fbpx
Wikipedia

Rotational spectroscopy

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 (power spectral density vs. rotational frequency) of polar molecules can be measured in absorption or emission by microwave spectroscopy[1] 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 (or just vibronic spectroscopy) where rotational, vibrational and electronic energy changes occur simultaneously.

Part of the rotational spectrum of trifluoroiodomethane, CF
3
I
.[notes 1] Each rotational transition is labeled with the quantum numbers, J, of the final and initial states, and is extensively split by the effects of nuclear quadrupole coupling with the 127I nucleus.

For rotational spectroscopy, molecules are classified according to symmetry into a spherical top, linear and symmetric top; analytical expressions can be derived for the rotational energy terms of these molecules. Analytical expressions can be derived for the fourth category, asymmetric top, for rotational levels up to J=3, but higher energy levels need to be determined using numerical methods. The rotational energies are derived theoretically by considering the molecules to be rigid rotors and then applying extra terms to account for centrifugal distortion, fine structure, hyperfine structure and Coriolis coupling. Fitting the spectra to the theoretical expressions gives numerical values of the angular moments of inertia from which very precise values of molecular bond lengths and angles can be derived in favorable cases. In the presence of an electrostatic field there is Stark splitting which allows molecular electric dipole moments to be determined.

An important application of rotational spectroscopy is in exploration of the chemical composition of the interstellar medium using radio telescopes.

Applications edit

Rotational spectroscopy has primarily been used to investigate fundamental aspects of molecular physics. It is a uniquely precise tool for the determination of molecular structure in gas-phase molecules. It can be used to establish barriers to internal rotation such as that associated with the rotation of the CH
3
group relative to the C
6
H
4
Cl
group in chlorotoluene (C
7
H
7
Cl
).[2] When fine or hyperfine structure can be observed, the technique also provides information on the electronic structures of molecules. Much of current understanding of the nature of weak molecular interactions such as van der Waals, hydrogen and halogen bonds has been established through rotational spectroscopy. In connection with radio astronomy, the technique has a key role in exploration of the chemical composition of the interstellar medium. Microwave transitions are measured in the laboratory and matched to emissions from the interstellar medium using a radio telescope. NH
3
was the first stable polyatomic molecule to be identified in the interstellar medium.[3] The measurement of chlorine monoxide[4] is important for atmospheric chemistry. Current projects in astrochemistry involve both laboratory microwave spectroscopy and observations made using modern radiotelescopes such as the Atacama Large Millimeter/submillimeter Array (ALMA).[5]

Overview edit

A molecule in the gas phase is free to rotate relative to a set of mutually orthogonal axes of fixed orientation in space, centered on the center of mass of the molecule. Free rotation is not possible for molecules in liquid or solid phases due to the presence of intermolecular forces. Rotation about each unique axis is associated with a set of quantized energy levels dependent on the moment of inertia about that axis and a quantum number. Thus, for linear molecules the energy levels are described by a single moment of inertia and a single quantum number,  , which defines the magnitude of the rotational angular momentum.

For nonlinear molecules which are symmetric rotors (or symmetric tops - see next section), there are two moments of inertia and the energy also depends on a second rotational quantum number,  , which defines the vector component of rotational angular momentum along the principal symmetry axis.[6] Analysis of spectroscopic data with the expressions detailed below results in quantitative determination of the value(s) of the moment(s) of inertia. From these precise values of the molecular structure and dimensions may be obtained.

For a linear molecule, analysis of the rotational spectrum provides values for the rotational constant[notes 2] and the moment of inertia of the molecule, and, knowing the atomic masses, can be used to determine the bond length directly. For diatomic molecules this process is straightforward. For linear molecules with more than two atoms it is necessary to measure the spectra of two or more isotopologues, such as 16O12C32S and 16O12C34S. This allows a set of simultaneous equations to be set up and solved for the bond lengths).[notes 3] A bond length obtained in this way is slightly different from the equilibrium bond length. This is because there is zero-point energy in the vibrational ground state, to which the rotational states refer, whereas the equilibrium bond length is at the minimum in the potential energy curve. The relation between the rotational constants is given by

 

where v is a vibrational quantum number and α is a vibration-rotation interaction constant which can be calculated if the B values for two different vibrational states can be found.[7]

For other molecules, if the spectra can be resolved and individual transitions assigned both bond lengths and bond angles can be deduced. When this is not possible, as with most asymmetric tops, all that can be done is to fit the spectra to three moments of inertia calculated from an assumed molecular structure. By varying the molecular structure the fit can be improved, giving a qualitative estimate of the structure. Isotopic substitution is invaluable when using this approach to the determination of molecular structure.

Classification of molecular rotors edit

In quantum mechanics the free rotation of a molecule is quantized, so that the rotational energy and the angular momentum can take only certain fixed values, which are related simply to the moment of inertia,  , of the molecule. For any molecule, there are three moments of inertia:  ,   and   about three mutually orthogonal axes A, B, and C with the origin at the center of mass of the system. The general convention, used in this article, is to define the axes such that  , with axis   corresponding to the smallest moment of inertia. Some authors, however, define the   axis as the molecular rotation axis of highest order.

The particular pattern of energy levels (and, hence, of transitions in the rotational spectrum) for a molecule is determined by its symmetry. A convenient way to look at the molecules is to divide them into four different classes, based on the symmetry of their structure. These are

Spherical tops (spherical rotors)
All three moments of inertia are equal to each other:  . Examples of spherical tops include phosphorus tetramer (P
4
)
, carbon tetrachloride (CCl
4
)
and other tetrahalides, methane (CH
4
)
, silane, (SiH
4
)
, sulfur hexafluoride (SF
6
)
and other hexahalides. The molecules all belong to the cubic point groups Td or Oh.
Linear molecules
For a linear molecule the moments of inertia are related by  . For most purposes,   can be taken to be zero. Examples of linear molecules include dioxygen (O
2
)
, dinitrogen (N
2
)
, carbon monoxide (CO), hydroxy radical (OH), carbon dioxide (CO2), hydrogen cyanide (HCN), carbonyl sulfide (OCS), acetylene (ethyne (HC≡CH) and dihaloethynes. These molecules belong to the point groups C∞v or D∞h.
Symmetric tops (symmetric rotors)
A symmetric top is a molecule in which two moments of inertia are the same,   or  . By definition a symmetric top must have a 3-fold or higher order rotation axis. As a matter of convenience, spectroscopists divide molecules into two classes of symmetric tops, Oblate symmetric tops (saucer or disc shaped) with   and Prolate symmetric tops (rugby football, or cigar shaped) with  . The spectra look rather different, and are instantly recognizable. Examples of symmetric tops include
Oblate
Benzene, C
6
H
6
; ammonia, NH
3
; xenon tetrafluoride, XeF
4
Prolate
Chloromethane, CH
3
Cl
, propyne, CH
3
C≡CH
As a detailed example, ammonia has a moment of inertia IC = 4.4128 × 10−47 kg m2 about the 3-fold rotation axis, and moments IA = IB = 2.8059 × 10−47 kg m2 about any axis perpendicular to the C3 axis. Since the unique moment of inertia is larger than the other two, the molecule is an oblate symmetric top.[8]
Asymmetric tops (asymmetric rotors)
The three moments of inertia have different values. Examples of small molecules that are asymmetric tops include water, H
2
O
and nitrogen dioxide, NO
2
whose symmetry axis of highest order is a 2-fold rotation axis. Most large molecules are asymmetric tops.

Selection rules edit

Microwave and far-infrared spectra edit

Transitions between rotational states can be observed in molecules with a permanent electric dipole moment.[9][notes 4] A consequence of this rule is that no microwave spectrum can be observed for centrosymmetric linear molecules such as N
2
(dinitrogen) or HCCH (ethyne), which are non-polar. Tetrahedral molecules such as CH
4
(methane), which have both a zero dipole moment and isotropic polarizability, would not have a pure rotation spectrum but for the effect of centrifugal distortion; when the molecule rotates about a 3-fold symmetry axis a small dipole moment is created, allowing a weak rotation spectrum to be observed by microwave spectroscopy.[10]

With symmetric tops, the selection rule for electric-dipole-allowed pure rotation transitions is ΔK = 0, ΔJ = ±1. Since these transitions are due to absorption (or emission) of a single photon with a spin of one, conservation of angular momentum implies that the molecular angular momentum can change by at most one unit.[11] Moreover, the quantum number K is limited to have values between and including +J to -J.[12]

Raman spectra edit

For Raman spectra the molecules undergo transitions in which an incident photon is absorbed and another scattered photon is emitted. The general selection rule for such a transition to be allowed is that the molecular polarizability must be anisotropic, which means that it is not the same in all directions.[13] Polarizability is a 3-dimensional tensor that can be represented as an ellipsoid. The polarizability ellipsoid of spherical top molecules is in fact spherical so those molecules show no rotational Raman spectrum. For all other molecules both Stokes and anti-Stokes lines[notes 5] can be observed and they have similar intensities due to the fact that many rotational states are thermally populated. The selection rule for linear molecules is ΔJ = 0, ±2. The reason for the values ±2 is that the polarizability returns to the same value twice during a rotation.[14] The value ΔJ = 0 does not correspond to a molecular transition but rather to Rayleigh scattering in which the incident photon merely changes direction.[15]

The selection rule for symmetric top molecules is

ΔK = 0
If K = 0, then ΔJ = ±2
If K ≠ 0, then ΔJ = 0, ±1, ±2

Transitions with ΔJ = +1 are said to belong to the R series, whereas transitions with ΔJ = +2 belong to an S series.[15] Since Raman transitions involve two photons, it is possible for the molecular angular momentum to change by two units.

