Download A Survey of Molecules With Large Nuclear

Document related concepts

Tensor wikipedia , lookup

Nuclear physics wikipedia , lookup

Circular dichroism wikipedia , lookup

Transcript
Wesleyan University
The Honors College
A SURVEY OF MOLECULES WITH LARGE
NUCLEAR QUADRUPOLES
AND OTHER PROJECTS
by
Eric A. Arsenault
Class of 2017
A thesis (or essay) submitted to the
faculty of Wesleyan University
in partial fulfillment of the requirements for the
Degree of Bachelor of Arts
with Departmental Honors in Chemistry
Middletown, Connecticut
April, 2017
“Let me show ya something.” —Old Gregg
For Dr. Dan Obenchain.
i
Acknowledgements
As a member of the Novick/Pringle/Cooke research group, I have grown
fantastically. With this group, I experienced my first flight and I became a published scientist. There are quite a few people to thank:
I want to thank Professor Stew Novick, Professor Wallace “Pete” Pringle,
and Professor Steve Cooke for all of their support, wisdom, and guidance and for
allowing me to join their research group despite the fact that at the time, I had
not the slightest idea of what the hell they were doing.
Over the past three years, Professor Novick has been exceptionally important in shaping the researcher that I am now and the one that I aspire to be.
Thank you.
I want to thank Dr. Dan Obenchain for his perpetual support, patience,
encouragement, and for showing me my first episode of The Mighty Boosh. I also
would like to thank Dr. Dan for editing and re-editing my work, for listening to
practice presentation after practice presentation, for his questions (“Where are
the numbers?”), and for continuously answering my questions. Thank you.
I want to thank the many scientists that I was fortunate to collaborate
with during my time in the Novick/Pringle/Cooke group, especially Dr. Thomas
Blake, Professor Karen Peterson, and Professor Wei Lin. I also want to give many
thanks to the current and former members of the Novick/Pringle/Cooke group
that helped to make my experience what is was, including Dr. Brittany Long,
Derek Frank, Yoon Jeong Choi, Angela Chung, Will Orellana, Robert Melchreit,
and Dr. Sue Stephens.
Outside of the laboratory, I would like to acknowledge the many professors
ii
that I have interacted with on a daily basis during my time at Wesleyan, especially
Professor Joseph Knee. Outside both the laboratory and the classroom, I am
so grateful for my family and friends. Shout-outs to Mom, Dad, Emily, Luke,
Grandma, Charlie, and Roxy (meow ). Thank you to Isabel for teaching me
about the present and life beyond equations. I also want to thank the Wesleyan
Cross Country team, specifically the many teammates that ran thousands and
thousands of miles alongside me over the past few years. Lastly, a hat tip to the
residents of 52 Home Avenue.
iii
Contents
1 Introduction
1.1
1
Microwave Spectroscopy . . . . . . . . . . . . . . . . . . . . . . .
1
1.1.1
The Rigid Rotor . . . . . . . . . . . . . . . . . . . . . . .
2
1.1.2
Centrifugal Distortion . . . . . . . . . . . . . . . . . . . .
3
1.1.3
Hyperfine Interactions . . . . . . . . . . . . . . . . . . . .
4
1.1.4
Nuclear Quadrupole Interaction . . . . . . . . . . . . . . .
5
1.1.5
Nuclear Spin-Rotation Interaction . . . . . . . . . . . . . .
7
1.1.6
The Full Hamiltonian . . . . . . . . . . . . . . . . . . . . .
8
2 A Study of 2-Iodobutane by Rotational Spectroscopy
10
2.1
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10
2.2
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11
2.3
Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
12
2.3.1
Instrumentation . . . . . . . . . . . . . . . . . . . . . . . .
12
2.3.2
Quantum Chemical Calculations . . . . . . . . . . . . . . .
14
2.3.3
Spectral Assignments . . . . . . . . . . . . . . . . . . . . .
16
2.3.4
Hyperfine Structure . . . . . . . . . . . . . . . . . . . . . .
18
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
22
2.4.1
Structural Determination . . . . . . . . . . . . . . . . . . .
22
2.4.2
Nuclear Quadrupole Coupling Tensor of Iodine . . . . . . .
25
2.5
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
29
2.6
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . .
30
2.7
Supplemental Information . . . . . . . . . . . . . . . . . . . . . .
30
2.4
iv
3 A Study of the Conformational Isomerism of 1-Iodobutane by
High Resolution Rotational Spectroscopy
33
3.1
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
33
3.2
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
34
3.3
Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
35
3.3.1
Quantum Chemical Calculations . . . . . . . . . . . . . . .
36
3.3.2
Spectral Assignments . . . . . . . . . . . . . . . . . . . . .
36
3.3.3
Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
42
3.4
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
43
3.5
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
48
3.6
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . .
48
3.7
Supplemental Material . . . . . . . . . . . . . . . . . . . . . . . .
49
4 Nuclear Quadrupole Coupling in SiH2 I2 due to the Presence of
Two Iodine Nuclei
53
4.1
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
53
4.2
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
54
4.3
Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
55
4.4
Spectral Assignments . . . . . . . . . . . . . . . . . . . . . . . . .
55
4.5
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
60
4.5.1
Structural Determination . . . . . . . . . . . . . . . . . . .
60
4.5.2
Nuclear Quadrupole Coupling Tensor of Iodine . . . . . . .
60
4.5.3
Chemical Nature of the Si−I Bond . . . . . . . . . . . . .
62
4.6
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
64
4.7
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . .
64
4.8
Supplemental Material . . . . . . . . . . . . . . . . . . . . . . . .
64
5 Assorted Projects
72
6 Appendix A
84
v
Chapter 1
Introduction
1.1
Microwave Spectroscopy
This section will provide a brief synopsis of the theory behind microwave
spectroscopy. Information pertaining to instrumentation, experimental setup, and
spectral analysis will be left to the following chapters, where the presentation of
these details has been tailored to the project under discussion.
Microwave spectroscopy is the study of the interaction between microwave
radiation, which spans the 300 MHz to 300 GHz portion of the electromagnetic
spectrum, and matter, typically in the gas phase[1]. In the simplest case, this
leads to the observation of a pure rotational spectrum, although in practice many
complexities arise. For brevity, a mere outline of the theory behind microwave
spectroscopy will be presented here. Molecular Rotation Spectra by H. W. Kroto,
Microwave Spectroscopy by C. H. Townes and A. L. Schawlow, and Microwave
Molecular Spectra by Walter Gordy and R. L. Cook can serve the reader as a
more comprehensive set of resources for the concepts presented below[4–6].
1
1.1.1
The Rigid Rotor
To investigate the origin of the rigid rotor Hamiltonian, it is best to start
with the classical expression for pure rotational kinetic energy given by
1
Tr = ω † Iω
2
where

I
0 0
 aa

I =  0 Ibb 0

0 0 Icc
(1.1)





(1.2)
is the diagonalized inertia tensor, I, which, by definition, is a projection of the
tensor in the principal axis system of the molecule under investigation[2]. Similarly,

