Fundamental series

Last updated

The fundamental series is a set of spectral lines in a set caused by transition between d and f orbitals in atoms.

Contents

Originally the series was discovered in the infrared by Fowler and independently by Arno Bergmann. [1] This resulted in the name Bergmann series used for such a set of lines in a spectrum. However the name was changed as Bergmann also discovered other series of lines. And other discoverers also established other such series. They became known as the fundamental series. [2] Bergmann observed lithium at 5347 cm−1, sodium at 5416 cm−1 potassium at 6592 cm−1. [2] Bergmann observed that the lines in the series in the caesium spectrum were double. His discovery was announced in Contributions to the Knowledge of the Infra-Red Emission Spectra of the Alkalies, Jena 1907. [3] Carl Runge called this series the "new series". He predicted that the lines of potassium and rubidium would be in pairs. [3] He expressed the frequencies of the series lines by a formula and predicted a connection of the series limit to the other known series. In 1909 W. M. Hicks produced approximate formulas for the various series and noticed that this series had a simpler formula than the others and thus called it the "fundamental series" and used the letter F. [1] [4]

The formula that more resembled the hydrogen spectrum calculations was because of a smaller quantum defect. There is no physical basis to call this fundamental. [5] The fundamental series was described as badly-named. [6] It is the last spectroscopic series to have a special designation. [6] The next series involving transitions between F and G subshells is known as the FG series. [6]

Frequencies of the lines in the series are given by this formula:

R is the Rydberg constant, is the series limit, represented by 3D, and is represented by mF. A shortened formula is then given by with values of m being integers from 4 upwards. [7] The two numbers separated by the "−" are called terms, that represent the energy level of an atom.

The limit of the fundamental series is the same as the 3D level. [5]

The terms can have different designations, mF for single line systems, mΦ for doublets and mf for triplets. [8]

Lines in the fundamental series are split into compound doublets, due to the D and F subshells having different spin possibilities. The splitting of the D subshell is very small and that of the F subshell even less so, so the fine structure in the fundamental series is harder to resolve than that in the sharp or diffuse series. [7]

Lithium

The quantum defect for lithium is 0. [5]

Sodium

Grotrian diagram for sodium. Fundamental series is due to 3D-mF transitions shown here in cyan (green). Energy levels of sodium atom.png
Grotrian diagram for sodium. Fundamental series is due to 3D-mF transitions shown here in cyan (green).

The fundamental series lines for sodium appear in the near infrared.

sodium fundamental series [9]
transitionwavelength 1 Å
3d-4f18465.3
3d-5f12679.2
3d-6f10834.9
3d-7f9961.28
3d-8f9465.94
3d-9f9153.88
3d-10f8942.96
3d-11f8796

Potassium

The fundamental series lines for potassium appear in the near infrared.

potassium fundamental series [10]
transitionwavelength 1 Åwavelength 2 Å
3d-4f15163.115168.4
3d-5f11022.3
3d-6f9565.69597.76
3d-7f8902.28902.4
3d-8f8503.58505.3
3d-9f8250.28251.7

Rubidium

The fundamental series lines for rubidium appear in the near infrared. The valence electron moves from the 4d level as the 3d is contained in an inner shell. They were observed by R von Lamb. Relevant energy levels are 4p64dj=5/2 19,355.282 cm−1 and j=3/2 19,355.623 cm−1, and the first f levels at 4p64fj=5/2 26,792.185 cm−1 and j=7/2 26,792.169 cm−1. [11]

rubidium fundamental series [11]
transitionvacuum wavelength 1 Å 5/2–7/2vacuum wavelength 2 Å 3/2-5/2
4d-4f13,446.48613,447.076
4d-5f10,078.03910,078.473
4d-6f8870.9438871.281
4d-7f8273.6848273.981
4d-8f7927.440