Units edit

The units used for rotational constants depend on the type of measurement. With infrared spectra in the wavenumber scale ( ), the unit is usually the inverse centimeter, written as cm−1, which is literally the number of waves in one centimeter, or the reciprocal of the wavelength in centimeters ( ). On the other hand, for microwave spectra in the frequency scale ( ), the unit is usually the gigahertz. The relationship between these two units is derived from the expression

 

where ν is a frequency, λ is a wavelength and c is the velocity of light. It follows that

 

As 1 GHz = 109 Hz, the numerical conversion can be expressed as

 

Effect of vibration on rotation edit

The population of vibrationally excited states follows a Boltzmann distribution, so low-frequency vibrational states are appreciably populated even at room temperatures. As the moment of inertia is higher when a vibration is excited, the rotational constants (B) decrease. Consequently, the rotation frequencies in each vibration state are different from each other. This can give rise to "satellite" lines in the rotational spectrum. An example is provided by cyanodiacetylene, H−C≡C−C≡C−C≡N.[16]

Further, there is a fictitious force, Coriolis coupling, between the vibrational motion of the nuclei in the rotating (non-inertial) frame. However, as long as the vibrational quantum number does not change (i.e., the molecule is in only one state of vibration), the effect of vibration on rotation is not important, because the time for vibration is much shorter than the time required for rotation. The Coriolis coupling is often negligible, too, if one is interested in low vibrational and rotational quantum numbers only.

Effect of rotation on vibrational spectra edit

Historically, the theory of rotational energy levels was developed to account for observations of vibration-rotation spectra of gases in infrared spectroscopy, which was used before microwave spectroscopy had become practical. To a first approximation, the rotation and vibration can be treated as separable, so the energy of rotation is added to the energy of vibration. For example, the rotational energy levels for linear molecules (in the rigid-rotor approximation) are

 

In this approximation, the vibration-rotation wavenumbers of transitions are

 

where   and   are rotational constants for the upper and lower vibrational state respectively, while   and   are the rotational quantum numbers of the upper and lower levels. In reality, this expression has to be modified for the effects of anharmonicity of the vibrations, for centrifugal distortion and for Coriolis coupling.[17]

For the so-called R branch of the spectrum,   so that there is simultaneous excitation of both vibration and rotation. For the P branch,   so that a quantum of rotational energy is lost while a quantum of vibrational energy is gained. The purely vibrational transition,  , gives rise to the Q branch of the spectrum. Because of the thermal population of the rotational states the P branch is slightly less intense than the R branch.

Rotational constants obtained from infrared measurements are in good accord with those obtained by microwave spectroscopy, while the latter usually offers greater precision.

Structure of rotational spectra edit

Spherical top edit

Spherical top molecules have no net dipole moment. A pure rotational spectrum cannot be observed by absorption or emission spectroscopy because there is no permanent dipole moment whose rotation can be accelerated by the electric field of an incident photon. Also the polarizability is isotropic, so that pure rotational transitions cannot be observed by Raman spectroscopy either. Nevertheless, rotational constants can be obtained by ro–vibrational spectroscopy. This occurs when a molecule is polar in the vibrationally excited state. For example, the molecule methane is a spherical top but the asymmetric C-H stretching band shows rotational fine structure in the infrared spectrum, illustrated in rovibrational coupling. This spectrum is also interesting because it shows clear evidence of Coriolis coupling in the asymmetric structure of the band.

Linear molecules edit

 
Energy levels and line positions calculated in the rigid rotor approximation

The rigid rotor is a good starting point from which to construct a model of a rotating molecule. It is assumed that component atoms are point masses connected by rigid bonds. A linear molecule lies on a single axis and each atom moves on the surface of a sphere around the centre of mass. The two degrees of rotational freedom correspond to the spherical coordinates θ and φ which describe the direction of the molecular axis, and the quantum state is determined by two quantum numbers J and M. J defines the magnitude of the rotational angular momentum, and M its component about an axis fixed in space, such as an external electric or magnetic field. In the absence of external fields, the energy depends only on J. Under the rigid rotor model, the rotational energy levels, F(J), of the molecule can be expressed as,

 

where   is the rotational constant of the molecule and is related to the moment of inertia of the molecule. In a linear molecule the moment of inertia about an axis perpendicular to the molecular axis is unique, that is,  , so

 

For a diatomic molecule

 

where m1 and m2 are the masses of the atoms and d is the distance between them.

Selection rules dictate that during emission or absorption the rotational quantum number has to change by unity; i.e.,  . Thus, the locations of the lines in a rotational spectrum will be given by

 

where   denotes the lower level and   denotes the upper level involved in the transition.

The diagram illustrates rotational transitions that obey the  =1 selection rule. The dashed lines show how these transitions map onto features that can be observed experimentally. Adjacent   transitions are separated by 2B in the observed spectrum. Frequency or wavenumber units can also be used for the x axis of this plot.

Rotational line intensities edit

 
Rotational level populations with Bhc/kT = 0.05. J is the quantum number of the lower rotational state.

The probability of a transition taking place is the most important factor influencing the intensity of an observed rotational line. This probability is proportional to the population of the initial state involved in the transition. The population of a rotational state depends on two factors. The number of molecules in an excited state with quantum number J, relative to the number of molecules in the ground state, NJ/N0 is given by the Boltzmann distribution as

 ,

where k is the Boltzmann constant and T the absolute temperature. This factor decreases as J increases. The second factor is the degeneracy of the rotational state, which is equal to 2J + 1. This factor increases as J increases. Combining the two factors[18]

 

The maximum relative intensity occurs at[19][notes 6]

 

The diagram at the right shows an intensity pattern roughly corresponding to the spectrum above it.

Centrifugal distortion edit

When a molecule rotates, the centrifugal force pulls the atoms apart. As a result, the moment of inertia of the molecule increases, thus decreasing the value of  , when it is calculated using the expression for the rigid rotor. To account for this a centrifugal distortion correction term is added to the rotational energy levels of the diatomic molecule.[20]

 

where   is the centrifugal distortion constant.

Therefore, the line positions for the rotational mode change to

 

In consequence, the spacing between lines is not constant, as in the rigid rotor approximation, but decreases with increasing rotational quantum number.

An assumption underlying these expressions is that the molecular vibration follows simple harmonic motion. In the harmonic approximation the centrifugal constant   can be derived as

 

where k is the vibrational force constant. The relationship between   and  

 

where   is the harmonic vibration frequency, follows. If anharmonicity is to be taken into account, terms in higher powers of J should be added to the expressions for the energy levels and line positions.[20] A striking example concerns the rotational spectrum of hydrogen fluoride which was fitted to terms up to [J(J+1)]5.[21]

Oxygen edit

The electric dipole moment of the dioxygen molecule, O
2
is zero, but the molecule is paramagnetic with two unpaired electrons so that there are magnetic-dipole allowed transitions which can be observed by microwave spectroscopy. The unit electron spin has three spatial orientations with respect to the given molecular rotational angular momentum vector, K, so that each rotational level is split into three states, J = K + 1, K, and K - 1, each J state of this so-called p-type triplet arising from a different orientation of the spin with respect to the rotational motion of the molecule. The energy difference between successive J terms in any of these triplets is about 2 cm−1 (60 GHz), with the single exception of J = 1←0 difference which is about 4 cm−1. Selection rules for magnetic dipole transitions allow transitions between successive members of the triplet (ΔJ = ±1) so that for each value of the rotational angular momentum quantum number K there are two allowed transitions. The 16O nucleus has zero nuclear spin angular momentum, so that symmetry considerations demand that K have only odd values.[22][23]

Symmetric top edit

For symmetric rotors a quantum number J is associated with the total angular momentum of the molecule. For a given value of J, there is a 2J+1- fold degeneracy with the quantum number, M taking the values +J ...0 ... -J. The third quantum number, K is associated with rotation about the principal rotation axis of the molecule. In the absence of an external electrical field, the rotational energy of a symmetric top is a function of only J and K and, in the rigid rotor approximation, the energy of each rotational state is given by

 

where   and   for a prolate symmetric top molecule or   for an oblate molecule.