ωa





ω =  ωb 


ωc
(1.3)
is the angular velocity about the principal axes of the molecule. The derivation
for Tr itself can be found in most texts on classical dynamics[3]. Next, defining
the angular momentum, P , as
P = Iω
(1.4)
leads to the rearrangement of Tr such that
1
1
1
Tr = ω † Iω = (Iω)† I −1 (Iω) = P † I −1 P .
2
2
2
(1.5)
When expressed in this way, Tr can be used directly to formulate the classical
rigid rotor Hamiltonian
1 † −1
1 Pa2 Pb2 Pc2
Ĥ = P I P = (
+
+
).
2
2 Iaa Ibb
Icc
(1.6)
By replacing the general angular momentum operator, P , with the corresponding conjugate (quantum mechanical) angular momentum operator, J , the quantum mechanical rigid rotor Hamiltonian can be easily determined to be of the
form
J2
J2
1 J2
ĤR = ( a + b + c )
2 Iaa Ibb Icc
2
(1.7)
Table 1.1: Classes of rigid rotors
A=B=C
spherical top
B=C
linear molecule
A>B=C
prolate symmetric top
A=B>C
oblate symmetric top
A>B>C
asymmetric rotor
[4–6]. By convention, so that the components of J are expressed in units of ~,
ĤR is written as
ĤR = AJa2 + BJb2 + CJc2
where A, B, and C, are the principal rotational constants given by
~2
,
2Icc
(1.8)
2
~2
, ~ ,
2Iaa 2Ibb
and
respectively. Also by convention, IA ≤ IB ≤ IC such that A ≥ B ≥ C. These
distinctions lead to several specific classes of rigid rotors, which are presented in
Table 1.1.
Lastly, the eigenstates of ĤR , for the most general rotor, which correspond
to the rotational energy levels, are labeled by quantum numbers J
KA
can have values of J to 0 and
KC
KA KC ,
where
can have values of 0 to J. For each class
of rigid rotor, specific selection rules determine the allowed transitions between
rotational energy levels, however, these rules will not be presented here.
1.1.2
Centrifugal Distortion
In experiment, observed rotational transitions rarely conform to the pre-
dictions yielded by the final form of ĤR in (1.8), as this Hamiltonian is based
on the approximation that the nuclear framework of the molecule is fixed. In
practice, centrifugal force produced by molecular rotation causes a distortion in
the geometry of the molecule, which is to say that it is best to treat molecular
systems as semi-rigid rotors. The consequence of this is that the principal moments of inertia are not longer constant parameters, rather they depend on the
3
rotational state of the molecule.
The effects of centrifugal distortion are treated as a perturbation to ĤR so
that the resulting Hamiltonian becomes
Ĥ = ĤR + ĤCD
(1.9)
where ĤCD is the Hamiltonian accounting for centrifugal distortion. The firstorder perturbation correction to HˆR is given by
(1)
ĤCD =
~4 X
ταβγδ Jα Jβ Jγ Jδ
4
(1.10)
α,β,γ,δ
where α, β, γ, δ = a, b, or c. After some manipulation this becomes
ĤCD = −DJ J 4 − DJK J 2 Jz2 − Dk Jz4 + d1 J 2 (J+2 + J−2 ) + d2 (J+4 + J−4 )
(1.11)
where DJ , DJK , Dk , d1 , and d2 are the quartic centrifugal distortion constants,
which are functions of ταβγδ [5, 7]. It should be noted that this form of ĤCD is
general and depending on the molecule and therefore the choice of basis Jz = Ja ,
Jb , or Jc [5, 7]. For the work presented in subsequent chapters, ĤCD was chosen
to be in the symmetric rotor basis and therefore Jz was taken to be Ja or Jc
depending upon whether the basis set was prolate or oblate.
Additionally, ĤCD can be treated by higher levels of perturbation theory,
which would return higher-order centrifugal distortion coefficients. That being
said, only quartic centrifugal distortion terms are pertinent to the work that will
be presented in the subsequent chapters.
Lastly, the eigenstates of (1.9) are still labeled by the quantum numbers
J
KA KC .
1.1.3
Hyperfine Interactions
Many of the molecules to be discussed contain nuclei with intrinsic nuclear
spins, I (not be confused with the moments of inertia in this context), greater
than 12 . The presence of these nuclei in the molecular framework of a species can
4
give rise to many additional interactions. Nuclei with I >
1
2
have a magnetic
dipole moment and an electric quadrupole moment, both of which can further
complicate the rotational spectrum of a molecule.
1.1.4
Nuclear Quadrupole Interaction
An atomic nucleus with I >
1
2
has a non-spherical distribution of nuclear
charge and therefore a non-spherical distribution of electronic charge around the
nucleus. When such a nucleus is introduced into a molecular system, the nuclear
spin couples with the rotational angular momentum of the molecule causing perturbations and splittings in the rotational energy levels. The coupling scheme for
the nuclear quadrupole interaction is
F =I +J
(1.12)
where F is the total angular momentum. The new total angular momentum
quantum numbers then become
F = J + I, J + I − 1, ..., |J − I|
(1.13)
such that the rotational energy energy levels are labeled by the quantum numbers
J
KA KC
F . In the case of a species with two quadrupolar nuclei, with spins of
I1 and I2 , the coupling scheme becomes
F1 = I1 + J
(1.14)
F2 = I2 + F1
(1.15)
where the rotational energy levels are labeled by J
KA KC
F1 F2 . In this manner, n
quadrupolar nuclei could be accounted for (yet assigning the rotational spectrum
of such a species would be a different beast).
To first order, for cylindrically symmetric molecules, the nuclear quadrupole
interaction is given by
EQ = eQqJ
2J + 3
Y (J, I, F )
J
5
(1.16)
where
Y (J, I, F ) =
3
C(C
4
+ 1) − I(I + 1)J(J + 1)
2I(2I − 1)(2J − 1)(2J + 3)
(1.17)
with
C = F (F + 1) − J(J + 1) − I(I + 1)
(1.18)
is Casimir’s function (Y (J, I, F )), e is the charge of a proton, Q is the nuclear
quadrupole moment (due to non-spherical nuclear charge), and qJ is the electric
field gradient at the nucleus (due to non-spherical electronic charge around the
nucleus)[4–6]. In this general form
X
2
qJ =
qgg Jg2
(J + 1)(2J + 3)
(1.19)
2 Jg = J K A K C |Jg2 |J K A K C
(1.20)
g=a,b,c
where
[4–6]. However, as will be the case in many of the studies to come, the molecule
is not cylindrically symmetric (χbb 6= χcc ) and off-diagonal elements of the nuclear
quadrupole coupling Hamiltonian, ĤQ , are not negligible. It therefore becomes
necessary to also include the second order and higher corrections in perturbation
theory to EQ (or to not rely on perturbation theory at all and diagonalize the full
Hamiltonian). Upon this inclusion, ĤQ becomes
ĤQ =
X
1
χαβ [I , I ]
α β +
2I(2I − 1)
(1.21)
α,β
where α, β = a, b, or c, [Iα , Iβ ]+ = Iα Iβ + Iβ Iα , and χαβ are elements of the
traceless nuclear quadrupole coupling tensor

