Home  

Random  

Nearby  



Log in  



Settings  



Donate  



About Wikipedia  

Disclaimers  



Wikipedia





Hydrogen spectral series





Article  

Talk  



Language  

Watch  

Edit  


(Redirected from Paschen series)
 


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 spectral series of hydrogen, on a logarithmic scale.

Physics

edit
 
Electron transitions and their resulting wavelengths for hydrogen. Energy levels are not to scale.

A hydrogen atom consists of an electron orbiting its nucleus. The electromagnetic force between the electron and the nuclear proton leads to a set of quantum states for the electron, each with its own energy. These states were visualized by the Bohr model of the hydrogen atom as being distinct orbits around the nucleus. Each energy level, or electron shell, or orbit, is designated by an integer, n as shown in the figure. The Bohr model was later replaced by quantum mechanics in which the electron occupies an atomic orbital rather than an orbit, but the allowed energy levels of the hydrogen atom remained the same as in the earlier theory.

Spectral emission occurs when an electron transitions, or jumps, from a higher energy state to a lower energy state. To distinguish the two states, the lower energy state is commonly designated as n, and the higher energy state is designated as n. The energy of an emitted photon corresponds to the energy difference between the two states. Because the energy of each state is fixed, the energy difference between them is fixed, and the transition will always produce a photon with the same energy.

The spectral lines are grouped into series according to n. Lines are named sequentially starting from the longest wavelength/lowest frequency of the series, using Greek letters within each series. For example, the 2 → 1 line is called "Lyman-alpha" (Ly-α), while the 7 → 3 line is called "Paschen-delta" (Pa-δ).

 
Energy level diagram of electrons in hydrogen atom

There are emission lines from hydrogen that fall outside of these series, such as the 21 cm line. These emission lines correspond to much rarer atomic events such as hyperfine transitions.[1] The fine structure also results in single spectral lines appearing as two or more closely grouped thinner lines, due to relativistic corrections.[2]

In quantum mechanical theory, the discrete spectrum of atomic emission was based on the Schrödinger equation, which is mainly devoted to the study of energy spectra of hydrogen-like atoms, whereas the time-dependent equivalent Heisenberg equation is convenient when studying an atom driven by an external electromagnetic wave.[3]

In the processes of absorption or emission of photons by an atom, the conservation laws hold for the whole isolated system, such as an atom plus a photon. Therefore the motion of the electron in the process of photon absorption or emission is always accompanied by motion of the nucleus, and, because the mass of the nucleus is always finite, the energy spectra of hydrogen-like atoms must depend on the nuclear mass.[3]

Rydberg formula

edit

The energy differences between levels in the Bohr model, and hence the wavelengths of emitted or absorbed photons, is given by the Rydberg formula:[4]

 

where

  • n (often written  ) is the principal quantum number of the lower energy level,
  • n (or ) is the principal quantum number of the upper energy level, and
  •   is the Rydberg constant. (1.09677×107 m−1 for hydrogen and 1.09737×107 m−1 for heavy metals).[5][6]
  • The wavelength will always be positive because n is defined as the lower level and so is less than n. This equation is valid for all hydrogen-like species, i.e. atoms having only a single electron, and the particular case of hydrogen spectral lines is given by Z=1.

    Series

    edit

    Lyman series (n = 1)

    edit
     
    Lyman seriesofhydrogen atom spectral lines in the ultraviolet

    In the Bohr model, the Lyman series includes the lines emitted by transitions of the electron from an outer orbit of quantum number n > 1 to the 1st orbit of quantum number n' = 1.

    The series is named after its discoverer, Theodore Lyman, who discovered the spectral lines from 1906–1914. All the wavelengths in the Lyman series are in the ultraviolet band.[7][8]

    n λ, vacuum

    (nm)

    2 121.57
    3 102.57
    4 97.254
    5 94.974
    6 93.780
    91.175
    Source:[9]

    Balmer series (n = 2)

    edit
     
    The four visible hydrogen emission spectrum lines in the Balmer series. H-alpha is the red line at the right.

    The Balmer series includes the lines due to transitions from an outer orbit n > 2 to the orbit n' = 2.

    Named after Johann Balmer, who discovered the Balmer formula, an empirical equation to predict the Balmer series, in 1885. Balmer lines are historically referred to as "H-alpha", "H-beta", "H-gamma" and so on, where H is the element hydrogen.[10] Four of the Balmer lines are in the technically "visible" part of the spectrum, with wavelengths longer than 400 nm and shorter than 700 nm. Parts of the Balmer series can be seen in the solar spectrum. H-alpha is an important line used in astronomy to detect the presence of hydrogen.

    n λ, air

    (nm)

    3 656.3  
    4 486.1  
    5 434.0  
    6 410.2  
    7 397.0  
    364.6
    Source:[9]