This gives the transition wavenumbers as

 

which is the same as in the case of a linear molecule.[24] With a first order correction for centrifugal distortion the transition wavenumbers become

 

The term in DJK has the effect of removing degeneracy present in the rigid rotor approximation, with different K values.[25]

Asymmetric top edit

 
Pure rotation spectrum of atmospheric water vapour measured at Mauna Kea (33 cm−1 to 100 cm−1)

The quantum number J refers to the total angular momentum, as before. Since there are three independent moments of inertia, there are two other independent quantum numbers to consider, but the term values for an asymmetric rotor cannot be derived in closed form. They are obtained by individual matrix diagonalization for each J value. Formulae are available for molecules whose shape approximates to that of a symmetric top.[26]

The water molecule is an important example of an asymmetric top. It has an intense pure rotation spectrum in the far infrared region, below about 200 cm−1. For this reason far infrared spectrometers have to be freed of atmospheric water vapour either by purging with a dry gas or by evacuation. The spectrum has been analyzed in detail.[27]

Quadrupole splitting edit

When a nucleus has a spin quantum number, I, greater than 1/2 it has a quadrupole moment. In that case, coupling of nuclear spin angular momentum with rotational angular momentum causes splitting of the rotational energy levels. If the quantum number J of a rotational level is greater than I, 2I + 1 levels are produced; but if J is less than I, 2J + 1 levels result. The effect is one type of hyperfine splitting. For example, with 14N (I = 1) in HCN, all levels with J > 0 are split into 3. The energies of the sub-levels are proportional to the nuclear quadrupole moment and a function of F and J. where F = J + I, J + I − 1, …, |JI|. Thus, observation of nuclear quadrupole splitting permits the magnitude of the nuclear quadrupole moment to be determined.[28] This is an alternative method to the use of nuclear quadrupole resonance spectroscopy. The selection rule for rotational transitions becomes[29]

 

Stark and Zeeman effects edit

In the presence of a static external electric field the 2J + 1 degeneracy of each rotational state is partly removed, an instance of a Stark effect. For example, in linear molecules each energy level is split into J + 1 components. The extent of splitting depends on the square of the electric field strength and the square of the dipole moment of the molecule.[30] In principle this provides a means to determine the value of the molecular dipole moment with high precision. Examples include carbonyl sulfide, OCS, with μ = 0.71521 ± 0.00020 debye. However, because the splitting depends on μ2, the orientation of the dipole must be deduced from quantum mechanical considerations.[31]

A similar removal of degeneracy will occur when a paramagnetic molecule is placed in a magnetic field, an instance of the Zeeman effect. Most species which can be observed in the gaseous state are diamagnetic . Exceptions are odd-electron molecules such as nitric oxide, NO, nitrogen dioxide, NO
2
, some chlorine oxides and the hydroxyl radical. The Zeeman effect has been observed with dioxygen, O
2
[32]

Rotational Raman spectroscopy edit

Molecular rotational transitions can also be observed by Raman spectroscopy. Rotational transitions are Raman-allowed for any molecule with an anisotropic polarizability which includes all molecules except for spherical tops. This means that rotational transitions of molecules with no permanent dipole moment, which cannot be observed in absorption or emission, can be observed, by scattering, in Raman spectroscopy. Very high resolution Raman spectra can be obtained by adapting a Fourier Transform Infrared Spectrometer. An example is the spectrum of 15
N
2
. It shows the effect of nuclear spin, resulting in intensities variation of 3:1 in adjacent lines. A bond length of 109.9985 ± 0.0010 pm was deduced from the data.[33]

Instruments and methods edit

The great majority of contemporary spectrometers use a mixture of commercially available and bespoke components which users integrate according to their particular needs. Instruments can be broadly categorised according to their general operating principles. Although rotational transitions can be found across a very broad region of the electromagnetic spectrum, fundamental physical constraints exist on the operational bandwidth of instrument components. It is often impractical and costly to switch to measurements within an entirely different frequency region. The instruments and operating principals described below are generally appropriate to microwave spectroscopy experiments conducted at frequencies between 6 and 24 GHz.

Absorption cells and Stark modulation edit

A microwave spectrometer can be most simply constructed using a source of microwave radiation, an absorption cell into which sample gas can be introduced and a detector such as a superheterodyne receiver. A spectrum can be obtained by sweeping the frequency of the source while detecting the intensity of transmitted radiation. A simple section of waveguide can serve as an absorption cell. An important variation of the technique in which an alternating current is applied across electrodes within the absorption cell results in a modulation of the frequencies of rotational transitions. This is referred to as Stark modulation and allows the use of phase-sensitive detection methods offering improved sensitivity. Absorption spectroscopy allows the study of samples that are thermodynamically stable at room temperature. The first study of the microwave spectrum of a molecule (NH
3
) was performed by Cleeton & Williams in 1934.[34] Subsequent experiments exploited powerful sources of microwaves such as the klystron, many of which were developed for radar during the Second World War. The number of experiments in microwave spectroscopy surged immediately after the war. By 1948, Walter Gordy was able to prepare a review of the results contained in approximately 100 research papers.[35] Commercial versions[36] of microwave absorption spectrometer were developed by Hewlett-Packard in the 1970s and were once widely used for fundamental research. Most research laboratories now exploit either Balle-Flygare or chirped-pulse Fourier transform microwave (FTMW) spectrometers.

Fourier transform microwave (FTMW) spectroscopy edit

The theoretical framework[37] underpinning FTMW spectroscopy is analogous to that used to describe FT-NMR spectroscopy. The behaviour of the evolving system is described by optical Bloch equations. First, a short (typically 0-3 microsecond duration) microwave pulse is introduced on resonance with a rotational transition. Those molecules that absorb the energy from this pulse are induced to rotate coherently in phase with the incident radiation. De-activation of the polarisation pulse is followed by microwave emission that accompanies decoherence of the molecular ensemble. This free induction decay occurs on a timescale of 1-100 microseconds depending on instrument settings. Following pioneering work by Dicke and co-workers in the 1950s,[38] the first FTMW spectrometer was constructed by Ekkers and Flygare in 1975.[39]

Balle–Flygare FTMW spectrometer edit

Balle, Campbell, Keenan and Flygare demonstrated that the FTMW technique can be applied within a "free space cell" comprising an evacuated chamber containing a Fabry-Perot cavity.[40] This technique allows a sample to be probed only milliseconds after it undergoes rapid cooling to only a few kelvins in the throat of an expanding gas jet. This was a revolutionary development because (i) cooling molecules to low temperatures concentrates the available population in the lowest rotational energy levels. Coupled with benefits conferred by the use of a Fabry-Perot cavity, this brought a great enhancement in the sensitivity and resolution of spectrometers along with a reduction in the complexity of observed spectra; (ii) it became possible to isolate and study molecules that are very weakly bound because there is insufficient energy available for them to undergo fragmentation or chemical reaction at such low temperatures. William Klemperer was a pioneer in using this instrument for the exploration of weakly bound interactions. While the Fabry-Perot cavity of a Balle-Flygare FTMW spectrometer can typically be tuned into resonance at any frequency between 6 and 18 GHz, the bandwidth of individual measurements is restricted to about 1 MHz. An animation illustrates the operation of this instrument which is currently the most widely used tool for microwave spectroscopy.[41]

Chirped-Pulse FTMW spectrometer edit

Noting that digitisers and related electronics technology had significantly progressed since the inception of FTMW spectroscopy, B.H. Pate at the University of Virginia[42] designed a spectrometer[43] which retains many advantages of the Balle-Flygare FT-MW spectrometer while innovating in (i) the use of a high speed (>4 GS/s) arbitrary waveform generator to generate a "chirped" microwave polarisation pulse that sweeps up to 12 GHz in frequency in less than a microsecond and (ii) the use of a high speed (>40 GS/s) oscilloscope to digitise and Fourier transform the molecular free induction decay. The result is an instrument that allows the study of weakly bound molecules but which is able to exploit a measurement bandwidth (12 GHz) that is greatly enhanced compared with the Balle-Flygare FTMW spectrometer. Modified versions of the original CP-FTMW spectrometer have been constructed by a number of groups in the United States, Canada and Europe.[44][45] The instrument offers a broadband capability that is highly complementary to the high sensitivity and resolution offered by the Balle-Flygare design.

Notes edit

  1. ^ The spectrum was measured over a couple of hours with the aid of a chirped-pulse Fourier transform microwave spectrometer at the University of Bristol.
  2. ^ This article uses the molecular spectroscopist's convention of expressing the rotational constant   in cm−1. Therefore   in this article corresponds to   in the Rigid rotor article.
  3. ^ For a symmetric top, the values of the 2 moments of inertia can be used to derive 2 molecular parameters. Values from each additional isotopologue provide the information for one more molecular parameter. For asymmetric tops a single isotopologue provides information for at most 3 molecular parameters.
  4. ^ Such transitions are called electric dipole-allowed transitions. Other transitions involving quadrupoles, octupoles, hexadecapoles etc. may also be allowed but the spectral intensity is very much smaller, so these transitions are difficult to observe. Magnetic-dipole-allowed transitions can occur in paramagnetic molecules such as dioxygen, O
    2
    and nitric oxide, NO
  5. ^ In Raman spectroscopy the photon energies for Stokes and anti-Stokes scattering are respectively less than and greater than the incident photon energy. See the energy-level diagram at Raman spectroscopy.
  6. ^ This value of J corresponds to the maximum of the population considered as a continuous function of J. However, since only integer values of J are allowed, the maximum line intensity is observed for a neighboring integer J.