χ
χ
χ
 aa ab ac

χ =  χba χbb χbc

χca χcb χcc





(1.22)
such that the final form of this Hamiltonian is[4–6, 8]
ĤQ =
3
1
1
1
{ χaa [Ia2 − I 2 ] + (χbb − χcc )[I+2 + I−2 ] + χab [Ia Ib + Ib Ia ]
2I(2I − 1) 2
3
4
+χac [Ia Ic + Ic Ia ] + χbc [Ib Ic + Ic Ib ]}.
(1.23)
6
In the case of two (or more) quadrupolar nuclei, the quadrupolar interaction can be treated in two ways: a) if I1 > I2 then ĤQ2 is treated as a perturbation
to ĤQ1 or b) if I1 is of equal or similar magnitude to I2 then the characteristic
equation must be solved. The latter case is considerably more complex and was
characterized in-depth by Robinson et al. and Meyers et al.[9, 10].
1.1.5
Nuclear Spin-Rotation Interaction
Nuclei with I ≥
1
2
exhibit (often subtle) magnetic hyperfine interactions
(nuclear spin-rotation interactions) that further perturb the rotational energy
levels of a system. Classically, the interaction between a dipole moment and a
magnetic field is given by
ĤM = −µ· H.
(1.24)
In this context, the dipole moment, µ, is the nuclear spin magnetic moment µI ,
which is given by the nuclear magneton, βI , the gyromagnetic ratio of the nucleus,
gI , and the nuclear spin, I, such that
µ = µI = βI gI I
(1.25)
and H is the magnetic field created via molecular rotation. As the frequency of
molecular rotation is so much greater than the the precession of the nuclear spin,
µI is approximated to interact only with the average of the magnetic field generated by the angular momentum of the molecule in the direction of J . Considering
this
HJ J
Hef f = p
J(J + 1)
(1.26)
and
ĤSR = −βI gI I· Hef f
−βI gI HJ
= p
I· J .
J(J + 1)
(1.27)
After some manipulation, for the most general rotor,
ĤSR = CJ KA KC I· J
7
(1.28)
where
CJ KA KC
X
1
Cgg Jg2 .
=
J(J + 1)
(1.29)
g=a,b,c
It is important to note here that the CJ KA KC ’s depend on a specific rotational state
and during spectral analysis process, these are the values used to calculate the
Cgg ’s. Again after some rearrangement, to first-order, the nuclear-spin rotation
Hamiltonian is given by
ĤSR = Caa Ia Ja + Cbb Ib Jb + Ccc Ic Jc
(1.30)
where Caa , Cbb , and Ccc are the nuclear spin-rotation coupling constants respective
to the a-, b-, and c-principal axes[4–6]. In the case of a species that has more
than one nuclei with I ≥ 21 ,
ĤSR total = ĤSR 1 + ĤSR 2 + ... + ĤSR n .
(1.31)
It should be noted that no additional quantum numbers are necessary to account
specifically for the nuclear spin-rotation interaction, as it is merely a perturbative
interaction. Additionally, this outline of the nuclear spin-rotation interaction is
P
specifically for molecules in 1
states.
1.1.6
The Full Hamiltonian
The full Hamiltonian, for a species that exhibits all of the aforementioned
interactions, as will be the case for the molecules to be discusses, then has the
form
Ĥ = AJa2 + BJb2 + CJc2 − DJ J 4 − DJK J 2 Jz2 − Dk Jz4 + d1 J 2 (J+2 + J−2 )
+d2 (J+4 + J−4 ) +
1
1
3
{ χaa [Ia2 − I 2 ]
2I(2I − 1) 2
3
1
+ (χbb − χcc )[I+2 + I−2 ] + χab [Ia Ib + Ib Ia ] (1.32)
4
+χac [Ia Ic + Ic Ia ] + χbc [Ib Ic + Ic Ib ]}
+Caa Ia Ja + Cbb Ib Jb + Ccc Ic Jc .
8
References
[1] J. E. Wollrab, Rotational Spectra and Molecular Structure: Physical Chemistry: a Series of Monographs, volume 13, Academic Press, 2016.
[2] R. G. Mortimer, Physical chemistry. 3rd, 2008.
[3] D. T. Greenwood, Classical dynamics, Courier Corporation, 1977.
[4] H. W. Kroto, Molecular Rotation Spectra, Dover, 1992.
[5] W. Gordy, R. L. Cook, Microwave Molecular Spectra, Wiley, New York,
1984.
[6] C. H. Townes, A. L. Schawlow, Microwave Spectroscopy, Courier Corporation, 2013.
[7] J. K. Watson, The Journal of Chemical Physics 46 (1967) 1935–1949.
[8] E. Hirota, J. M. Brown, J. Hougen, T. Shida, N. Hirota, Pure Appl. Chem.
66 (1994) 571–576.
[9] G. W. Robinson, C. Cornwell, The Journal of Chemical Physics 21 (1953)
1436–1442.
[10] R. J. Myers, W. D. Gwinn, The Journal of Chemical Physics 20 (1952) 1420–
1427.
9
Chapter 2
A Study of 2-Iodobutane by
Rotational Spectroscopy
This chapter has been published in the Journal of Physical Chemistry A
prior to the compilation of this thesis. The author list is as follows: Eric A. Arsenault, Daniel A. Obenchain, Yoon Jeong Choi, Thomas A. Blake, S. A. Cooke,
and Stewart E. Novick. The publication is E. A. Arsenault, D. A. Obenchain, Y.
J. Choi, T. A. Blake, S. A. Cooke, S. E. Novick, J. Phys. Chem. A 120 (2016)
7145–7151.
2.1
Abstract
The rotational transitions belonging to 2-iodobutane (sec-butyl-iodide,
CH3 CHICH2 CH3 ) have been measured over the frequency range 5.5-16.5 GHz
via jet-pulsed Fourier transform microwave (FTMW) spectroscopy. The complete nuclear quadrupole coupling tensor of iodine, χ, has been obtained for the
gauche (g)-, anti (a)-, and gauche0 (g0 )-conformers, as well as the four
13
C iso-
topologues of the gauche species. Rotational constants, centrifugal distortion
constants, quadrupole coupling constants, and nuclear spin-rotation constants
were determined for each species. Changes in the χ of the iodine nucleus, resulting from conformational and isotopic differences, will be discussed. Isotopic
10
substitution of g-2-iodobutane allowed for a rs structure to be determined for the
carbon backbone. Additionally, isotopic substitution, in conjunction with an ab
initio structure, allowed for a fit of various r0 structural parameters belonging to
g-2-iodobutane.
2.2
Introduction
The electrophilic addition of hydrogen halides to alkenes is a fundamental
two-step reaction discussed in most introductory organic texts[1]: first a carbonium ion is formed with the transfer of the proton from the hydrogen halide to
the alkene, and second the halide ion bonds to the carbonium ion. Where on the
carbon backbone the halide ion attaches is dictated by the relative stability of
the carbonium ion formed in the first step. An ion formed at a tertiary carbon
position is more stable than that at a secondary position, which in turn is more
stable than an ion formed at a primary carbon position. These relative stabilities
lead to Markovnikov’s empirical rule explaining the preponderance of one product
isomer formed relative to another in hydrogen halide additions to alkenes. The
overall energetics of these reactions are further complicated by conformational
equilibria.[1, 2] Our interest is in examining the structures of these haloalkane
conformers and, eventually, their relationship to the carbonium ion transition
state.
Halobutanes represent a sufficiently complex, yet tractable, conformer system to study using free-jet supersonic cooling and Fourier transform microwave
spectroscopy. We begin here by examining the conformers of 2-iodobutane. Iodine, with its large nuclear quadrupole moment and possible spin-rotation coupling, makes the resulting microwave spectrum more complex and challenging to
assign and fit, but it affords the opportunity of gathering more detailed information about the molecule’s electronic wave function parameters.[3]
The first spectroscopic investigation of 2-iodobutane belonged to a larger
infrared (IR) study of 2-haloalkanes carried out by Benedetti and Cecchi[4] over
11
40 years ago. The IR spectra for each haloalkane revealed the presence of three
conformational isomers, related via a rotation about the C2 −C3 bond of the four
carbon chain. The conformations of 2-iodobutane, found in this work, agree
with the conclusions of Benedetti and Cecchi. Additionally, the
13
C isotopo-
logues of g-2-iodobutane were observed in natural abundance, leading to a more
rigorous determination of the structure of this particular conformation. The nuclear quadrupole coupling constants (NQCCs) belonging to the spectroscopically
observed 2-iodobutane species in the present work will be compared with each
other and to previously investigated iodoalkanes, in order to gain insight into
how substitution affects the electronic environment at the nucleus of the iodine
atom.
2.3
2.3.1
Experimental
Instrumentation
The initial rotational spectrum of 2-iodobutane was collected over a fre-
quency range of 7-13 GHz, shown in Figure 2.1, on a chirp-pulsed Fourier transform microwave (FTMW) spectrometer. This instrument is based on the design
by Pate and coworkers[5] and has been outlined in detail elsewhere[6]. The circuitry of this instrument will be described briefly. An 8, 10, or 12 GHz microwave
center frequency, ν, is mixed with a 6 µs linear frequency sweep, from DC to
1 GHz. The successive microwave radiation, ν ± 1 GHz, is amplified and then
broadcast through a microwave horn antenna into the molecular beam. The radiation then induces a polarization in the coincident supersonic expansion of the
gas-phase molecular sample. Following a 1 µs delay, a second microwave horn
antenna collects the free induction decay (FID). The FID is fast Fourier transformed and directly digitized on a Tektronix TDS6124C Digital Oscilloscope.
800,000 points of the FID are collected over a time period of 20 µs, where one
point is obtained every 25 ps. Measured rotational transitions have an average
12
line width of 80 kHz with an uncertainty of ± 8 kHz in the center frequency.
Transitions for all four 13 C isotopologues, as well as ancillary transitions for
all of the conformers, were measured with a Balle-Flygare type spectrometer[7].
This instrument has also been formerly described in detail[8, 9]. In short, microwave radiation, lasting 0.9 µs, polarizes the sample concurrently undergoing
supersonic expansion. After a delay of 26 µs, the FID of the polarized sample
is collected for 102.4 µs and digitized. Between a few hundred and a few thousand averages were collected for each molecular transition, in order to improve
the signal-to-noise ratio. The measured transitions have an average line width of
5 kHz with an uncertainty of ± 3 kHz in the center frequency.
The sample was purchased from Sigma-Aldrich (≥ 98% CH3 CHICH2 CH3 )
and used without further purification. The volatile liquid sample (boiling point
119-120 ◦ C) was pipetted into a glass U-form tube containing copper beads as a
stabilizer. One atmosphere of dry argon (99.999%, Airgas) was bubbled through