    Paschen series (Bohr series, n = 3)

    edit

    Named after the German physicist Friedrich Paschen who first observed them in 1908. The Paschen lines all lie in the infrared band.[11] This series overlaps with the next (Brackett) series, i.e. the shortest line in the Brackett series has a wavelength that falls among the Paschen series. All subsequent series overlap.

    n λ, air

    (nm)

    4 1875
    5 1282
    6 1094
    7 1005
    8 954.6
    820.4
    Source:[9]

    Brackett series (n = 4)

    edit

    Named after the American physicist Frederick Sumner Brackett who first observed the spectral lines in 1922.[12] The spectral lines of Brackett series lie in far infrared band.

    n λ, air

    (nm)

    5 4051
    6 2625
    7 2166
    8 1944
    9 1817
    1458
    Source:[9]

    Pfund series (n = 5)

    edit

    Experimentally discovered in 1924 by August Herman Pfund.[13]

    n λ, vacuum

    (nm)

    6 7460
    7 4654
    8 3741
    9 3297
    10 3039
    2279
    Source:[14]

    Humphreys series (n = 6)

    edit

    Discovered in 1953 by American physicist Curtis J. Humphreys.[15]

    n λ, vacuum

    (μm)

    7 12.37
    8 7.503
    9 5.908
    10 5.129
    11 4.673
    3.282
    Source:[14]

    Further series (n > 6)

    edit

    Further series are unnamed, but follow the same pattern and equation as dictated by the Rydberg equation. Series are increasingly spread out and occur at increasing wavelengths. The lines are also increasingly faint, corresponding to increasingly rare atomic events. The seventh series of atomic hydrogen was first demonstrated experimentally at infrared wavelengths in 1972 by Peter Hansen and John Strong at the University of Massachusetts Amherst.[16]

    Extension to other systems

    edit

    The concepts of the Rydberg formula can be applied to any system with a single particle orbiting a nucleus, for example a He+ ion or a muonium exotic atom. The equation must be modified based on the system's Bohr radius; emissions will be of a similar character but at a different range of energies. The Pickering–Fowler series was originally attributed to an unknown form of hydrogen with half-integer transition levels by both Pickering[17][18][19] and Fowler,[20] but Bohr correctly recognised them as spectral lines arising from the He+ nucleus.[21][22][23]

    All other atoms have at least two electrons in their neutral form and the interactions between these electrons makes analysis of the spectrum by such simple methods as described here impractical. The deduction of the Rydberg formula was a major step in physics, but it was long before an extension to the spectra of other elements could be accomplished.