References edit

  1. ^ Gordy, W. (1970). A. Weissberger (ed.). Microwave Molecular Spectra in Technique of Organic Chemistry. Vol. IX. New York: Interscience.
  2. ^ Nair, K.P.R.; Demaison, J.; Wlodarczak, G.; Merke, I. (236). "Millimeterwave rotational spectrum and internal rotation in o-chlorotoluene". Journal of Molecular Spectroscopy. 237 (2): 137–142. Bibcode:2006JMoSp.237..137N. doi:10.1016/j.jms.2006.03.011.
  3. ^ Cheung, A.C.; Rank, D.M.; Townes, C.H.; Thornton, D.D. & Welch, W.J. (1968). "Detection of NH
    3
    molecules in the interstellar medium by their microwave emission spectra". Physical Review Letters. 21 (25): 1701–5. Bibcode:1968PhRvL..21.1701C. doi:10.1103/PhysRevLett.21.1701.
  4. ^ Ricaud, P.; Baron, P; de La Noë, J. (2004). "Quality assessment of ground-based microwave measurements of chlorine monoxide, ozone, and nitrogen dioxide from the NDSC radiometer at the Plateau de Bure". Ann. Geophys. 22 (6): 1903–15. Bibcode:2004AnGeo..22.1903R. doi:10.5194/angeo-22-1903-2004.
  5. ^ "Astrochemistry in Virginia". Retrieved 2 December 2012.
  6. ^ Atkins & de Paula 2006, p. 444
  7. ^ Banwell & McCash 1994, p. 99
  8. ^ Moment of inertia values from Atkins & de Paula 2006, p. 445
  9. ^ Hollas 1996, p. 95
  10. ^ Hollas 1996, p. 104 shows part of the observed rotational spectrum of silane
  11. ^ Atkins & de Paula 2006, p. 447
  12. ^ Banwell & McCash 1994, p. 49
  13. ^ Hollas 1996, p. 111
  14. ^ Atkins & de Paula 2006, pp. 474–5
  15. ^ a b Banwell & McCash 1994, Section 4.2, p. 105, Pure Rotational Raman Spectra
  16. ^ Alexander, A. J.; Kroto, H. W.; Walton, D. R. M. (1967). "The microwave spectrum, substitution structure and dipole moment of cyanobutadiyne". J. Mol. Spectrosc. 62 (2): 175–180. Bibcode:1976JMoSp..62..175A. doi:10.1016/0022-2852(76)90347-7. Illustrated in Hollas 1996, p. 97
  17. ^ Banwell & McCash 1994, p. 63.
  18. ^ Banwell & McCash 1994, p. 40
  19. ^ Atkins & de Paula 2006, p. 449
  20. ^ a b Banwell & McCash 1994, p. 45
  21. ^ Jennings, D.A.; Evenson, K.M; Zink, L.R.; Demuynck, C.; Destombes, J.L.; Lemoine, B; Johns, J.W.C. (April 1987). "High-resolution spectroscopy of HF from 40 to 1100 cm−1: Highly accurate rotational constants". Journal of Molecular Spectroscopy. 122 (2): 477–480. Bibcode:1987JMoSp.122..477J. doi:10.1016/0022-2852(87)90021-X.pdf
  22. ^ Strandberg, M. W. P.; Meng, C. Y.; Ingersoll, J. G. (1949). "The Microwave Absorption Spectrum of Oxygen". Phys. Rev. 75 (10): 1524–8. Bibcode:1949PhRv...75.1524S. doi:10.1103/PhysRev.75.1524.pdf
  23. ^ Krupenie, Paul H. (1972). "The Spectrum of Molecular Oxygen" (PDF). Journal of Physical and Chemical Reference Data. 1 (2): 423–534. Bibcode:1972JPCRD...1..423K. doi:10.1063/1.3253101.
  24. ^ Hollas 1996, p. 101
  25. ^ Hollas 1996, p. 102 shows the effect on the microwave spectrum of H
    3
    SiNCS
    .
  26. ^ Hollas 1996, p. 103
  27. ^ Hall, Richard T.; Dowling, Jerome M. (1967). "Pure Rotational Spectrum of Water Vapor". J. Chem. Phys. 47 (7): 2454–61. Bibcode:1967JChPh..47.2454H. doi:10.1063/1.1703330. Hall, Richard T.; Dowling, Jerome M. (1971). "Erratum: Pure Rotational Spectrum of Water Vapor". J. Chem. Phys. 54 (11): 4968. Bibcode:1971JChPh..54.4968H. doi:10.1063/1.1674785.
  28. ^ Simmons, James W.; Anderson, Wallace E.; Gordy, Walter (1950). "Microwave Spectrum and Molecular Constants of Hydrogen Cyanide". Phys. Rev. 77 (1): 77–79. Bibcode:1950PhRv...77...77S. doi:10.1103/PhysRev.77.77.
  29. ^ Chang, Raymond (1971). Basic Principles of Spectroscopy. McGraw-Hill. p139
  30. ^ Hollas 1996, p. 102 gives the equations for diatomic molecules and symmetric tops
  31. ^ Hollas 1996, p. 102
  32. ^ Burkhalter, James H.; Roy S. Anderson; William V. Smith; Walter Gordy (1950). "The Fine Structure of the Microwave Absorption Spectrum of Oxygen". Phys. Rev. 79 (4): 651–5. Bibcode:1950PhRv...79..651B. doi:10.1103/PhysRev.79.651.
  33. ^ Hollas 1996, p. 113, illustrates the spectrum of 15
    N
    2
    obtained using 476.5 nm radiation from an argon ion laser.
  34. ^ Cleeton, C.E.; Williams, N.H. (1934). "Electromagnetic waves of 1.1 cm wave-length and the absorption spectrum of ammonia". Physical Review. 45 (4): 234–7. Bibcode:1934PhRv...45..234C. doi:10.1103/PhysRev.45.234.
  35. ^ Gordy, W. (1948). "Microwave spectroscopy". Reviews of Modern Physics. 20 (4): 668–717. Bibcode:1948RvMP...20..668G. doi:10.1103/RevModPhys.20.668.
  36. ^ "June 1971, Hewlett-Packard Journal" (PDF).
  37. ^ Schwendemann, R.H. (1978). "Transient Effects in Microwave Spectroscopy". Annu. Rev. Phys. Chem. 29: 537–558. Bibcode:1978ARPC...29..537S. doi:10.1146/annurev.pc.29.100178.002541.
  38. ^ Dicke, R.H.; Romer, R.H. (1955). "Pulse Techniques in Microwave Spectroscopy". Rev. Sci. Instrum. 26 (10): 915–928. Bibcode:1955RScI...26..915D. doi:10.1063/1.1715156.
  39. ^ Ekkers, J.; Flygare, W.H. (1976). "Pulsed microwave Fourier transform spectrometer". Rev. Sci. Instrum. 47 (4): 448–454. Bibcode:1976RScI...47..448E. doi:10.1063/1.1134647.
  40. ^ Balle, T.J.; Campbell, E.J.; Keenan, M.R.; Flygare, W.H. (1980). "A new method for observing the rotational spectra of weak molecular complexes: KrHCl". J. Chem. Phys. 72 (2): 922–932. Bibcode:1980JChPh..72..922B. doi:10.1063/1.439210.
  41. ^ Jager, W. "Balle-Flygare FTMW spectrometer animation".
  42. ^ "Web page of B.H. Pate Research Group, Department of Chemistry, University of Virginia".
  43. ^ Brown, G.G.; Dian, B.C.; Douglass, K.O.; Geyer, S.M.; Pate, B.H. (2006). "The rotational spectrum of epifluorohydrin measured by chirped-pulse Fourier transform microwave spectroscopy". J. Mol. Spectrosc. 238 (2): 200–212. Bibcode:2006JMoSp.238..200B. doi:10.1016/j.jms.2006.05.003.
  44. ^ Grubbs, G.S.; Dewberry, C.T.; Etchison, K.C.; Kerr, K.E.; Cooke, S.A. (2007). "A search accelerated correct intensity Fourier transform microwave spectrometer with pulsed laser ablation source". Rev. Sci. Instrum. 78 (9): 096106–096106–3. Bibcode:2007RScI...78i6106G. doi:10.1063/1.2786022. PMID 17902981.
  45. ^ Wilcox, D.S.; Hotopp, K.M.; Dian, B.C. (2011). "Two-Dimensional Chirped-Pulse Fourier Transform Microwave Spectroscopy". J. Phys. Chem. A. 115 (32): 8895–8905. Bibcode:2011JPCA..115.8895W. doi:10.1021/jp2043202. PMID 21728367.

Bibliography edit

External links edit

  • infrared gas spectra simulator
  • Hyperphysics article on Rotational Spectrum
  • A list of microwave spectroscopy research groups around the world