the sample. This mixture was then pulsed through a solenoid valve into the
chamber of the spectrometer, which is held at a pressure of ∼ 10−6 Torr. The
molecules in the gas pulse undergo supersonic expansion, leaving them rotationally cold (1-2 K) and in only the lowest energy conformations, but yet, g-, a-, and
g0 -conformers were observed. Sample treatment, as described above, was identical
for both instruments.
13
Figure 2.1: The experimental rotational spectrum of 2-iodobutane is shown above the baseline
and the simulated spectra of the various conformers are shown below. This portion of spectrum
illustrates the heavy overlap of the hyperfine structure due to the iodine nucleus in each of
the three conformers. The predicted gauche-, anti-, and gauche0 -2-iodobutane transitions are
represented by purple, red, and green, respectively. Theoretical transition intensities have been
scaled using the relative ab initio energies.
2.3.2
Quantum Chemical Calculations
Figure 2.2: Illustrations of the three ab initio structures. The C1 −C2 −C3 −C4 dihedral angles
for the a-, g0 -, and g-conformers are 64◦ , -62◦ , and 171◦ , respectively.
Quantum chemical calculations were performed using the GAUSSIAN09
Revision A suite[10] to obtain ab initio structures for CH3 CHICH2 CH3 . A coordinate scan at the APFD/321G* level was first employed, in order to determine
likely ground state molecular geometries. From this scan, multiple low-energy
structures were found. However, only two structures from this calculation were
observed in the spectra, which have been labeled as the g- and g0 -conformers.
The subsequent structures were optimized at the MP2 level of theory using a
14
6-311G* basis set for the iodine atom, which was imported from the EMSL Basis
Set Library[11, 12] and a 6-311G++(2d,2p) basis set for the remaining hydrogen
and carbon atoms. An additional calculation for the a-conformer, not obtained
in the coordinate scan, was performed at the same level of theory as the other
optimizations. Zero point energy (ZPE) corrections were calculated for each optimized structure. Results from the ZPE corrections did not contain any imaginary
frequencies, which indicates that these three structures are all true local minima
on the potential energy surface. Images of these three ab initio structures are
presented in Figure 2.2 and the results of the calculation are presented in Table
2.1. The rotational constants, belonging to the a-, g0 -, and g-conformers, are in
basic agreement with the experimental results.
Table 2.1: Ab initio results of 2-iodobutane at the MP2 level of theory
Parameters
a
g
g0
A (MHz)
6072
3612
4273
B (MHz)
1160
1645
1479
C (MHz)
1014
1189
1248
χaa a (MHz)
-812
-684
-650
χbb (MHz)
395
272
287
χcc (MHz)
417
411
363
χab (MHz)
-256
-441
-416
χac (MHz)
-207
-212
-311
χbc (MHz)
-43
-85
-118
∆Eb (cm−1 )
166
0
210
∆EZP E c (cm−1 )
174
0
237
Dihedral Angle (◦ )
64
171
-62
a
NQCCs resulting from the presence of iodine.
b
Energies relative to lowest energy conformer.
c
Energies relative to lowest energy conformer with ZPE corrections.
15
2.3.3
Spectral Assignments
The microwave assignments for the three conformers and all four
13
C
isotopologues of gauche-2-iodobutane were completed with the aid of Pickett’s
SPFIT/SPCAT[13] software. All of the conformers were fit initially from broadband spectra gathered on the chirp-pulse FTMW instrument, where the AABS
package[14] was used jointly with Pickett’s programs. Tables 2.2 and 2.3 contain
the final spectroscopic constants for the three conformations observed and for the
gauche species and its four
13
C isotopologues, respectively. It can be noted that
there is approximately a 4% difference between all of the ab initio and experimental rotational constants. The ab initio NQC tensors of iodine are all quite a
bit different than the experimental results. The only agreement is in the trend
in relative magnitude of the diagonal and off-diagonal NQCCs. The rather poor
prediction of the NQC tensor is due to the fact that a core potential was used to
estimate the interaction energies of the many core electrons of iodine. For consistency, the experimentally determined off-diagonal NQCCs are presented with
signs in agreement with the ab initio values. However, the only certainty that
exists about the signs of the off-diagonal terms is that the product of the three
off-diagonal elements must be negative[15].
The g-conformer, an asymmetric top with κ = −0.62, contains a dipole
moment that projects onto the a-, b-, and c-principal axes. Both the a- and g0 species are near prolate, with κ = −0.94 and κ = −0.85, and also contain a
dipole moment with three projections in the principal axis system. Due to the
asymmetry and large quadrupolar nuclei present in each species, a rich variety
of transition types were observed. Spectral assignments for the a-conformer were
based on Q- and R- branch transitions. For the g0 - species, Q-, R-, and S-branch
transitions (∆J=+2) were also observed. The fit for the g-conformer included
P-, Q-, R-, and S-branch transitions. Figure 2.3 presents an a-type S-branch
transition belonging to g-2-iodobutane.
The assignments for the four gauche-13 C isotopologues, all present in nat-
16
ural abundance, were based primarily on a-type R-branch transitions. A few
b-type R-branch transitions and a S-branch transition, also included in the spectral fits for each species, helped significantly to better determine the A rotational
constant. All four of these fits consist of fifteen parameters. However, six parameters, namely the centrifugal distortion constants, DK , d1 , and d2 , and the
nuclear spin-molecular rotation constants, Caa , Cbb , and Ccc , were held constant
to the parent values. These terms were unable to be fit from the small number
of observed transitions belonging to the lowest energy conformation, the g-13 C
isotopologues.
13
C isotopologues belonging to the two other conformers were not
assigned. The transitions belonging to these species were both lacking in intensity
and heavily tangled in a slew of other transitions.
17
Figure 2.3: The Doppler doublet of the a-type S-branch transition, 404
3
2
← 221
1
2,
of the
parent conformer of g-2-iodobutane is shown, where the average of each peak is taken to be the
transition frequency, 11011.621 MHz. Labels on the y-axis of the plot were omitted, since the
intensity of the transition is arbitrary. The transition was measured on a Balle-Flygare type
spectrometer with 100 averages, with a signal-to-noise ratio of 76.
2.3.4
Hyperfine Structure
The I =
5
2
nuclear spin of iodine, results in the presence of hyperfine struc-
ture in the rotational spectrum of 2-iodobutane. The observed hyperfine structure
is a result of nuclear quadrupole and spin-rotation coupling. The Hamiltonian
that accounts for these complications is of the form[16–18]:
18
Ĥ = ĤR + ĤCD + ĤQ + ĤSR .
(2.1)
ĤR and ĤCD are the Hamiltonian terms accounting for molecular rotation
and centrifugal distortion, respectively. ĤQ is the nuclear quadrupole coupling
Hamiltonian, which can be written[19]:
X
1
χαβ [I , I ]
ĤQ =
α β +
2I(2I − 1)
(2.2)
α,β
and then, with some manipulation, this can be written in a form appropriate for use with Pickett’s SPFIT/SPCAT[13]:
ĤQ =
1
3
1
1
{ χaa [Ia2 − I 2 ] + (χbb − χcc )[I+2 + I−2 ] + χab [Ia Ib + Ib Ia ]
2I(2I − 1) 2
3
4
+χac [Ia Ic + Ic Ia ] + χbc [Ib Ic + Ic Ib ]}
(2.3)
where the χij terms correspond to the components of the nuclear electric
quadrupole coupling tensor. ĤSR , the Hamiltonian that accounts for nuclear
spin-rotation coupling can be expanded as[20]:
ĤSR = Caa Ia Ja + Cbb Ib Jb + Ccc Ic Jc
(2.4)
where Cii are the diagonal nuclear spin-rotation constants. Rotational
0
0
00
00
transitions are labeled by quantum numbers of the form JK
← JK
0
0F
00 00 F ,
a Kc
a Kc
where F is the total angular momentum quantum number that includes the coupling of spin angular momentum with the rotational angular momentum of the
molecule, given by F = I + J .
19
Table 2.2: Spectroscopic parameters of three conformations of 2-iodobutane
Experimental
a
ab initio
Parameters
a
g
g0
a
g
g0
A (MHz)
6276.1041(8)a
3726.51649(11)
4433.8638(7)
6072
3612
4273
B (MHz)
1200.61256(18)
1706.60968(12)
1520.53128(26)
1160
1645
1479
C (MHz)
1049.42041(11)
1231.19182(7)
1287.89606(38)
1014
1189
1248
DJ (kHz)
0.0908(16)
0.3488(14)
0.3546(29)
–
–
–
DJK (kHz)
–
–
-0.618(18)
–
–
–
DK (kHz)
-6.12(25)
1.034(7)
4.54(7)
–
–
–
d1 (kHz)
-0.0141(11)
-0.1256(10)
-0.088(4)
–
–
–
d2 (kHz)
–
-0.01970(35)
–
–
–
–
χaa b (MHz)
-1550.634(13)
-1329.651(2)
-1256.176(6)
-812
-684
-650
χbb (MHz)
779.401(12)
582.497(14)
569.007(11)
395
272
287
χcc (MHz)
771.233(17)
747.154(14)
687.169(12)
417
411
363
χab c (MHz)
-497.892(35)
-822.076(20)
-792.17(4)
-256
-441
-416
χac (MHz)
-452.170(28)
-456.84(5)
-615.34(6)
-207
-212
-311
χbc (MHz)
-92.25(7)
-176.840(32)
-227.51(4)
-43
-85
-118
Caa (kHz)
4.0(7)
2.85(11)
2.7(5)
–
–
–
Cbb (kHz)
3.74(22)
4.69(12)
4.18(30)
–
–
–
Ccc (kHz)
4.47(24)
3.65(7)
3.52(27)
–
–
–
Nd
102
212
72
–
–
–
RMSe (kHz)
3.8
2.7
3.0
–
–
–
Numbers in parentheses give standard errors (1σ, 67% confidence level)
in units of the least significant figure.
b NQCCs
c The
resulting from the presence of iodine.
relative signs of the off-diagonal NQCCs can not be determined. It is only known that the product of the
three, χab , χac , and χbc , must be negative. These terms are presented with signs in agreement with the ab
initio results.
d Number
e Root
of transitions used in the fit. r
P h
mean square deviation of the fit,
i
(obs − calc)2 /N .
20
Table 2.3: Spectroscopic parameters for g-2-iodobutane
Parameters
Predictiona
g-Parent
13
A (MHz)
3612
3726.51649(11)b
3606.32964(26)
3710.83914(30)
3722.6838(8)
3659.4255(8)
B (MHz)
1645
1706.60968(12)
1697.90412(9)
1697.83972(12)
1677.48112(18)
1674.57492(14)
C (MHz)
1189
1231.19182(7)
1213.32896(9)
1225.72244(13)
1216.26864(20)
1207.29040(14)
0.3478(24)
C1
13
C2
13
C3
13
C4
DJ (kHz)
–
0.3488(14)
0.3386(15)
0.3441(20)
0.3385(32)
DK (kHz)
–
1.034(7)
[1.034]g
[1.034]
[1.034]
[1.034]
d1 (kHz)
–
-0.1256(10)
[-0.1256]
[-0.1256]
[-0.1256]
[-0.1256]
d2 (kHz)
–
-0.01970(35)
[-0.0197]
[-0.0197]