Caesium

caesium fundamental series [12]
transitionwavelengths Å
5/2–7/25/2-5/23/2-5/2
5d-4f10,126.376 710,126.18810,027.103 3
5d-5f8,079.036 38,078.9368,015.726 3
5d-6f7,279.956 87,279.8897,228.533 6
5d-7f6,870.454 46,824.651 3
5d-8f6,628.657 66,586.021 6
5d-9f6,472.619 66,431.966 2
5d-10f6,365.522 66,326.203 4
5d-11f6,288.592 76,250.214 9
5d-12f6,231.349 06,193.668 9
5d-13f6,187.544 26,150.38
5d-14f6,153.238 16,116.52

Related Research Articles

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

In atomic theory and quantum mechanics, an atomic orbital is a function describing the location and wave-like behavior of an electron in an atom. This function can be used to calculate the probability of finding any electron of an atom in any specific region around the atom's nucleus. The term atomic orbital may also refer to the physical region or space where the electron can be calculated to be present, as predicted by the particular mathematical form of the orbital.

<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">Infrared spectroscopy</span> Measurement of infrared radiations interaction with matter

Infrared spectroscopy is the measurement of the interaction of infrared radiation with matter by absorption, emission, or reflection. It is used to study and identify chemical substances or functional groups in solid, liquid, or gaseous forms. It can be used to characterize new materials or identify and verify known and unknown samples. The method or technique of infrared spectroscopy is conducted with an instrument called an infrared spectrometer which produces an infrared spectrum. An IR spectrum can be visualized in a graph of infrared light absorbance on the vertical axis vs. frequency, wavenumber or wavelength on the horizontal axis. Typical units of wavenumber used in IR spectra are reciprocal centimeters, with the symbol cm−1. Units of IR wavelength are commonly given in micrometers, symbol μm, which are related to the wavenumber in a reciprocal way. A common laboratory instrument that uses this technique is a Fourier transform infrared (FTIR) spectrometer. Two-dimensional IR is also possible as discussed below.

<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 Balmer series, or Balmer lines in atomic physics, is one of a set of six named series describing the spectral line emissions of the hydrogen atom. The Balmer series is calculated using the Balmer formula, an empirical equation discovered by Johann Balmer in 1885.

In physics and chemistry, the Lyman series is a hydrogen spectral series of transitions and resulting ultraviolet emission lines of the hydrogen atom as an electron goes from n ≥ 2 to n = 1, the lowest energy level of the electron. The transitions are named sequentially by Greek letters: from n = 2 to n = 1 is called Lyman-alpha, 3 to 1 is Lyman-beta, 4 to 1 is Lyman-gamma, and so on. The series is named after its discoverer, Theodore Lyman. The greater the difference in the principal quantum numbers, the higher the energy of the electromagnetic emission.

Rotational–vibrational spectroscopy is a branch of molecular spectroscopy concerned with infrared and Raman spectra of molecules in the gas phase. Transitions involving changes in both vibrational and rotational states can be abbreviated as rovibrational transitions. When such transitions emit or absorb photons, the frequency is proportional to the difference in energy levels and can be detected by certain kinds of spectroscopy. Since changes in rotational energy levels are typically much smaller than changes in vibrational energy levels, changes in rotational state are said to give fine structure to the vibrational spectrum. For a given vibrational transition, the same theoretical treatment as for pure rotational spectroscopy gives the rotational quantum numbers, energy levels, and selection rules. In linear and spherical top molecules, rotational lines are found as simple progressions at both higher and lower frequencies relative to the pure vibration frequency. In symmetric top molecules the transitions are classified as parallel when the dipole moment change is parallel to the principal axis of rotation, and perpendicular when the change is perpendicular to that axis. The ro-vibrational spectrum of the asymmetric rotor water is important because of the presence of water vapor in the atmosphere.

<span class="mw-page-title-main">Hyperfine structure</span> Small shifts and splittings in the energy levels of atoms, molecules and ions

In atomic physics, hyperfine structure is defined by small shifts in otherwise degenerate energy levels and the resulting splittings in those energy levels of atoms, molecules, and ions, due to electromagnetic multipole interaction between the nucleus and electron clouds.