rotational, spectroscopy, concerned, with, measurement, energies, transitions, between, quantized, rotational, states, molecules, phase, rotational, spectrum, power, spectral, density, rotational, frequency, polar, molecules, measured, absorption, emission, mi. 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 power spectral density vs rotational frequency of polar molecules can be measured in absorption or emission by microwave spectroscopy 1 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 or just vibronic spectroscopy where rotational vibrational and electronic energy changes occur simultaneously Part of the rotational spectrum of trifluoroiodomethane CF3 I notes 1 Each rotational transition is labeled with the quantum numbers J of the final and initial states and is extensively split by the effects of nuclear quadrupole coupling with the 127I nucleus For rotational spectroscopy molecules are classified according to symmetry into a spherical top linear and symmetric top analytical expressions can be derived for the rotational energy terms of these molecules Analytical expressions can be derived for the fourth category asymmetric top for rotational levels up to J 3 but higher energy levels need to be determined using numerical methods The rotational energies are derived theoretically by considering the molecules to be rigid rotors and then applying extra terms to account for centrifugal distortion fine structure hyperfine structure and Coriolis coupling Fitting the spectra to the theoretical expressions gives numerical values of the angular moments of inertia from which very precise values of molecular bond lengths and angles can be derived in favorable cases In the presence of an electrostatic field there is Stark splitting which allows molecular electric dipole moments to be determined An important application of rotational spectroscopy is in exploration of the chemical composition of the interstellar medium using radio telescopes Contents 1 Applications 2 Overview 2 1 Classification of molecular rotors 2 2 Selection rules 2 2 1 Microwave and far infrared spectra 2 2 2 Raman spectra 2 3 Units 2 4 Effect of vibration on rotation 2 5 Effect of rotation on vibrational spectra 3 Structure of rotational spectra 3 1 Spherical top 3 2 Linear molecules 3 2 1 Rotational line intensities 3 2 2 Centrifugal distortion 3 2 3 Oxygen 3 3 Symmetric top 3 4 Asymmetric top 4 Quadrupole splitting 5 Stark and Zeeman effects 6 Rotational Raman spectroscopy 7 Instruments and methods 7 1 Absorption cells and Stark modulation 7 2 Fourier transform microwave FTMW spectroscopy 7 2 1 Balle Flygare FTMW spectrometer 7 2 2 Chirped Pulse FTMW spectrometer 8 Notes 9 References 10 Bibliography 11 External linksApplications editRotational spectroscopy has primarily been used to investigate fundamental aspects of molecular physics It is a uniquely precise tool for the determination of molecular structure in gas phase molecules It can be used to establish barriers to internal rotation such as that associated with the rotation of the CH3 group relative to the C6 H4 Cl group in chlorotoluene C7 H7 Cl 2 When fine or hyperfine structure can be observed the technique also provides information on the electronic structures of molecules Much of current understanding of the nature of weak molecular interactions such as van der Waals hydrogen and halogen bonds has been established through rotational spectroscopy In connection with radio astronomy the technique has a key role in exploration of the chemical composition of the interstellar medium Microwave transitions are measured in the laboratory and matched to emissions from the interstellar medium using a radio telescope NH3 was the first stable polyatomic molecule to be identified in the interstellar medium 3 The measurement of chlorine monoxide 4 is important for atmospheric chemistry Current projects in astrochemistry involve both laboratory microwave spectroscopy and observations made using modern radiotelescopes such as the Atacama Large Millimeter submillimeter Array ALMA 5 Overview editA molecule in the gas phase is free to rotate relative to a set of mutually orthogonal axes of fixed orientation in space centered on the center of mass of the molecule Free rotation is not possible for molecules in liquid or solid phases due to the presence of intermolecular forces Rotation about each unique axis is associated with a set of quantized energy levels dependent on the moment of inertia about that axis and a quantum number Thus for linear molecules the energy levels are described by a single moment of inertia and a single quantum number J displaystyle J nbsp which defines the magnitude of the rotational angular momentum For nonlinear molecules which are symmetric rotors or symmetric tops see next section there are two moments of inertia and the energy also depends on a second rotational quantum number K displaystyle K nbsp which defines the vector component of rotational angular momentum along the principal symmetry axis 6 Analysis of spectroscopic data with the expressions detailed below results in quantitative determination of the value s of the moment s of inertia From these precise values of the molecular structure and dimensions may be obtained For a linear molecule analysis of the rotational spectrum provides values for the rotational constant notes 2 and the moment of inertia of the molecule and knowing the atomic masses can be used to determine the bond length directly For diatomic molecules this process is straightforward For linear molecules with more than two atoms it is necessary to measure the spectra of two or more isotopologues such as 16O12C32S and 16O12C34S This allows a set of simultaneous equations to be set up and solved for the bond lengths notes 3 A bond length obtained in this way is slightly different from the equilibrium bond length This is because there is zero point energy in the vibrational ground state to which the rotational states refer whereas the equilibrium bond length is at the minimum in the potential energy curve The relation between the rotational constants is given by B v B a v 1 2 displaystyle B v B alpha left v frac 1 2 right nbsp where v is a vibrational quantum number and a is a vibration rotation interaction constant which can be calculated if the B values for two different vibrational states can be found 7 For other molecules if the spectra can be resolved and individual transitions assigned both bond lengths and bond angles can be deduced When this is not possible as with most asymmetric tops all that can be done is to fit the spectra to three moments of inertia calculated from an assumed molecular structure By varying the molecular structure the fit can be improved giving a qualitative estimate of the structure Isotopic substitution is invaluable when using this approach to the determination of molecular structure Classification of molecular rotors edit In quantum mechanics the free rotation of a molecule is quantized so that the rotational energy and the angular momentum can take only certain fixed values which are related simply to the moment of inertia I displaystyle I nbsp of the molecule For any molecule there are three moments of inertia I A displaystyle I A nbsp I B displaystyle I B nbsp and I C displaystyle I C nbsp about three mutually orthogonal axes A B and C with the origin at the center of mass of the system The general convention used in this article is to define the axes such that I A I B I C displaystyle I A leq I B leq I C nbsp with axis A displaystyle A nbsp corresponding to the smallest moment of inertia Some authors however define the A displaystyle A nbsp axis as the molecular rotation axis of highest order The particular pattern of energy levels and hence of transitions in the rotational spectrum for a molecule is determined by its symmetry A convenient way to look at the molecules is to divide them into four different classes based on the symmetry of their structure These are Spherical tops spherical rotors All three moments of inertia are equal to each other I A I B I C displaystyle I A I B I C nbsp Examples of spherical tops include phosphorus tetramer P4 carbon tetrachloride CCl4 and other tetrahalides methane CH4 silane SiH4 sulfur hexafluoride SF6 and other hexahalides The molecules all belong to the cubic point groups Td or Oh Linear moleculesFor a linear molecule the moments of inertia are related by I A I B I C displaystyle I A ll I B I C nbsp For most purposes I A displaystyle I A nbsp can be taken to be zero Examples of linear molecules include dioxygen O2 dinitrogen N2 carbon monoxide CO hydroxy radical OH carbon dioxide CO2 hydrogen cyanide HCN carbonyl sulfide OCS acetylene ethyne HC CH and dihaloethynes These molecules belong to the point groups C v or D h Symmetric tops symmetric rotors A symmetric top is a molecule in which two moments of inertia are the same I A I B displaystyle I A I B nbsp or I B I C displaystyle I B I C nbsp By definition a symmetric top must have a 3 fold or higher order rotation axis As a matter of convenience spectroscopists divide molecules into two classes of symmetric tops Oblate symmetric tops saucer or disc shaped with I A I B lt I C displaystyle I A I B lt I C nbsp and Prolate symmetric tops rugby football or cigar shaped with I A lt I B I C displaystyle I A lt I B I C nbsp The spectra look rather different and are instantly recognizable Examples of symmetric tops includeOblate Benzene C6 H6 ammonia NH3 xenon tetrafluoride XeF4 Prolate Chloromethane CH3 Cl propyne CH3 C CH As a detailed example ammonia has a moment of inertia IC 4 4128 10 47 kg m2 about the 3 fold rotation axis and moments IA IB 2 8059 10 47 kg m2 about any axis perpendicular to the C3 axis Since the unique moment of inertia is larger than the other two the molecule is an oblate symmetric top 8 dd Asymmetric tops asymmetric rotors The three moments of inertia have different values Examples of small molecules that are asymmetric tops include water H2 O and nitrogen dioxide NO2 whose symmetry axis of highest order is a 2 fold rotation axis Most large molecules are asymmetric tops Selection rules edit Main article selection rules Microwave and far infrared spectra edit Transitions between rotational states can be observed in molecules with a permanent electric dipole moment 9 notes 4 A consequence of this rule is that no microwave spectrum can be observed for centrosymmetric linear molecules such as N2 dinitrogen or HCCH ethyne which are non polar Tetrahedral molecules such as CH4 methane which have both a zero dipole moment and isotropic polarizability would not have a pure rotation spectrum but for the effect of centrifugal distortion when the molecule rotates about a 3 fold symmetry axis a small dipole moment is created allowing a weak