[-0.0197]
[-0.0197]
χaa c (MHz)
-684
-1329.651(2)
-1356.097(13)
-1339.391(6)
-1323.515(23)
-1291.649(17)
χbb (MHz)
272
582.497(14)
609.264(14)
589.897(11)
578.842(27)
544.173(20)
χcc (MHz)
411
747.154(14)
746.833(19)
749.494(13)
744.672(36)
747.476(27)
χab d (MHz)
-441
-822.076(20)
-789.72(13)
-814.05(12)
-826.10(33)
-864.80(20)
χac (MHz)
-212
-456.84(5)
-460.40(31)
-452.34(31)
-461.4(7)
-451.6(5)
χbc (MHz)
-85
-176.840(32)
-169.526(30)
-173.04(17)
-180.07(44)
-187.00(28)
Caa (kHz)
–
2.85(11)
[2.85]
[2.85]
[2.85]
[2.85]
Cbb (kHz)
–
4.69(12)
[4.69]
[4.69]
[4.69]
[4.69]
Ccc (kHz)
–
3.65(7)
[3.65]
[3.65]
[3.65]
[3.65]
Ne
–
212
39
58
38
40
f
–
2.7
0.8
1.7
1.6
1.2
RMS (kHz)
a
From MP2 level calculation.
b
Numbers in parentheses give standard errors (1σ, 67% confidence level)
in units of the least significant figure.
c
NQCCs resulting from the presence of iodine.
d
The relative signs of the off-diagonal NQCCs can not be determined. It is only known that
the product of the three, χab , χac , and χbc , must be negative. These terms are presented with
signs in agreement with the ab initio results.
e
f
g
Number of transitions used in the fit. r
P h
i 2
Root mean square deviation of the fit,
(obs − calc) /N .
Numbers in square brackets indicate values held constant to those obtained for parent.
21
2.4
2.4.1
Discussion
Structural Determination
The five unique sets of rotational constants for g-2-iodobutane, from the
parent species and four 13 C isotopologues, allowed for structural determination via
isotopic substitution. With the aid of the STRFIT structural fitting program[21],
eight selected geometric parameters were obtained using these fifteen spectroscopic rotational constants, where ab initio coordinates served as an initial structure. Table 2.4 lists the four bond lengths, two bond angles, and two dihedral
angles that give the structure of the carbon backbone and C−I bond. The C1 C2 -C3 -C4 dihedral angle was determined to be 172.7(16)◦ , thus indicating that
the carbon chain is non-planar. The r0 coordinates obtained from this structural
fit were used in later calculations, namely when performing a rotation of the
quadrupole tensor into the C−I bond.
A Kraitchman analysis[22] yielded rs coordinates for the carbon chain,
with respect to the principal axes of g-2-iodobutane. This analysis was performed
to serve as a confirmation that the positions of the assigned carbons are correct.
A comparison between these experimentally derived coordinates and ab initio
coordinates is presented in Table 2.5. It should be noted that the c-coordinate
for C1 is an imaginary number. Although this coordinate is included in Table
2.5, it is best to assume that it is simply near zero. Table 2.5 also offers a
comparison between the Kraitchman analysis and ab initio results, which are in
good agreement.
22
Table 2.4: STRFIT r0 structural parameters of gauche-2-iodobutane
Bond Length (Å)
C1 -C2
1.536(12)a
C2 -C3
1.497(6)
C3 -C4
1.548(6)
I-C2
2.166(4)
Bond Angle (◦ )
∠(C1 -C2 -C3 )
112.9(10)
∠(C2 -C3 -C4 )
114.17(23)
Dihedral (◦ )
(C1 -C2 -C3 -C4 )
172.7(16)
(I-C2 -C3 -C4 )
-63.85(31)
Structural Fit Error
a
χ2
0.0024
σ
0.018
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
23
24
a
2.2422(7)
2.3616(6)
C3
C4
1.6019(9)
0.132(11)
0.6646(23)
2.1389(7)
|b|
0.113(13)
0.354(4)
0.370(4)
0.052(29)i
|c|
Values in parentheses give absolute Costain errors of the least significant figure.
1.1801(13)
C2
a
1.2170(12)
|a|
Kraitchman Coordinates
C1
Atoms
2.4
2.3
1.2
1.2
a
-1.6
-0.15
0.67
2.2
b
0.13
-0.36
0.37
-0.023
c
ab initio Gauche Coordinates
Table 2.5: Kraitchman versus ab initio coordinates for gauche-2-iodobutane
2.4.2
Nuclear Quadrupole Coupling Tensor of Iodine
Upon inspection of Tables 2.2 and 2.3, it is immediately obvious that the
elements of the NQC tensor are quite different for each species of 2-iodobutane.
These differences are a result of the different orientations of the principal inertial
axes in the conformations and isotopologues with respect to the C−I bond. In order to make a meaningful comparison of these tensors, they should all be expressed
in individual coordinate systems that are not dependent upon the conformations
or the isotopic variations. One such set of frames are those in which the χ tensors
themselves are diagonalized. Utilizing Kisiel’s program, QDIAG[23], the complete NQC tensor of iodine was diagonalized for each species of 2-iodobutane.
Diagonalization transforms the NQC tensor from the inertial axis system of the
molecule to the principal axis system of the quadrupolar nucleus, iodine, in this
case.
Table 2.6: Conformational comparison of the diagonalized NQC tensor of iodine in
2-iodobutane
a
Parameters
a
g
g0
χzz (MHz)
-1737.458(21)a
-1731.291(23)
-1730.09(4)
χyy (MHz)
881.337(18)
888.384(20)
862.95(4)
χxx (MHz)
856.121(20)
842.907(28)
867.14(5)
ηχ b
0.014513(16)
0.026268(20)
-0.00242(4)
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
b
ηχ is a measure of the asymmetry of the nuclear quadrupole coupling tensor, where
ηχ =
χxx −χyy
.
χzz
A comparison of the diagonalized quadrupole tensor, χ, for iodine in each
of the three observed conformations is presented in Table 2.6. Comparing values
of η χ , which is a measure of the asymmetry of the tensor, in Table 2.6 reveals
subtle changes in the electronic nature of the C−I bond, or more specifically,
25
changes in the electric field gradient at the nucleus of the iodine atom, due to
conformational differences. A NQC tensor with an η χ of 0 would indicate a
cylindrically symmetric tensor. Values of η χ for the a-, g-, and g0 -conformers
were 0.014513(16), 0.026268(20), and -0.00242(4), respectively. This parameter
indicates that the g0 -conformer is 6 times more “symmetric” than the a-conformer,
which is just under twice as “symmetric” as the g-conformer. Interestingly, the
order in increasing symmetry between χ for the three conformers follows the trend
in the increasing relative ab initio energies between the three conformers. The
affect of geometric changes on the NQC tensor can also be seen upon a comparison
of χzz between the three conformers. After comparing the final values of χzz
for each conformer, at most, only a 0.4% difference was observed. However,
more powerful conclusions can be drawn from the values of χzz obtained for
2-iodobutane upon contrasting this work with a series of previous studies on
iodoalkanes. Table 2.7 presents a selection of such work. There is a notable trend,
namely that the magnitude of χzz decreases with increasing carbon substitution
in the iodoalkane. This change in magnitude is more pronounced when the degree
of substitution increases on the carbon directly bonded to the iodine atom. More
simply put, |χzz | is less for an iodine atom bonded to a secondary carbon than it
is for an iodine atom bonded to a primary carbon. This diagonalization is only
possible because the complete tensor was determined, which, in turn, was only
possible because the iodine χ was large enough that even the off-diagonal terms
had spectroscopic consequences. There is only a 0.23% difference between χzz of
a-2-iodobutane and isopropyl iodide, which is just under three times less than the
difference between χzz of g0 -2-iodobutane and isopropyl iodide, 0.65%. This factor
of three can be rationalized by comparing geometric differences between these
species. The g0 -conformer simply differs more from isopropyl iodide geometrically
than the a-conformer does. These comparisons serve quite nicely to show the
sensitivity of the nuclear quadrupole of the iodine atom and its ability to serve
as a probe of subtle chemically relevant differences.
However, even more subtle changes in the NQC tensor of iodine can be
26
Table 2.7: Comparison of the diagonalized NQC tensor of iodine 2-iodobutane with other
iodoalkanes
Molecule
χzz (MHz)
Reference
CH3 I
-1934.080(10)
Wlodarczak et al. [24]
CH3 CH2 I
-1815.693(210)
Boucher et al. [25]
trans-CH3 CH2 CH2 I
-1814.55(55)
Fujitake and Hayashi [26]
gauche-CH3 CH2 CH2 I
-1805.16(56)
Fujitake and Hayashi [26]
CH3 CHICH3
-1741.47(75)
Ikeda et al. [27]
a−CH3 CH2 CHICH3
-1737.458(21)
This work
g−CH3 CH2 CHICH3
-1731.291(33)
This work
g0 -CH3 CH2 CHICH3
-1730.09(4)
This work
noticed by comparing χ of the parent species of g-2-iodobutane with its four
13
C
isotopologues. This can be seen in Table 2.8. A comparison of ηχ between these
five isotopic species reveals agreement, at least within range of their respective
errors.
13
C isotopic substitution seems to have no noticeable affect on the asym-
metry of the NQC tensor of iodine in g-2-iodobutane. Although there may be no
affect in this regard, the values of χzz seem to suggest a change in the projection
of the NQC tensor along the z-axis of the iodine nucleus when the
directly bonded to iodine is substituted with a
the difference in χzz between the parent and
13
13
12
C nucleus
C nucleus. When comparing
C2 isotopologue, the change in
χzz is 200 kHz. This is just over a factor of two greater than the largest change
observed for any of the other isotopologues, where the values of χzz between the
parent and
13
C1 ,
13
C3 , and
13
C4 isotopologues only vary by 40 to 80 kHz. Multi-
ple causes for this discrepancy were investigated in order to rationalize this very
small difference.
First, the presence of additional nuclear spin-rotation interactions, due to
the the magnetic moment of the
13
C nucleus, were investigated. However, the
27
Table 2.8: Rotation of the diagonalized NQC tensor of iodine into the principal axis system
of g-2-iodobutane
a
13
13
C1
13
C2
13
Parameters
g-Parent
C3
C4
χzz (MHz)
-1731.291(23)a
-1731.21(14)
-1731.49(13)
-1731.24(34)
-1731.25(23)
χyy (MHz)
888.384(20)
888.34(8)
888.61(9)
888.60(24)
888.51(16)
χxx (MHz)
842.907(28)
842.88(14)
842.89(14)
842.63(35)
842.74(25)
ηχ b
0.026268(20)
0.02626(9)
0.02640(10)
0.02655(25)
0.02644(17)
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
b
ηχ is a measure of the asymmetry of the nuclear quadrupole coupling tensor, where
ηχ =
χxx −χyy
.
χzz
inclusion of these terms in the Hamiltonian of this isotopologue both did not fit
or counterbalance the change in χzz of the