    See also

    edit

    References

    edit
    1. ^ "The Hydrogen 21-cm Line". Hyperphysics. Georgia State University. 2005-10-30. Retrieved 2009-03-18.
  • ^ Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-Wesley. ISBN 978-0-8053-8714-8.
  • ^ a b Andrew, A. V. (2006). "2. Schrödinger equation". Atomic spectroscopy. Introduction of theory to Hyperfine Structure. p. 274. ISBN 978-0-387-25573-6.
  • ^ Bohr, Niels (1985), "Rydberg's discovery of the spectral laws", in Kalckar, J. (ed.), N. Bohr: Collected Works, vol. 10, Amsterdam: North-Holland Publ., pp. 373–9
  • ^ Mohr, Peter J.; Taylor, Barry N.; Newell, David B. (2008). "CODATA Recommended Values of the Fundamental Physical Constants: 2006" (PDF). Reviews of Modern Physics. 80 (2): 633–730. arXiv:0801.0028. Bibcode:2008RvMP...80..633M. CiteSeerX 10.1.1.150.3858. doi:10.1103/RevModPhys.80.633.
  • ^ "Hydrogen energies and spectrum". hyperphysics.phy-astr.gsu.edu. Retrieved 2020-06-26.
  • ^ Lyman, Theodore (1906), "The Spectrum of Hydrogen in the Region of Extremely Short Wave-Length", Memoirs of the American Academy of Arts and Sciences, New Series, 23 (3): 125–146, Bibcode:1906MAAAS..13..125L, doi:10.2307/25058084, JSTOR 25058084. Also in The Astrophysical Journal, 23: 181, 1906, Bibcode:1906ApJ....23..181L, doi:10.1086/141330{{citation}}: CS1 maint: untitled periodical (link).
  • ^ Lyman, Theodore (1914), "An Extension of the Spectrum in the Extreme Ultra-Violet", Nature, 93 (2323): 241, Bibcode:1914Natur..93..241L, doi:10.1038/093241a0
  • ^ a b c d Wiese, W. L.; Fuhr, J. R. (2009), "Accurate Atomic Transition Probabilities for Hydrogen, Helium, and Lithium", Journal of Physical and Chemical Reference Data, 38 (3): 565, Bibcode:2009JPCRD..38..565W, doi:10.1063/1.3077727
  • ^ Balmer, J. J. (1885), "Notiz uber die Spectrallinien des Wasserstoffs", Annalen der Physik, 261 (5): 80–87, Bibcode:1885AnP...261...80B, doi:10.1002/andp.18852610506
  • ^ Paschen, Friedrich (1908), "Zur Kenntnis ultraroter Linienspektra. I. (Normalwellenlängen bis 27000 Å.-E.)", Annalen der Physik, 332 (13): 537–570, Bibcode:1908AnP...332..537P, doi:10.1002/andp.19083321303, archived from the original on 2012-12-17
  • ^ Brackett, Frederick Sumner (1922), "Visible and Infra-Red Radiation of Hydrogen", Astrophysical Journal, 56: 154, Bibcode:1922ApJ....56..154B, doi:10.1086/142697, hdl:2027/uc1.$b315747, S2CID 122252244
  • ^ Pfund, A. H. (1924), "The emission of nitrogen and hydrogen in infrared", J. Opt. Soc. Am., 9 (3): 193–196, Bibcode:1924JOSA....9..193P, doi:10.1364/JOSA.9.000193
  • ^ a b Kramida, A. E.; et al. (November 2010). "A critical compilation of experimental data on spectral lines and energy levels of hydrogen, deuterium, and tritium". Atomic Data and Nuclear Data Tables. 96 (6): 586–644. Bibcode:2010ADNDT..96..586K. doi:10.1016/j.adt.2010.05.001.
  • ^ Humphreys, C.J. (1953), "The Sixth Series in the Spectrum of Atomic Hydrogen", Journal of Research of the National Bureau of Standards, 50: 1, doi:10.6028/jres.050.001
  • ^ Hansen, Peter; Strong, John (1973). "Seventh Series of Atomic Hydrogen". Applied Optics. 12 (2): 429–430. Bibcode:1973ApOpt..12..429H. doi:10.1364/AO.12.000429. PMID 20125315.
  • ^ Pickering, E. C. (1896). "Stars having peculiar spectra. New variable stars in Crux and Cygnus". Harvard College Observatory Circular. 12: 1–2. Bibcode:1896HarCi..12....1P. Also published as: Pickering, E. C.; Fleming, W. P. (1896). "Stars having peculiar spectra. New variable stars in Crux and Cygnus". Astrophysical Journal. 4: 369–370. Bibcode:1896ApJ.....4..369P. doi:10.1086/140291.
  • ^ Pickering, E. C. (1897). "Stars having peculiar spectra. New variable Stars in Crux and Cygnus". Astronomische Nachrichten. 142 (6): 87–90. Bibcode:1896AN....142...87P. doi:10.1002/asna.18971420605.
  • ^ Pickering, E. C. (1897). "The spectrum of zeta Puppis". Astrophysical Journal. 5: 92–94. Bibcode:1897ApJ.....5...92P. doi:10.1086/140312.
  • ^ Fowler, A. (1912). "Observations of the Principal and other Series of Lines in the Spectrum of Hydrogen". Monthly Notices of the Royal Astronomical Society. 73 (2): 62–63. Bibcode:1912MNRAS..73...62F. doi:10.1093/mnras/73.2.62.
  • ^ Bohr, N. (1913). "The Spectra of Helium and Hydrogen". Nature. 92 (2295): 231–232. Bibcode:1913Natur..92..231B. doi:10.1038/092231d0. S2CID 11988018.
  • ^ Hoyer, Ulrich (1981). "Constitution of Atoms and Molecules". In Hoyer, Ulrich (ed.). Niels Bohr – Collected Works: Volume 2 – Work on Atomic Physics (1912–1917). Amsterdam: North Holland Publishing Company. pp. 103–316 (esp. pp. 116–122). ISBN 978-0720418002.
  • ^ Robotti, Nadia (1983). "The Spectrum of ζ Puppis and the Historical Evolution of Empirical Data". Historical Studies in the Physical Sciences. 14 (1): 123–145. doi:10.2307/27757527. JSTOR 27757527.
  • edit

      Media related to Hydrogen spectral series at Wikimedia Commons


    Retrieved from "https://en.wikipedia.org/w/index.php?title=Hydrogen_spectral_series&oldid=1227646491#Paschen_series_(Bohr_series,_n′_=_3)"
     



    Last edited on 7 June 2024, at 00:41  





    Languages

     


    العربية

    Беларуская
    Български
    Català
    Español
    Euskara
    فارسی
    Français
    Gaeilge
    Galego

    Hrvatski
    Bahasa Indonesia
    Italiano
    Nederlands

    Oʻzbekcha / ўзбекча
    Polski
    Português
    Română
    Русский
    Slovenščina
    Српски / srpski
    Srpskohrvatski / српскохрватски
    ி
    Türkçe
    Українська

     

    Wikipedia


    This page was last edited on 7 June 2024, at 00:41 (UTC).

    Content is available under CC BY-SA 4.0 unless otherwise noted.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Terms of Use

    Desktop