rotation spectrum to be observed by microwave spectroscopy 10 With symmetric tops the selection rule for electric dipole allowed pure rotation transitions is DK 0 DJ 1 Since these transitions are due to absorption or emission of a single photon with a spin of one conservation of angular momentum implies that the molecular angular momentum can change by at most one unit 11 Moreover the quantum number K is limited to have values between and including J to J 12 Raman spectra edit For Raman spectra the molecules undergo transitions in which an incident photon is absorbed and another scattered photon is emitted The general selection rule for such a transition to be allowed is that the molecular polarizability must be anisotropic which means that it is not the same in all directions 13 Polarizability is a 3 dimensional tensor that can be represented as an ellipsoid The polarizability ellipsoid of spherical top molecules is in fact spherical so those molecules show no rotational Raman spectrum For all other molecules both Stokes and anti Stokes lines notes 5 can be observed and they have similar intensities due to the fact that many rotational states are thermally populated The selection rule for linear molecules is DJ 0 2 The reason for the values 2 is that the polarizability returns to the same value twice during a rotation 14 The value DJ 0 does not correspond to a molecular transition but rather to Rayleigh scattering in which the incident photon merely changes direction 15 The selection rule for symmetric top molecules is DK 0 If K 0 then DJ 2 If K 0 then DJ 0 1 2Transitions with DJ 1 are said to belong to the R series whereas transitions with DJ 2 belong to an S series 15 Since Raman transitions involve two photons it is possible for the molecular angular momentum to change by two units Units edit The units used for rotational constants depend on the type of measurement With infrared spectra in the wavenumber scale n displaystyle tilde nu nbsp the unit is usually the inverse centimeter written as cm 1 which is literally the number of waves in one centimeter or the reciprocal of the wavelength in centimeters n 1 l displaystyle tilde nu 1 lambda nbsp On the other hand for microwave spectra in the frequency scale n displaystyle nu nbsp the unit is usually the gigahertz The relationship between these two units is derived from the expression n l c displaystyle nu cdot lambda c nbsp where n is a frequency l is a wavelength and c is the velocity of light It follows that n cm 1 1 l cm n s 1 c cm s 1 n s 1 2 99792458 10 10 displaystyle tilde nu text cm 1 frac 1 lambda text cm frac nu text s 1 c left text cm cdot mathrm s 1 right frac nu text s 1 2 99792458 times 10 10 nbsp As 1 GHz 109 Hz the numerical conversion can be expressed as n cm 1 n GHz 30 displaystyle tilde nu text cm 1 approx frac nu text GHz 30 nbsp Effect of vibration on rotation edit The population of vibrationally excited states follows a Boltzmann distribution so low frequency vibrational states are appreciably populated even at room temperatures As the moment of inertia is higher when a vibration is excited the rotational constants B decrease Consequently the rotation frequencies in each vibration state are different from each other This can give rise to satellite lines in the rotational spectrum An example is provided by cyanodiacetylene H C C C C C N 16 Further there is a fictitious force Coriolis coupling between the vibrational motion of the nuclei in the rotating non inertial frame However as long as the vibrational quantum number does not change i e the molecule is in only one state of vibration the effect of vibration on rotation is not important because the time for vibration is much shorter than the time required for rotation The Coriolis coupling is often negligible too if one is interested in low vibrational and rotational quantum numbers only Effect of rotation on vibrational spectra edit Main article Rotational vibrational spectroscopy Historically the theory of rotational energy levels was developed to account for observations of vibration rotation spectra of gases in infrared spectroscopy which was used before microwave spectroscopy had become practical To a first approximation the rotation and vibration can be treated as separable so the energy of rotation is added to the energy of vibration For example the rotational energy levels for linear molecules in the rigid rotor approximation are E rot h c B J J 1 displaystyle E text rot hcBJ J 1 nbsp In this approximation the vibration rotation wavenumbers of transitions are n n vib B J J 1 B J J 1 displaystyle tilde nu tilde nu text vib B J J 1 B J J 1 nbsp where B displaystyle B nbsp and B displaystyle B nbsp are rotational constants for the upper and lower vibrational state respectively while J displaystyle J nbsp and J displaystyle J nbsp are the rotational quantum numbers of the upper and lower levels In reality this expression has to be modified for the effects of anharmonicity of the vibrations for centrifugal distortion and for Coriolis coupling 17 For the so called R branch of the spectrum J J 1 displaystyle J J 1 nbsp so that there is simultaneous excitation of both vibration and rotation For the P branch J J 1 displaystyle J J 1 nbsp so that a quantum of rotational energy is lost while a quantum of vibrational energy is gained The purely vibrational transition D J 0 displaystyle Delta J 0 nbsp gives rise to the Q branch of the spectrum Because of the thermal population of the rotational states the P branch is slightly less intense than the R branch Rotational constants obtained from infrared measurements are in good accord with those obtained by microwave spectroscopy while the latter usually offers greater precision Structure of rotational spectra editSpherical top edit Spherical top molecules have no net dipole moment A pure rotational spectrum cannot be observed by absorption or emission spectroscopy because there is no permanent dipole moment whose rotation can be accelerated by the electric field of an incident photon Also the polarizability is isotropic so that pure rotational transitions cannot be observed by Raman spectroscopy either Nevertheless rotational constants can be obtained by ro vibrational spectroscopy This occurs when a molecule is polar in the vibrationally excited state For example the molecule methane is a spherical top but the asymmetric C H stretching band shows rotational fine structure in the infrared spectrum illustrated in rovibrational coupling This spectrum is also interesting because it shows clear evidence of Coriolis coupling in the asymmetric structure of the band Linear molecules edit nbsp Energy levels and line positions calculated in the rigid rotor approximationThe rigid rotor is a good starting point from which to construct a model of a rotating molecule It is assumed that component atoms are point masses connected by rigid bonds A linear molecule lies on a single axis and each atom moves on the surface of a sphere around the centre of mass The two degrees of rotational freedom correspond to the spherical coordinates 8 and f which describe the direction of the molecular axis and the quantum state is determined by two quantum numbers J and M J defines the magnitude of the rotational angular momentum and M its component about an axis fixed in space such as an external electric or magnetic field In the absence of external fields the energy depends only on J Under the rigid rotor model the rotational energy levels F J of the molecule can be expressed as F J B J J 1 J 0 1 2 displaystyle F left J right BJ left J 1 right qquad J 0 1 2 nbsp where B displaystyle B nbsp is the rotational constant of the molecule and is related to the moment of inertia of the molecule In a linear molecule the moment of inertia about an axis perpendicular to the molecular axis is unique that is I B I C I A 0 displaystyle I B I C I A 0 nbsp so B h 8 p 2 c I B h 8 p 2 c I C displaystyle B h over 8 pi 2 cI B h over 8 pi 2 cI C nbsp For a diatomic molecule I m 1 m 2 m 1 m 2 d 2 displaystyle I frac m 1 m 2 m 1 m 2 d 2 nbsp where m1 and m2 are the masses of the atoms and d is the distance between them Selection rules dictate that during emission or absorption the rotational quantum number has to change by unity i e D J J J 1 displaystyle Delta J J prime J prime prime pm 1 nbsp Thus the locations of the lines in a rotational spectrum will be given by n J J F J F J 2 B J 1 J 0 1 2 displaystyle tilde nu J prime leftrightarrow J prime prime F left J prime right F left J prime prime right 2B left J prime prime 1 right qquad J prime prime 0 1 2 nbsp where J displaystyle J prime prime nbsp denotes the lower level and J displaystyle J prime nbsp denotes the upper level involved in the transition The diagram illustrates rotational transitions that obey the D J displaystyle Delta J nbsp 1 selection rule The dashed lines show how these transitions map onto features that can be observed experimentally Adjacent J J displaystyle J prime prime leftarrow J prime nbsp transitions are separated by 2B in the observed spectrum Frequency or wavenumber units can also be used for the x axis of this plot Rotational line intensities edit nbsp Rotational level populations with Bhc kT 0 05 J is the quantum number of the lower rotational state The probability of a transition taking place is the most important factor influencing the intensity of an observed rotational line This probability is proportional to the population of the initial state involved in the transition The population of a rotational state depends on two factors The number of molecules in an excited state with quantum number J relative to the number of molecules in the ground state NJ N0 is given by the Boltzmann distribution as N J N 0 e E J k T e B h c J J 1 k T displaystyle frac N J N 0 e frac E J kT e frac BhcJ J 1 kT nbsp where k is the Boltzmann constant and T the absolute temperature This factor decreases as J increases The second factor is the degeneracy of the rotational state which is equal to 2J 1 This factor increases as J increases Combining the two factors 18 population 2 J 1 e E J k T displaystyle text population propto 2J 1 e frac E J kT nbsp The maximum relative intensity occurs at 19 notes 6 J k T 2 h c B 1 2 displaystyle J sqrt frac kT 2hcB frac 1 2 nbsp The diagram at the right shows an intensity pattern roughly corresponding to the spectrum above it Centrifugal distortion edit When a molecule rotates the centrifugal force pulls the atoms apart As a result the moment of inertia of the molecule increases thus decreasing the value of B displaystyle B nbsp when it is calculated using the expression for the rigid rotor To account for this a centrifugal distortion correction term is added to the rotational energy levels of the diatomic molecule 20 F J B J J 