13
C2 isotopologue. Further evidence
against the presence of this interaction was provided by the fact that the other
three isotopic species yielded values of χzz nearly identical to the parent value
without the addition of these terms in their respective Hamiltonians.
Second, an additional spin-spin interaction between
(I = 52 ) may be present. No change in χzz of the
13
13
C (I = 12 ) and
127
I
C2 isotopologue ensued as a
result of including this term in the Hamiltonian. The term was deemed negligible,
as it did not fit well and was only on the order of 2 kHz.
After a missing term in the Hamiltonian of the
13
C2 isotopologue was
eliminated as a cause for this change, the discrepancy in χzz , between the parent
species and this isotopologue, can be attributed to the most obvious effect, the
increased mass of the carbon atom directly bonded to the iodine. Upon
13
C
isotopic substitution, the vibrationally averaged C−I bond length should decrease.
The fact that the z-axis of the diagonalized nuclear electronic quadrupole tensor
is under 2◦ from the C−I bond axis certainly helped to make such an observation
possible. A further ab initio study was performed, to determine the normal mode
vibration frequencies of g-2-iodobutane. The bond length was scaled using the
normal mode frequency that is closest to the “pure” C−I stretch, 599 cm−1 , and
28
the approximation that the reduced mass of the stretch was the reduced mass of
12
C − 127 I or
13
C − 127 I. A decrease in bond length of 1.8 mÅ was found, which
was used to scale the
13
C − 127 I bond length. An energy calculation was then
performed on this slightly altered structure, at the same level of theory as all
previous quantum chemical calculations, to predict the change in the NQCCs.
Table 2.9 presents χ for each of the two species in question with more significant
figures than are typically presented because the comparison being made is only
among ab initio values, so the errors in the magnitudes cancels out to first order.
A decrease in χzz of 300 kHz was determined for a bond length decrease of 1.8
mÅ. Thus our measured decrease of 199(13) kHz suggests a C−I bond length
decrease of approximately 1.2 mÅ upon
13
C isotopic substitution.
Table 2.9: Changes in the ab initio NQC tensor due to isotopic substitution
Parent
2.5
13
C Bond Length Correction
χzz (MHz)
-896.4
-896.7
χyy (MHz)
444.5
444.7
χxx (MHz)
451.9
452.1
Conclusion
Utilizing high-resolution rotational spectroscopy, an intensive investigation
of three conformers and four
13
C isotopologues of 2-iodobutane allowed for many
chemically relevant parameters of this haloalkane to be determined. The observed
hyperfine structure led to the complete determination of the NQC tensor of iodine
in seven different species of 2-iodobutane. In this way, iodine served as a probe of
the subtle differences in these species, resulting from both isotopic and geometric
differences.
29
2.6
Acknowledgements
The authors thank Wallace (Pete) Pringle for many useful discussions.
The cluster at Wesleyan University is supported by the NSF under CNS-0619508.
The Pacific Northwest National Laboratory is operated for the United States
Department of Energy by the Battelle Memorial Institute under contract DEAC05-76RLO 1830.
2.7
Supplemental Information
Final fit outputs for the a-, g-, and g0 -parent species, in addition to the four
13
C isotopologues of the g0 -conformer, can be found at doi:10.1021/acs.jpca.6b06938.
References
[1] R. T. Morrison, R. N. Boyd, Organic Chemistry, 3 ed., Allyn and Bacon,
Inc., Boston, 1973.
[2] W. E. Steinmetz, F. Hickernell, I. K. Mun, L. H. Scharpen, J. Mol. Spectrosc.
68 (1977) 173–182.
[3] J. Gripp, H. Dreizler, Z. Naturforsch., A: Phys. Sci. 43 (1988) 971–976.
[4] E. Benedetti, P. Cecchi, Spectrochim. Acta, Part A 28 (1972) 1007–1017.
[5] G. G. Brown, B. C. Dian, K. O. Douglass, S. M. Geyer, S. T. Shipman, B. H.
Pate, Rev. Sci. Instrum. 79 (2008) 053103.
[6] G. S. Grubbs II, C. T. Dewberry, K. C. Etchison, K. E. Kerr, S. A. Cooke,
Rev. Sci. Instrum. 78 (2007) 096106.
[7] T. Balle, W. Flygare, Rev. Sci. Instrum. 52 (1981) 33–45.
[8] G. Grubbs II, D. A. Obenchain, H. M. Pickett, S. E. Novick, J. Chem. Phys.
141 (2014) 114306.
30
[9] A. H. Walker, W. Chen, S. E. Novick, B. D. Bean, M. D. Marshall, J. Chem.
Phys. 102 (1995) 7298–7305.
[10] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb,
J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov,
J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota,
R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao,
H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro,
M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov,
R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S.
Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox,
J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski,
R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador,
J. J. Dannenberg, S. Dapprich, A. D. Daniels, Farkas, J. B. Foresman, J. V.
Ortiz, J. Cioslowski, D. J. Fox, Gaussian-09 revision d.01, 2013. Gaussian
Inc. Wallingford, CT 2009.
[11] D. Feller, J. Comput. Chem. 17 (1996) 1571–1586.
[12] K. L. Schuchardt, B. T. Didier, T. Elsethagen, L. Sun, V. Gurumoorthi,
J. Chase, J. Li, T. L. Windus, J. Chem. Inf. Model. 47 (2007) 1045–1052.
[13] H. M. Pickett, J. Mol. Spectrosc. 148 (1991) 371–377.
[14] Z. Kisiel, L. Pszczólkowski, I. R. Medvedev, M. Winnewisser, F. C. De Lucia,
E. Herbst, J. Mol. Spectrosc. 233 (2005) 231–243.
[15] U. Spoerel, H. Dreizler, W. Stahl, Zeitschrift für Naturforschung A 49 (1994)
645–646.
[16] P. Thaddeus, L. Krisher, J. Loubser, J. Chem. Phys. 40 (1964) 257–273.
[17] D. Posener, Austr. J. Phys. 11 (1958) 1–17.
31
[18] D. Boucher, J. Burie, D. Dangoisse, J. Demaison, A. Dubrulle, J. Chem.
Phys. 29 (1978) 323–330.
[19] E. Hirota, J. M. Brown, J. Hougen, T. Shida, N. Hirota, Pure Appl. Chem.
66 (1994) 571–576.
[20] W. Gordy, R. L. Cook, Microwave Molecular Spectra, Wiley, New York,
1984.
[21] Z. Kisiel, J. Mol. Spectrosc. 218 (2003) 58–67.
[22] J. Kraitchman, Am. J. Phys. 21 (1953) 17–24.
[23] Z. Kisiel, Prospe–programs for rotational spectroscopy, 2000.
[24] G. Wlodarczak, D. Boucher, R. Bocquet, J. Demaison, J. Mol. Spectrosc.
124 (1987) 53–65.
[25] D. Boucher, A. Dubrulle, J. Demaison, J. Mol. Spectrosc. 84 (1980) 375–387.
[26] M. Fujitake, M. Hayashi, J. Mol. Spectrosc. 127 (1988) 112–124.
[27] C. Ikeda, T. Inagusa, M. Hayashi, J. Mol. Spectrosc. 135 (1989) 334–348.
32
Chapter 3
A Study of the Conformational
Isomerism of 1-Iodobutane by
High Resolution Rotational
Spectroscopy
This chapter has been published in the Journal of Molecular Spectroscopy
prior to the compilation of this thesis. The author list is as follows: Eric A.
Arsenault, Daniel A. Obenchain, Thomas A. Blake, S. A. Cooke, and Stewart E.
Novick. The publication is E. A. Arsenault, D. A. Obenchain, T. A. Blake, S. A.
Cooke, S. E. Novick, J. Mol. Spec. (2017). doi:10.1016/j.jms.2017.03.014.
3.1
Abstract
The first microwave study of 1-iodobutane, performed by Steinmetz et
al. in 1977, led to the determination of the B + C parameter for the antianti- and gauche-anti-conformers. Nearly 40 years later, this reinvestigation of
1-iodobutane, by high-resolution microwave spectroscopy, led to the determination of rotational constants, centrifugal distortion constants, nuclear quadrupole
coupling constants (NQCCs), and nuclear-spin rotation constants belonging to
33
both of the two previously mentioned conformers, in addition to the gauchegauche-conformer, which was observed in this frequency regime for the first time.
Comparisons between the three conformers of 1-iodobutane and other iodo- and
bromoalkanes are made, specifically through an analysis of the nuclear quadrupole
coupling constants belonging to the iodine and bromine atoms in the respective
chemical environments.
3.2
Introduction
It has been commonly accepted that the electric field gradient of a nu-
cleus remains unchanged in most simple molecules even as they take on different
conformations or as they form van der Waals complexes. A field gradient caused
by a charge is proportional to 1/r3 , where r is the distance from the charge to
the nucleus. It is presented by Townes and Dailey[1] and outlined in Gordy and
Cook[2], how the field gradient found at a nucleus can be assumed to have been
produced by the bonds made by that atom. This analysis was originally for bonding p orbitals, but was later extended to include hybrid orbital contributions by
Novick in 2011[3]. This unchanging nature of a field gradient was recently used
to correctly determine the structure of HOD−N2 O[4].
The assumption that the field gradient will remain unchanged upon the
formation of a complex, conformational change, or isotopic substitution has, of
course, exceptions. Choosing N2 O and the complexes it forms as an example[5–
16], it was shown repeatedly that there are sublte changes in the electronic environment of near atoms as complexes are formed. For HCCH· · · N2 O[5, 7], as
studied by Leung and coworkers, it was shown that in forming the complex there
is a significant change in the electric field gradient of the central nitrogen, while
the field gradient at terminal nitrogen remained unchanged. Through molecular
multipole analysis, the authors showed that this change is caused by redistribution
of electrons about the central nitrogen.
To observe changes in electronic structure from perturbations smaller in
34
magnitude than those observed upon forming a van der Waals complex, such as
conformational changes, a more sensitive nucleus is required to act as a probe for
this change. Iodine, with its large nuclear electric quadrupole moment, −69.6(12)
f m2 [17] for
127
I, compared to the
14
N value of 2.001(10) f m2 [18], makes the
observed nuclear quadrupole coupling constants significantly more sensitive to
small changes in electron distribution near the iodine nucleus. For example, in
a recent study of iodobenzene and the Ne-iodobenzene complex[19], there is a
< 0.3% change in the iodine NQCC upon forming the complex with neon. We
