with the interatomic distance and the atoms repel one another. In such a state there are no discrete vibrational energy levels; instead, these levels form a continuum. This should not be confused with an excited state, which is a metastable electronic state containing a minimum in the potential energy, and may be short or long-lived.

<span class="mw-page-title-main">Chromium(I) hydride</span> Chemical compound

Chromium(I) hydride, systematically named chromium hydride, is an inorganic compound with the chemical formula (CrH)
n
. It occurs naturally in some kinds of stars where it has been detected by its spectrum. However, molecular chromium(I) hydride with the formula CrH has been isolated in solid gas matrices. The molecular hydride is very reactive. As such the compound is not well characterised, although many of its properties have been calculated via computational chemistry.

<span class="mw-page-title-main">Iron(I) hydride</span> Chemical compound

Iron(I) hydride, systematically named iron hydride and poly(hydridoiron) is a solid inorganic compound with the chemical formula (FeH)
n
(also written ([FeH])
n
or FeH). It is both thermodynamically and kinetically unstable toward decomposition at ambient temperature, and as such, little is known about its bulk properties.

Bond softening is an effect of reducing the strength of a chemical bond by strong laser fields. To make this effect significant, the strength of the electric field in the laser light has to be comparable with the electric field the bonding electron "feels" from the nuclei of the molecule. Such fields are typically in the range of 1–10 V/Å, which corresponds to laser intensities 1013–1015 W/cm2. Nowadays, these intensities are routinely achievable from table-top Ti:Sapphire lasers.

<span class="mw-page-title-main">Magnesium monohydride</span> Chemical compound

Magnesium monohydride is a molecular gas with formula MgH that exists at high temperatures, such as the atmospheres of the Sun and stars. It was originally known as magnesium hydride, although that name is now more commonly used when referring to the similar chemical magnesium dihydride.

<span class="mw-page-title-main">Helium dimer</span> Chemical compound

The helium dimer is a van der Waals molecule with formula He2 consisting of two helium atoms. This chemical is the largest diatomic molecule—a molecule consisting of two atoms bonded together. The bond that holds this dimer together is so weak that it will break if the molecule rotates, or vibrates too much. It can only exist at very low cryogenic temperatures.

Argon compounds, the chemical compounds that contain the element argon, are rarely encountered due to the inertness of the argon atom. However, compounds of argon have been detected in inert gas matrix isolation, cold gases, and plasmas, and molecular ions containing argon have been made and also detected in space. One solid interstitial compound of argon, Ar1C60 is stable at room temperature. Ar1C60 was discovered by the CSIRO.

<span class="mw-page-title-main">Argonium</span> Chemical compound

Argonium (also called the argon hydride cation, the hydridoargon(1+) ion, or protonated argon; chemical formula ArH+) is a cation combining a proton and an argon atom. It can be made in an electric discharge, and was the first noble gas molecular ion to be found in interstellar space.

Hydrogen compounds are compounds containg the element hydrogen. In these compounds, hydrogen can form in the +1 and -1 oxidation states. Hydrogen can form compounds both ionically and in covalent substances. It is a part of many organic compounds such as hydrocarbons as well as water and other organic substances. The H+ ion is often called a proton because it has one proton and no electrons, although the proton does not move freely. Brønsted–Lowry acids are capable of donating H+ ions to bases.

References

  1. 1 2 3 4 5 6 7 8 9 10 11 Lembo, L. J.; H. Helm; D. L. Huestis (1989). "Measurement of vibrational frequencies of the H3 molecule using two-step photoionization". The Journal of Chemical Physics. 90 (10): 5299. Bibcode:1989JChPh..90.5299L. doi:10.1063/1.456434. ISSN   0021-9606.
  2. Binder, J.L.; Filby, E.A.; Grubb, A.C. (1930). "Triatomic Hydrogen". Nature . 126 (3166): 11–12. Bibcode:1930Natur.126...11B. doi:10.1038/126011c0. S2CID   4142737.
  3. 1 2 3 4 5 Figger, H.; W. Ketterle; H. Walther (1989). "Spectroscopy of triatomic hydrogen". Zeitschrift für Physik D. 13 (2): 129–137. Bibcode:1989ZPhyD..13..129F. doi:10.1007/bf01398582. ISSN   0178-7683. S2CID   124478004.
  4. Laperle, Christopher M; Jennifer E Mann; Todd G Clements; Robert E Continetti (2005). "Experimentally probing the three-body predissociation dynamics of the low-lying Rydberg states of H3 and D3". Journal of Physics: Conference Series. 4 (1): 111–117. Bibcode:2005JPhCS...4..111L. doi: 10.1088/1742-6596/4/1/015 . ISSN   1742-6588.
  5. 1 2 3 Helm H. et al.:of Bound States to Continuum States in Neutral Triatomic Hydrogen. in: Dissociative Recombination, ed. S. Guberman, Kluwer Academic, Plenum Publishers, USA, 275-288 (2003) ISBN   0-306-47765-3
  6. 1 2 Tashiro, Motomichi; Shigeki Kato (2002). "Quantum dynamics study on predissociation of H3 Rydberg states: Importance of indirect mechanism". The Journal of Chemical Physics. 117 (5): 2053. Bibcode:2002JChPh.117.2053T. doi:10.1063/1.1490918. hdl: 2433/50519 . ISSN   0021-9606.
  7. Helm, Hanspeter (1988). "Measurement of the ionization potential of triatomic hydrogen". Physical Review A. 38 (7): 3425–3429. Bibcode:1988PhRvA..38.3425H. doi:10.1103/PhysRevA.38.3425. ISSN   0556-2791. PMID   9900777.
  8. 1 2 3 4 5 6 7 8 9 10 11 Kragh, Helge (2010). "The childhood of H3 and H3+". Astronomy & Geophysics. 51 (6): 6.25–6.27. Bibcode:2010A&G....51f..25K. doi: 10.1111/j.1468-4004.2010.51625.x . ISSN   1366-8781.
  9. Shelley Littin (April 11, 2012). "H3+ The Molecule that Made the Universe". Archived from the original on February 12, 2015. Retrieved 23 July 2013.{{cite web}}: CS1 maint: unfit URL (link)
  10. Defranceschi, M.; M. Suard; G. Berthier (1984). "Numerical solution of Hartree–Fock equations for a polyatomic molecule: Linear H3 in momentum space". International Journal of Quantum Chemistry. 25 (5): 863–867. doi:10.1002/qua.560250508. ISSN   0020-7608.
  11. Keiling, Andreas; Donovan, Eric; Bagenal, Fran; Karlsson, Tomas (2013-05-09). Auroral Phenomenology and Magnetospheric Processes: Earth and Other Planets. John Wiley & Sons. p. 376. ISBN   978-1-118-67153-5 . Retrieved 18 January 2014.
  12. 1 2 Wendt, Gerald L.; Landauer, Robert S. (1920). "Triatomic Hydrogen". Journal of the American Chemical Society. 42 (5): 930–946. doi:10.1021/ja01450a009.
  13. 1 2 Kragh, Helge (2011). "A Controversial Molecule: The Early History of Triatomic Hydrogen". Centaurus. 53 (4): 257–279. doi:10.1111/j.1600-0498.2011.00237.x. ISSN   0008-8994.