1 D J 2 J 1 2 J 0 1 2 displaystyle F left J right BJ left J 1 right DJ 2 left J 1 right 2 qquad J 0 1 2 nbsp where D displaystyle D nbsp is the centrifugal distortion constant Therefore the line positions for the rotational mode change to n J J 2 B J 1 4 D J 1 3 J 0 1 2 displaystyle tilde nu J prime leftrightarrow J prime prime 2B left J prime prime 1 right 4D left J prime prime 1 right 3 qquad J prime prime 0 1 2 nbsp In consequence the spacing between lines is not constant as in the rigid rotor approximation but decreases with increasing rotational quantum number An assumption underlying these expressions is that the molecular vibration follows simple harmonic motion In the harmonic approximation the centrifugal constant D displaystyle D nbsp can be derived as D h 3 32 p 4 I 2 r 2 k c displaystyle D frac h 3 32 pi 4 I 2 r 2 kc nbsp where k is the vibrational force constant The relationship between B displaystyle B nbsp and D displaystyle D nbsp D 4 B 3 w 2 displaystyle D frac 4B 3 tilde omega 2 nbsp where w displaystyle tilde omega nbsp is the harmonic vibration frequency follows If anharmonicity is to be taken into account terms in higher powers of J should be added to the expressions for the energy levels and line positions 20 A striking example concerns the rotational spectrum of hydrogen fluoride which was fitted to terms up to J J 1 5 21 Oxygen edit The electric dipole moment of the dioxygen molecule O2 is zero but the molecule is paramagnetic with two unpaired electrons so that there are magnetic dipole allowed transitions which can be observed by microwave spectroscopy The unit electron spin has three spatial orientations with respect to the given molecular rotational angular momentum vector K so that each rotational level is split into three states J K 1 K and K 1 each J state of this so called p type triplet arising from a different orientation of the spin with respect to the rotational motion of the molecule The energy difference between successive J terms in any of these triplets is about 2 cm 1 60 GHz with the single exception of J 1 0 difference which is about 4 cm 1 Selection rules for magnetic dipole transitions allow transitions between successive members of the triplet DJ 1 so that for each value of the rotational angular momentum quantum number K there are two allowed transitions The 16O nucleus has zero nuclear spin angular momentum so that symmetry considerations demand that K have only odd values 22 23 Symmetric top edit For symmetric rotors a quantum number J is associated with the total angular momentum of the molecule For a given value of J there is a 2J 1 fold degeneracy with the quantum number M taking the values J 0 J The third quantum number K is associated with rotation about the principal rotation axis of the molecule In the absence of an external electrical field the rotational energy of a symmetric top is a function of only J and K and in the rigid rotor approximation the energy of each rotational state is given by F J K B J J 1 A B K 2 J 0 1 2 and K J 0 J displaystyle F left J K right BJ left J 1 right left A B right K 2 qquad J 0 1 2 ldots quad mbox and quad K J ldots 0 ldots J nbsp where B h 8 p 2 c I B displaystyle B h over 8 pi 2 cI B nbsp and A h 8 p 2 c I A displaystyle A h over 8 pi 2 cI A nbsp for a prolate symmetric top molecule or A h 8 p 2 c I C displaystyle A h over 8 pi 2 cI C nbsp for an oblate molecule This gives the transition wavenumbers as n J J K F J K F J K 2 B J 1 J 0 1 2 displaystyle tilde nu J prime leftrightarrow J prime prime K F left J prime K right F left J prime prime K right 2B left J prime prime 1 right qquad J prime prime 0 1 2 nbsp which is the same as in the case of a linear molecule 24 With a first order correction for centrifugal distortion the transition wavenumbers become n J J K F J K F J K 2 B 2 D J K K 2 J 1 4 D J J 1 3 J 0 1 2 displaystyle tilde nu J prime leftrightarrow J prime prime K F left J prime K right F left J prime prime K right 2 left B 2D JK K 2 right left J prime prime 1 right 4D J left J prime prime 1 right 3 qquad J prime prime 0 1 2 nbsp The term in DJK has the effect of removing degeneracy present in the rigid rotor approximation with different K values 25 Asymmetric top edit nbsp Pure rotation spectrum of atmospheric water vapour measured at Mauna Kea 33 cm 1 to 100 cm 1 The quantum number J refers to the total angular momentum as before Since there are three independent moments of inertia there are two other independent quantum numbers to consider but the term values for an asymmetric rotor cannot be derived in closed form They are obtained by individual matrix diagonalization for each J value Formulae are available for molecules whose shape approximates to that of a symmetric top 26 The water molecule is an important example of an asymmetric top It has an intense pure rotation spectrum in the far infrared region below about 200 cm 1 For this reason far infrared spectrometers have to be freed of atmospheric water vapour either by purging with a dry gas or by evacuation The spectrum has been analyzed in detail 27 Quadrupole splitting editWhen a nucleus has a spin quantum number I greater than 1 2 it has a quadrupole moment In that case coupling of nuclear spin angular momentum with rotational angular momentum causes splitting of the rotational energy levels If the quantum number J of a rotational level is greater than I 2I 1 levels are produced but if J is less than I 2J 1 levels result The effect is one type of hyperfine splitting For example with 14N I 1 in HCN all levels with J gt 0 are split into 3 The energies of the sub levels are proportional to the nuclear quadrupole moment and a function of F and J where F J I J I 1 J I Thus observation of nuclear quadrupole splitting permits the magnitude of the nuclear quadrupole moment to be determined 28 This is an alternative method to the use of nuclear quadrupole resonance spectroscopy The selection rule for rotational transitions becomes 29 D J 1 D F 0 1 displaystyle Delta J pm 1 Delta F 0 pm 1 nbsp Stark and Zeeman effects editIn the presence of a static external electric field the 2J 1 degeneracy of each rotational state is partly removed an instance of a Stark effect For example in linear molecules each energy level is split into J 1 components The extent of splitting depends on the square of the electric field strength and the square of the dipole moment of the molecule 30 In principle this provides a means to determine the value of the molecular dipole moment with high precision Examples include carbonyl sulfide OCS with m 0 71521 0 00020 debye However because the splitting depends on m2 the orientation of the dipole must be deduced from quantum mechanical considerations 31 A similar removal of degeneracy will occur when a paramagnetic molecule is placed in a magnetic field an instance of the Zeeman effect Most species which can be observed in the gaseous state are diamagnetic Exceptions are odd electron molecules such as nitric oxide NO nitrogen dioxide NO2 some chlorine oxides and the hydroxyl radical The Zeeman effect has been observed with dioxygen O2 32 Rotational Raman spectroscopy editMolecular rotational transitions can also be observed by Raman spectroscopy Rotational transitions are Raman allowed for any molecule with an anisotropic polarizability which includes all molecules except for spherical tops This means that rotational transitions of molecules with no permanent dipole moment which cannot be observed in absorption or emission can be observed by scattering in Raman spectroscopy Very high resolution Raman spectra can be obtained by adapting a Fourier Transform Infrared Spectrometer An example is the spectrum of 15 N2 It shows the effect of nuclear spin resulting in intensities variation of 3 1 in adjacent lines A bond length of 109 9985 0 0010 pm was deduced from the data 33 Instruments and methods editThe great majority of contemporary spectrometers use a mixture of commercially available and bespoke components which users integrate according to their particular needs Instruments can be broadly categorised according to their general operating principles Although rotational transitions can be found across a very broad region of the electromagnetic spectrum fundamental physical constraints exist on the operational bandwidth of instrument components It is often impractical and costly to switch to measurements within an entirely different frequency region The instruments and operating principals described below are generally appropriate to microwave spectroscopy experiments conducted at frequencies between 6 and 24 GHz Absorption cells and Stark modulation edit A microwave spectrometer can be most simply constructed using a source of microwave radiation an absorption cell into which sample gas can be introduced and a detector such as a superheterodyne receiver A spectrum can be obtained by sweeping the frequency of the source while detecting the intensity of transmitted radiation A simple section of waveguide can serve as an absorption cell An important variation of the technique in which an alternating current is applied across electrodes within the absorption cell results in a modulation of the frequencies of rotational transitions This is referred to as Stark modulation and allows the use of phase sensitive detection methods offering improved sensitivity Absorption spectroscopy allows the study of samples that are thermodynamically stable at room temperature The first study of the microwave spectrum of a molecule NH3 was performed by Cleeton amp Williams in 1934 34 Subsequent experiments exploited powerful sources of microwaves such as the klystron many of which were developed for radar during the Second World War The number of experiments in microwave spectroscopy surged immediately after the war By 1948 Walter Gordy was able to prepare a review of the results contained in approximately 100 research papers 35 Commercial versions 36 of microwave absorption spectrometer were developed by Hewlett Packard in the 1970s and were once widely used for fundamental research Most research laboratories now exploit either Balle Flygare or chirped pulse Fourier transform microwave FTMW spectrometers Fourier transform microwave FTMW spectroscopy edit The theoretical framework 37 underpinning FTMW spectroscopy is analogous to that used to describe FT NMR spectroscopy The behaviour of the evolving system is described by optical Bloch equations First a short typically 0 3 microsecond duration microwave pulse is introduced on resonance with a rotational transition Those molecules that absorb the energy from this pulse are induced to rotate coherently in phase with the incident radiation De activation of the polarisation pulse is followed by microwave emission that accompanies decoherence of the