have recently reported a small, but significant, change in electronic structure from
carbon-13 isotopic substitution in 2-iodobutane[20].
Continuing a series of studies on the subtle changes in electric structure determined by changes in nuclear quadrupole coupling constants, we report
here on the conformational effects on a terminal iodine group in a hydrocarbon
chain.
3.3
Experimental
The high-resolution rotational spectrum of 1-iodobutane was measured
from 7-13 GHz with a chirped-pulse Fourier transform microwave (FTMW) spectrometer. Detailed specifications of this spectrometer, which is based on the
design of Pate and coworkers[21], have been presented previously[22]. In short,
a chosen microwave center frequency, ν, and a 6 µs linear frequency sweep are
mixed from DC to 1 GHz. The resulting radiation, ν ± 1 GHz, is broadcast
directly into a vacuum chamber through a microwave horn antenna. This transmitted radiation then induces a polarization in the coincident molecular beam.
A second microwave horn antenna collects the free induction decay (FID) after a
1 µs delay. With the aid of a Tektronix TDS6124C Digital Oscilloscope, the FID
is fast Fourier transformed and directly digitized. A total of 800,000 points, over
a time of 20 µs, are collected from the FID. On average, the molecular rotational
transitions have a line width of 80 kHz with an uncertainty of |8| kHz in the center
35
frequency.
R
The sample was acquired from Sigma-Aldrich
(≥ 99% CIH2 CH2 CH2 CH3 ).
Further purification was not necessary. The sample (bp 130-131 ◦ C) was contained in a glass U-form tube at room temperature and one atmosphere of dry
R
argon (99.999%, Airgas
) was bubbled directly through the liquid. The final
mixture of carrier gas and sample was pulsed through a solenoid valve into the
chamber, held at an ambient pressure of 10−6 Torr, allowing the molecules to
undergo supersonic expansion, where the molecules become rotationally cold (1-2
K).
3.3.1
Quantum Chemical Calculations
Using a 321G* basis set at the APFD level of theory, a coordinate scan of
the C−C−C−C and C−C−C−I dihedral angles was performed, in order to identify the most probable ground state molecular geometries. Three of the lowest
energy structures from the scan were optimized at the MP2 level of theory. A
6311G* basis set was imported from the EMSL Basis Set Library[23, 24] specifically chosen to handle the iodine atom, while a 6311G++(2d,2p) basis set was
used for the remaining carbon and hydrogen atoms. All calculations were performed with the GAUSSIAN09 Revision D suite[25]. Results from the ab initio
optimizations are presented in Table 3.1. Illustrations of the corresponding structures can be found in Figure 3.1. More on predicting the NQCCs will be discussed
in the subsequent sections.
3.3.2
Spectral Assignments
All three of the lowest energy conformers obtained from the ab initio in-
vestigation were successfully assigned from the rotational spectrum collected in
the frequency range of 7-13 GHz. A 70 MHz portion of this spectrum is shown in
Figure 3.2. The three broadband assignments were accomplished with the help of
both the AABS package[26] and Pickett’s programs, SPFIT/SPCAT[27, 28]. The
36
Table 3.1: Rotational constants and NQCCs for three conformers of 1-iodobutane as
determined from ab initio optimization at the MP2 level of theory
Parameters
gg
ga
aa
A (MHz)
6033
7388
15123
B (MHz)
1077
937
706
C (MHz)
1013
867
686
-397
-274
-682
χbb (MHz)
74
-167
223
χcc (MHz)
322
441
459
χab (MHz)
-570
-670
-511
χac (MHz)
-330
-109
0b
χbc (MHz)
-226
-104
0
∆Ea (cm−1 )
385
160
0
127 I
χaa (MHz)
127 I
127 I
127 I
127 I
127 I
a
Relative energies from MP2 optimizations.
b
χac and χbc are zero by symmetry.
Figure 3.1:
Calculated structures of the anti-anti (aa)-, gauche-anti (ga)-, and gauche-
gauche (gg)-conformers of 1-iodobutane from an ab initio optimization. The C−C−C−C and
C−C−C−I dihedral angles for the aa-, ga-, and gg-species were calculated to be 180◦ , 180◦ ;
179◦ , 66◦ ; and -65◦ , -63◦ .
37
Figure 3.2: A small portion of the experimental spectrum of 1-iodobutane is shown in black,
with the final simulated spectra for the aa-, ga-, and gg-conformers shown below in green,
purple, and red, respectively. The quantum numbers associated with each rotational transition
00
0
00
0
← JK
are also presented above the experimental spectrum, in the form JK
00
00 F .
0
0F
a Kc
a Kc
Figure 3.3: A 100 MHz portion of three different predictions of the rotational spectrum of
aa-1-iodobutane. The top orange spectrum is based on a hybrid tensor (discussed in Spectral
Assignments) and ab initio rotational constants, the middle blue spectrum is the prediction
using the experimental results belonging to the aa-conformer, and the bottom orange spectrum
is based completely on ab initio results.
38
Table 3.2: Spectroscopic parameters of 1-iodobutane
a
Parameters
gg
ga
aa
A (MHz)
5917.262(13)a
7532.3121(18)
15052.043(9)
B (MHz)
1162.2040(14)
970.51215(26)
732.93002(18)
C (MHz)
1082.7921(12)
896.60780(18) 711.34567(20)
DJ (kHz)
0.816(10)
0.2058(24)
0.0459(11)
DK (kHz)
34.9(38)
45.8(4)
–
DJK (kHz)
-7.18(5)
-4.450(18)
-2.867(17)
d1 (kHz)
-0.133(12)
-0.0302(6)
–
d2 (kHz)
–
0.0122(14)
–
χaa (MHz)
-752.593(28)
-521.973(12)
-1294.041(34)
χbb (MHz)
157.01(5)
-337.242(15)
380.05(6)
χcc (MHz)
595.58(6)
859.216(19)
913.99(7)
χab b (MHz)
-1116.74(13)
-1331.251(23)
-1070.28(6)
χac (MHz)
-692.17(19)
-226.66(13)
0e
χbc (MHz)
-460.66(12)
-235.95(6)
0
Caa (kHz)
–
6.4(15)
15(4)
Cbb (kHz)
3.43(7)
0.89(27)
1.8(8)
Ccc (kHz)
–
1.66(26)
2.7(8)
Nc
74
198
136
RMSd (kHz)
6.4
7.3
6.4
Numbers in parentheses give standard errors (1σ, 67% confidence level)
in units of the least significant figure.
b The
signs of the off-diagonal nuclear quadrupole coupling constants can not be exactly determined. It is only
known that the product of the three, χab , χac , and χbc , must be negative. These terms are presented with
signs in accordance with the ab initio results.
c Number
d Root
eχ
ac
of transitions used in the fit. r
P h
i
(obs − calc)2 /N .
mean square deviation of the fit,
and χbc are zero by symmetry.
39
final rotational constants, centrifugal distortion constants, NQCCs, and nuclear
spin-rotation constants for each conformer can be found in Table 3.2. Although
there is a large discrepancy between the ab initio and experimental NQC tensors
belonging to the iodine atom in each of the three species, there is at most only a
7% difference between the rotational constants obtained from the ab initio study
and those determined experimentally. This seems to suggest that the actual geometries of the three conformers present in the molecular beam are quite similar
to those that were calculated, whereas the only agreement in the NQCCs was in
the trend of their respective magnitudes.
The poor ab initio NQCC values made the first assignment, of the gaconformer, quite challenging. In order to make the assignment process more
efficient and less tedious, alternate methods of prediction were employed. Taking advantage of the fact that changes in the geometry of the butane chain have
a very small effect on the electric field gradient at the iodine nucleus (more on
this to come later), the experimental NQC tensor of iodine in the ga-conformer
was used to make predictions of the NQC tensors of iodine in both of the two
remaining unassigned conformers. This predictive process was simply an exercise
in tensor rotation. Using QDIAG[29], the ab initio tensors of the two unassigned
conformers were diagonalized. In doing this, the rotation matrices, specific to
the ab initio geometries of these species, were obtained. Then, these respective
rotation matrices were used to transform the experimental NQC tensor of the
ga-conformer into the inertial axes systems of the gg- and aa-conformers. The
resulting hybrid tensors were based on ab initio geometries, specific to the unassigned conformers, and experimental NQCCs, belonging to the already assigned
ga-conformer. To best illustrate the power of this method, the NQC tensor predictions of the aa-conformer will used as an example. Equation (1) contains the ab
initio tensor belonging to the aa-conformer, Equation (2) contains the hybrid tensor, and Equation (3) contains the experimental tensor based only on rotational
transitions belonging to the aa-conformer. It is perhaps immediately obvious that
the hybrid tensor, again based on the ab initio geometry of the aa-conformer and
40
the rotated experimental NQC tensor of the ga-conformer, offers a much better
prediction than the purely ab initio tensor. To further highlight this point, Figure 3.3 presents small portions of the rotational spectra corresponding to these
respective tensors. It is worth noting here that the similarities between the NQC
tensors of iodine in various iodoalkanes, such as iodoethane, t-1-iodopropane, and
aa-1-iodobutane, which will be discussed later, suggest that this method can be
applied to quite a large range of problems. This method, based on straightforward linear algebra, can lead to very accurate predictions without the need for
expensive computations.
It should be noted there are other methods used to make accurate predictions of NQQCs, especially for species with large quadrupoles, as shown by
Professor W. C. Bailey[30]. Post-analysis, we received calculations from Professor Bailey, which can be found in Table 3.3 alongside the predictions from our
hybrid approach[31, 32]. These calculations involve the calibration of a specific
combination of level of theory and basis set, by linear regression, of the calculated electric field gradients versus the experimental NQCCs of a selected group
of molecules. Once calibrated, the specific combination of level of theory and
basis set can be applied to other systems. Upon comparison of Table 3.2 and
Table 3.3, it can be seen that this method yields excellent results. The differences
between the hybrid method explained previously and these calculations are small.
Although it is clear that these calculations are quite accurate, the hybrid method
can rapidly provide good predictions based on very computationally inexpensive
optimizations.