<span class="mw-page-title-main">Rotational spectroscopy</span> Spectroscopy of quantized rotational states of gases

Rotational spectroscopy is concerned with the measurement of the energies of transitions between quantized rotational states of molecules in the gas phase. The rotational spectrum of polar molecules can be measured in absorption or emission by microwave spectroscopy or by far infrared spectroscopy. The rotational spectra of non-polar molecules cannot be observed by those methods, but can be observed and measured by Raman spectroscopy. Rotational spectroscopy is sometimes referred to as pure rotational spectroscopy to distinguish it from rotational-vibrational spectroscopy where changes in rotational energy occur together with changes in vibrational energy, and also from ro-vibronic spectroscopy where rotational, vibrational and electronic energy changes occur simultaneously.

In physics and chemistry, a selection rule, or transition rule, formally constrains the possible transitions of a system from one quantum state to another. Selection rules have been derived for electromagnetic transitions in molecules, in atoms, in atomic nuclei, and so on. The selection rules may differ according to the technique used to observe the transition. The selection rule also plays a role in chemical reactions, where some are formally spin-forbidden reactions, that is, reactions where the spin state changes at least once from reactants to products.

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

The Aufbau principle, also called the Aufbau rule, states that in the ground state of an atom or ion, electrons fill subshells of the lowest available energy, then they 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.

<span class="mw-page-title-main">Moseley's law</span> Law concerning X-rays emitted by atoms

Moseley's law is an empirical law concerning the characteristic X-rays emitted by atoms. The law had been discovered and published by the English physicist Henry Moseley in 1913–1914. Until Moseley's work, "atomic number" was merely an element's place in the periodic table and was not known to be associated with any measurable physical quantity. In brief, the law states that the square root of the frequency of the emitted X-ray is approximately proportional to the atomic number:

<span class="mw-page-title-main">Einstein coefficients</span> Quantities describing probability of absorption or emission of light

Einstein coefficients are quantities describing the probability of absorption or emission of a photon by an atom or molecule. The Einstein A coefficients are related to the rate of spontaneous emission of light, and the Einstein B coefficients are related to the absorption and stimulated emission of light. Throughout this article, "light" refers to any electromagnetic radiation, not necessarily in the visible spectrum.

<span class="mw-page-title-main">Hydrogen spectral series</span> Important atomic emission spectra

The emission spectrum of atomic hydrogen has been divided into a number of spectral series, with wavelengths given by the Rydberg formula. These observed spectral lines are due to the electron making transitions between two energy levels in an atom. The classification of the series by the Rydberg formula was important in the development of quantum mechanics. The spectral series are important in astronomical spectroscopy for detecting the presence of hydrogen and calculating red shifts.

The Rydberg–Ritz combination principle is an empirical rule proposed by Walther Ritz in 1908 to describe the relationship of the spectral lines for all atoms, as a generalization of an earlier rule by Johannes Rydberg for the hydrogen atom and the alkali metals. The principle states that the spectral lines of any element include frequencies that are either the sum or the difference of the frequencies of two other lines. Lines of the spectra of elements could be predicted from existing lines. Since the frequency of light is proportional to the wavenumber or reciprocal wavelength, the principle can also be expressed in terms of wavenumbers which are the sum or difference of wavenumbers of two other lines.

A molecular vibration is a periodic motion of the atoms of a molecule relative to each other, such that the center of mass of the molecule remains unchanged. The typical vibrational frequencies range from less than 1013 Hz to approximately 1014 Hz, corresponding to wavenumbers of approximately 300 to 3000 cm−1 and wavelengths of approximately 30 to 3 µm.

Vibronic spectroscopy is a branch of molecular spectroscopy concerned with vibronic transitions: the simultaneous changes in electronic and vibrational energy levels of a molecule due to the absorption or emission of a photon of the appropriate energy. In the gas phase, vibronic transitions are accompanied by changes in rotational energy also.

<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.