molecular ensemble This free induction decay occurs on a timescale of 1 100 microseconds depending on instrument settings Following pioneering work by Dicke and co workers in the 1950s 38 the first FTMW spectrometer was constructed by Ekkers and Flygare in 1975 39 Balle Flygare FTMW spectrometer edit Balle Campbell Keenan and Flygare demonstrated that the FTMW technique can be applied within a free space cell comprising an evacuated chamber containing a Fabry Perot cavity 40 This technique allows a sample to be probed only milliseconds after it undergoes rapid cooling to only a few kelvins in the throat of an expanding gas jet This was a revolutionary development because i cooling molecules to low temperatures concentrates the available population in the lowest rotational energy levels Coupled with benefits conferred by the use of a Fabry Perot cavity this brought a great enhancement in the sensitivity and resolution of spectrometers along with a reduction in the complexity of observed spectra ii it became possible to isolate and study molecules that are very weakly bound because there is insufficient energy available for them to undergo fragmentation or chemical reaction at such low temperatures William Klemperer was a pioneer in using this instrument for the exploration of weakly bound interactions While the Fabry Perot cavity of a Balle Flygare FTMW spectrometer can typically be tuned into resonance at any frequency between 6 and 18 GHz the bandwidth of individual measurements is restricted to about 1 MHz An animation illustrates the operation of this instrument which is currently the most widely used tool for microwave spectroscopy 41 Chirped Pulse FTMW spectrometer edit Noting that digitisers and related electronics technology had significantly progressed since the inception of FTMW spectroscopy B H Pate at the University of Virginia 42 designed a spectrometer 43 which retains many advantages of the Balle Flygare FT MW spectrometer while innovating in i the use of a high speed gt 4 GS s arbitrary waveform generator to generate a chirped microwave polarisation pulse that sweeps up to 12 GHz in frequency in less than a microsecond and ii the use of a high speed gt 40 GS s oscilloscope to digitise and Fourier transform the molecular free induction decay The result is an instrument that allows the study of weakly bound molecules but which is able to exploit a measurement bandwidth 12 GHz that is greatly enhanced compared with the Balle Flygare FTMW spectrometer Modified versions of the original CP FTMW spectrometer have been constructed by a number of groups in the United States Canada and Europe 44 45 The instrument offers a broadband capability that is highly complementary to the high sensitivity and resolution offered by the Balle Flygare design Notes edit The spectrum was measured over a couple of hours with the aid of a chirped pulse Fourier transform microwave spectrometer at the University of Bristol This article uses the molecular spectroscopist s convention of expressing the rotational constant B displaystyle B nbsp in cm 1 Therefore B displaystyle B nbsp in this article corresponds to B B h c displaystyle bar B B hc nbsp in the Rigid rotor article For a symmetric top the values of the 2 moments of inertia can be used to derive 2 molecular parameters Values from each additional isotopologue provide the information for one more molecular parameter For asymmetric tops a single isotopologue provides information for at most 3 molecular parameters Such transitions are called electric dipole allowed transitions Other transitions involving quadrupoles octupoles hexadecapoles etc may also be allowed but the spectral intensity is very much smaller so these transitions are difficult to observe Magnetic dipole allowed transitions can occur in paramagnetic molecules such as dioxygen O2 and nitric oxide NO In Raman spectroscopy the photon energies for Stokes and anti Stokes scattering are respectively less than and greater than the incident photon energy See the energy level diagram at Raman spectroscopy This value of J corresponds to the maximum of the population considered as a continuous function of J However since only integer values of J are allowed the maximum line intensity is observed for a neighboring integer J References edit Gordy W 1970 A Weissberger ed Microwave Molecular Spectra in Technique of Organic Chemistry Vol IX New York Interscience Nair K P R Demaison J Wlodarczak G Merke I 236 Millimeterwave rotational spectrum and internal rotation in o chlorotoluene Journal of Molecular Spectroscopy 237 2 137 142 Bibcode 2006JMoSp 237 137N doi 10 1016 j jms 2006 03 011 Cheung A C Rank D M Townes C H Thornton D D amp Welch W J 1968 Detection of NH3 molecules in the interstellar medium by their microwave emission spectra Physical Review Letters 21 25 1701 5 Bibcode 1968PhRvL 21 1701C doi 10 1103 PhysRevLett 21 1701 Ricaud P Baron P de La Noe J 2004 Quality assessment of ground based microwave measurements of chlorine monoxide ozone and nitrogen dioxide from the NDSC radiometer at the Plateau de Bure Ann Geophys 22 6 1903 15 Bibcode 2004AnGeo 22 1903R doi 10 5194 angeo 22 1903 2004 Astrochemistry in Virginia Retrieved 2 December 2012 Atkins amp de Paula 2006 p 444 Banwell amp McCash 1994 p 99 Moment of inertia values from Atkins amp de Paula 2006 p 445 Hollas 1996 p 95 Hollas 1996 p 104 shows part of the observed rotational spectrum of silane Atkins amp de Paula 2006 p 447 Banwell amp McCash 1994 p 49 Hollas 1996 p 111 Atkins amp de Paula 2006 pp 474 5 a b Banwell amp McCash 1994 Section 4 2 p 105 Pure Rotational Raman Spectra Alexander A J Kroto H W Walton D R M 1967 The microwave spectrum substitution structure and dipole moment of cyanobutadiyne J Mol Spectrosc 62 2 175 180 Bibcode 1976JMoSp 62 175A doi 10 1016 0022 2852 76 90347 7 Illustrated in Hollas 1996 p 97 Banwell amp McCash 1994 p 63 Banwell amp McCash 1994 p 40 Atkins amp de Paula 2006 p 449 a b Banwell amp McCash 1994 p 45 Jennings D A Evenson K M Zink L R Demuynck C Destombes J L Lemoine B Johns J W C April 1987 High resolution spectroscopy of HF from 40 to 1100 cm 1 Highly accurate rotational constants Journal of Molecular Spectroscopy 122 2 477 480 Bibcode 1987JMoSp 122 477J doi 10 1016 0022 2852 87 90021 X pdf Strandberg M W P Meng C Y Ingersoll J G 1949 The Microwave Absorption Spectrum of Oxygen Phys Rev 75 10 1524 8 Bibcode 1949PhRv 75 1524S doi 10 1103 PhysRev 75 1524 pdf Krupenie Paul H 1972 The Spectrum of Molecular Oxygen PDF Journal of Physical and Chemical Reference Data 1 2 423 534 Bibcode 1972JPCRD 1 423K doi 10 1063 1 3253101 Hollas 1996 p 101 Hollas 1996 p 102 shows the effect on the microwave spectrum of H3 SiNCS Hollas 1996 p 103 Hall Richard T Dowling Jerome M 1967 Pure Rotational Spectrum of Water Vapor J Chem Phys 47 7 2454 61 Bibcode 1967JChPh 47 2454H doi 10 1063 1 1703330 Hall Richard T Dowling Jerome M 1971 Erratum Pure Rotational Spectrum of Water Vapor J Chem Phys 54 11 4968 Bibcode 1971JChPh 54 4968H doi 10 1063 1 1674785 Simmons James W Anderson Wallace E Gordy Walter 1950 Microwave Spectrum and Molecular Constants of Hydrogen Cyanide Phys Rev 77 1 77 79 Bibcode 1950PhRv 77 77S doi 10 1103 PhysRev 77 77 Chang Raymond 1971 Basic Principles of Spectroscopy McGraw Hill p139 Hollas 1996 p 102 gives the equations for diatomic molecules and symmetric tops Hollas 1996 p 102 Burkhalter James H Roy S Anderson William V Smith Walter Gordy 1950 The Fine Structure of the Microwave Absorption Spectrum of Oxygen Phys Rev 79 4 651 5 Bibcode 1950PhRv 79 651B doi 10 1103 PhysRev 79 651 Hollas 1996 p 113 illustrates the spectrum of 15 N2 obtained using 476 5 nm radiation from an argon ion laser Cleeton C E Williams N H 1934 Electromagnetic waves of 1 1 cm wave length and the absorption spectrum of ammonia Physical Review 45 4 234 7 Bibcode 1934PhRv 45 234C doi 10 1103 PhysRev 45 234 Gordy W 1948 Microwave spectroscopy Reviews of Modern Physics 20 4 668 717 Bibcode 1948RvMP 20 668G doi 10 1103 RevModPhys 20 668 June 1971 Hewlett Packard Journal PDF Schwendemann R H 1978 Transient Effects in Microwave Spectroscopy Annu Rev Phys Chem 29 537 558 Bibcode 1978ARPC 29 537S doi 10 1146 annurev pc 29 100178 002541 Dicke R H Romer R H 1955 Pulse Techniques in Microwave Spectroscopy Rev Sci Instrum 26 10 915 928 Bibcode 1955RScI 26 915D doi 10 1063 1 1715156 Ekkers J Flygare W H 1976 Pulsed microwave Fourier transform spectrometer Rev Sci Instrum 47 4 448 454 Bibcode 1976RScI 47 448E doi 10 1063 1 1134647 Balle T J Campbell E J Keenan M R Flygare W H 1980 A new method for observing the rotational spectra of weak molecular complexes KrHCl J Chem Phys 72 2 922 932 Bibcode 1980JChPh 72 922B doi 10 1063 1 439210 Jager W Balle Flygare FTMW spectrometer animation Web page of B H Pate Research Group Department of Chemistry University of Virginia Brown G G Dian B C Douglass K O Geyer S M Pate B H 2006 The rotational spectrum of epifluorohydrin measured by chirped pulse Fourier transform microwave spectroscopy J Mol Spectrosc 238 2 200 212 Bibcode 2006JMoSp 238 200B doi 10 1016 j jms 2006 05 003 Grubbs G S Dewberry C T Etchison K C Kerr K E Cooke S A 2007 A search accelerated correct intensity Fourier transform microwave spectrometer with pulsed laser ablation source Rev Sci Instrum 78 9 096106 096106 3 Bibcode 2007RScI 78i6106G doi 10 1063 1 2786022 PMID 17902981 Wilcox D S Hotopp K M Dian B C 2011 Two Dimensional Chirped Pulse Fourier Transform Microwave Spectroscopy J Phys Chem A 115 32 8895 8905 Bibcode 2011JPCA 115 8895W doi 10 1021 jp2043202 PMID 21728367 Bibliography editAtkins P W de Paula J 2006 Molecular Spectroscopy Section Pure rotation spectra Physical Chemistry 8th ed Oxford University Press pp 431 469 ISBN 0198700725 Banwell Colin N McCash Elaine M 1994 Fundamentals of Molecular Spectroscopy 4th ed McGraw Hill ISBN 0 07 707976 0 Brown John M Carrington Alan 2003 Rotational spectroscopy of diatomic molecule Cambridge University Press ISBN 0 521 53078 4 Hollas M J 1996 Modern Spectroscopy 3rd ed Wiley ISBN 0471965227 Kroto H W 2003 Molecular Rotation Spectroscopy Dover ISBN 0 486 49540 X McQuarrie Donald A 2008 Quantum Chemistry University Science Books ISBN 978 1 891389 50 4 Townes Charles H Schawlow Arthur L 1975 Microwave Spectroscopy Dover ISBN 978 0 486 61798 5 Kovacs Istvan 1969 Rotational Structure in the Spectra of Diatomic Molecules Adam Hilger ISBN 0852741421 Wollrab James E 1967 Rotational spectra and molecular structure Academic Press ISBN 148319485X External links editinfrared gas spectra simulator Hyperphysics article on Rotational Spectrum A list of microwave spectroscopy research groups around the world Retrieved from https en wikipedia org w index php title Rotational spectroscopy amp oldid 1172101249, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.