The sharp series is a series of spectral lines in the atomic emission spectrum caused when electrons descend from higher-energy s orbitals of an atom to the lowest available p orbital. The spectral lines include some in the visible light, and they extend into the ultraviolet. The lines get closer and closer together as the frequency increases never exceeding the series limit. The sharp series was important in the development of the understanding of electron shells and subshells in atoms. The sharp series has given the letter s to the s atomic orbital or subshell.

<span class="mw-page-title-main">Diffuse series</span> Series of lines in atomic spectra

The diffuse series is a series of spectral lines in the atomic emission spectrum caused when electrons jump between the lowest p orbital and d orbitals of an atom. The total orbital angular momentum changes between 1 and 2. The spectral lines include some in the visible light, and may extend into ultraviolet or near infrared. The lines get closer and closer together as the frequency increases never exceeding the series limit. The diffuse series was important in the development of the understanding of electron shells and subshells in atoms. The diffuse series has given the letter d to the d atomic orbital or subshell.

References

  1. 1 2 Saunders, F. A. (August 1919). "Review of Recent Work on the Series Spectra of Helium and of Hydrogen". Astrophysical Journal. 50: 151. Bibcode:1919ApJ....50..151S. doi: 10.1086/142490 .
  2. 1 2 Brand, J. C. D. (24 Nov 1995). Lines of Light. CRC Press. p. 135. ISBN   9782884491631.
  3. 1 2 Runge, Carl (December 1907). "The Spectra of the Alkalies". Astrophysical Journal. 27: 158–160. Bibcode:1908ApJ....27..158R. doi:10.1086/141538.
  4. Hicks, W. M. (9 December 1909). "A Critical Study of Spectral Series.-Part I. The Alkalies H and He". Philosophical Transactions of the Royal Society of London A. 210 (459–470): 57–111. doi: 10.1098/rsta.1911.0003 . JSTOR   90988.
  5. 1 2 3 Candler, A. C. (1937). Atomic Spectra and the Vector Model: Volume 1. Cambridge, UK: Cambridge University Press. p. 22. ISBN   9781001286228 . Retrieved 28 August 2015.
  6. 1 2 3 Fowler, A. (23 February 1928). "Spectra and atoms". Journal of the Chemical Society (Resumed): 764–780. doi:10.1039/JR9280000764. S2CID   37856069.(subscription required)
  7. 1 2 Herzberg, Gerhard (January 1944). Atomic Spectra and Atomic Structure. Courier Corporation. p.  56. ISBN   9780486601151.
  8. Saunders, F. A. (1915). "Some Recent Discoveries in Spectrum Series". Astrophysical Journal. 41: 323. Bibcode:1915ApJ....41..323S. doi:10.1086/142175.
  9. Wiese, W.; Smith, M. W.; Miles, B. M. (October 1969). Atomic Transition Probabilities Volume II Sodium Through Calcium A Critical Data Compilation (PDF). Washington: National Bureau of Standards. pp. 39–41. Archived from the original (PDF) on September 24, 2015.
  10. Wiese, W.; Smith, M. W.; Miles, B. M. (October 1969). Atomic Transition Probabilities Volume II Sodium Through Calcium A Critical Data Compilation (PDF). Washington: National Bureau of Standards. pp. 228–229. Archived from the original (PDF) on September 24, 2015.
  11. 1 2 Luna, F.R.T.; Cavalcanti, G.H.; Coutinho, L.H.; Trigueiros, A.G. (December 2002). "A compilation of wavelengths and energy levels for the spectrum of neutral rubidium (RbI)". Journal of Quantitative Spectroscopy and Radiative Transfer. 75 (5): 559–587. Bibcode:2002JQSRT..75..559L. doi:10.1016/S0022-4073(02)00030-4.(subscription required)
  12. Sansonetti, Jean E. (2009). "Wavelengths, Transition Probabilities, and Energy Levels for the Spectra of Cesium (Cs I–Cs LV)" (PDF). Journal of Physical and Chemical Reference Data. 38 (4): 768–769. Bibcode:2009JPCRD..38..761S. doi:10.1063/1.3132702. Archived from the original (PDF) on 4 March 2016. Retrieved 25 August 2015.