−682 −511
0





χM P 2 =  −511 223
0 


0
0
459


−1346 −1019 0




χhybrid =  −1019 462
0 


0
0
884
41
(3.1)
(3.2)

χexp
−1294.041(34) −1070.28(6)
0


=  −1070.28(6)
380.05(6)
0

0
0
913.99(7)





(3.3)
Table 3.3: Comparison of the two methods of prediction for the NQC tensors in 1-iodobutane
MP2/6311+G(d,p)a
gg
ga
aa
gg
gab
aa
-757
-546
-1317
-706
–
-1346
χbb (MHz)
164
-314
400
104
–
462
χcc (MHz)
593
860
917
602
–
884
χab (MHz)
-1119
-1336
-1062
-1025
–
-1019
χac (MHz)
-702
-234
0c
-632
–
0c
χbc (MHz)
-464
-238
0
-367
–
0
Parameters
127 I
χaa (MHz)
127 I
127 I
127 I
127 I
127 I
a
Hybrid Method
These calculations were performed by Professor W. C. Bailey with the calibrated
MP2/6311+G(d,p) combination[31, 32].
b
These values were not predicted via the hybrid method, as the experimental NQC tensor of
the ga-conformer was used to make the predictions for the other two conformers. Rather, the
NQC tensor presented in Table 1 was used.
c
Zero by symmetry.
3.3.3
Theory
The hyperfine structure in the rotational spectrum of 1-iodobutane is a
consequence of the nuclear spin of iodine (I = 52 ), which allows for the observation
of both nuclear quadrupole coupling and nuclear spin-rotation coupling. The
respective Hamiltonians, which account for these interactions, were combined
with both the rigid rotor and centrifugal distortion Hamiltonians. The final form
of the Hamiltonian then becomes[20, 33–35]:
Ĥ = ĤR + ĤCD + ĤQ + ĤSR .
42
(3.4)
Quantum labels for the rotational transitions belonging to 1-iodobutane
00
0
00
0
are as follows: JK
0 F ← JK 00 K 00 F , where F = I + J . The quantum number F
0
a c
a Kc
is the total angular momentum quantum number that accounts for the coupling
between the nuclear spin of iodine and the rotational angular momentum of the
molecule.
3.4
Discussion
Table 3.4: Comparison of this work with the previous study of 1-iodobutane
Parameters
a
This work.
Steinmetz et al.[36]
ga B + C (MHz)
1867.11995(32)
1868.4(30)a
aa B + C (MHz)
1444.27569(27)
1445.3(4)
gg B + C (MHz)
2244.9961(18)
Not observed.
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
For the first time since 1977, 1-iodobutane was reinvestigated via microwave spectroscopy. Table 3.4 presents a comparison between this study and
the work of Steinmetz et al.[36] In the previous low-resolution microwave study,
the B + C parameter was measured for the ga- and aa-conformers. Between studies, the measured values of B + C are within 0.07% for both the ga-conformer
and aa-conformer. In this study, the gg-conformer was spectroscopically detected
in this frequency range for the first time, although previous work[36] did indicate
that this species should likely be present in low abundance, as the highest energy
conformer of the three.
Due to the flexible nature of this molecule, the three observed conformers
possess quite different rotational constants. This is most evident upon noting
that the difference between the A rotational constant of the heavy atom planar aa-conformer and the highly asymmetric ga-conformer is over 9100 MHz, as
seen in Table 3.2. As expected, the large structural differences between these
43
Table 3.5: NQC tensor of iodine in 1-iodobutane
Parameters
gg
ga
aa
Energies (cm−1 )a
385
160
0
χzz (MHz)
-1758.57(14)b
-1804.133(35)
-1815.72(6)
-3.15%c
-0.64%
–
1182.40(19)
920.04(9)
901.73(7)
+31.13%
+2.03%
–
576.17(16)
884.09(11)
913.99(7)
-36.96%
-3.27%
–
0.34473(15)
0.01993(8)
-0.00675(6)
χyy (MHz)
χxx (MHz)
ηχ d
a
From Table 1.
b
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
c
Percentage change relative to the corresponding component of the lowest energy
aa-conformer.
d
ηχ is a measure of the asymmetry of the nuclear quadrupole coupling tensor, where
ηχ =
χxx −χyy
.
χzz
44
Table 3.6: NQC tensor of bromine in 1-bromobutane[37]
Parameters
79 Br
79 Br
79 Br
gg
ga
aa
χzz (MHz)
536.2(5)a
539.7(6)
543.45(34)
χyy (MHz)
-264.9(6)
-264.9(20)
-269.5(6)
χxx (MHz)
-271.29(36)
-274.8(15)
-274.0(9)
-0.0119(13)
-0.018(5)
-0.0083(21)
χzz (MHz)
448.3(47)
451.1(9)
453.95(35)
χyy (MHz)
-222.0(19)
-220.2(23)
-225.5(7)
χxx (MHz)
-226.3(39)
-231.0(19)
-228.4(9)
-0.01(1)
-0.024(7)
-0.0065(24)
ηχ b
81 Br
81 Br
81 Br
ηχ
a
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
b
ηχ is a measure of the asymmetry of the nuclear quadrupole coupling tensor, where
ηχ =
χxx −χyy
.
χzz
three conformational isomers, resulting from differences in the I−C−C−C and
C−C−C−C dihedral angles, do not translate to substantial variations in the
chemical environment at the iodine nucleus. The most meaningful quantification
of the differences between these chemical environments or more specifically, the
electric field gradient at the iodine nucleus, can be made through a comparison
of the various diagonalized nuclear quadrupole coupling tensors, χ, where χ is
a projection of the NQC tensor in the principal axis system of the quadrupolar
nucleus (e.g. iodine), as opposed to a projection of the tensor in the inertial axis
system of the molecule. The advantage of comparing these diagonalized tensors
is that they are all projections into an axis system which is independent of the
conformation. Table 3.5 presents the diagonalized χ tensor of iodine in each conformer. Diagonalization was performed with QDIAG[29]. It should be noted that
the χzz element is the best means of comparison because it is the projection of
χ onto the z-axis of the nuclear quadrupole, which is pointed nearly along the
45
Table 3.7: A comparison of this work with similar haloalkanes
127 I
127 I
127 I
iodoethane[38]
t-1-iodopropane[39]
aa-1-iodobutane
χzz (MHz)
-1815.22(85)a
-1814.55(55)
-1815.72(6)
χyy (MHz)
901.71(81)
900.44(47)
901.73(7)
χxx (MHz)
913.50(26)
914.12(44)
913.99(7)
-0.0065(6)
-0.0075(5)
-0.00675(6)
bromoethane[38]
t-1-bromopropane[40]
aa-1-bromobutane[37]
χzz (MHz)
544.03(168)
541.6(7)
543.45(34)
χyy (MHz)
-270.32(166)
-267.8(7)
-269.5(6)
χxx (MHz)
-273.71(17)
-273.79(4)
-274.0(9)
-0.0062(34)
-0.0111(12)
-0.0083(21)
bromoethane[38]
t-1-bromopropane[40]
aa-1-bromobutane[37]
χzz (MHz)
453.91(203)
451.9(7)
453.95(35)
χyy (MHz)
-225.25(201)
223.1(7)
-225.5(7)
χxx (MHz)
-228.65(18)
-228.77(4)
-228.4(9)
-0.0075(48)
-0.0125(16)
-0.0065(24)
ηχ b
79 Br
79 Br
79 Br
ηχ
81 Br
81 Br
81 Br
ηχ
a Numbers
in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
b
ηχ is a measure of the asymmetry of the nuclear quadrupole coupling tensor, where
ηχ =
χxx −χyy
.
χzz
46
C−I bond, whereas the orientation of the x- and y-axis are not known. While the
differences in the rotated NQC tensor elements of the three conformers are small,
they are never the less significant, as can be seen in Table 3.5. Upon inspection,
only a 3.15% difference between χzz of the gg- and aa-species is observed, even
though the geometry of the alkane chain is drastically different in each case. Even
smaller differences in χzz , of 2.56% and 0.64%, are present when comparing the
remaining two pairs of conformers, namely the gg- versus the ga-species, and the
ga- versus the aa-species, respectively. This is purely a reflection of the fact that
the two latter pairs of conformers possess more similar geometries than the first
pair that was mentioned. This indicates that there is hardly a change in the
electric field gradient at the iodine nucleus upon rather significant conformational
changes.
Similar comparisons, based on a study by Kim et al., between the χzz
element of χ belonging to both 79 Br and 81 Br in the gg-, ga-, and aa-species of 1bromobutane were made[37]. These tensors are presented in Table 3.6. At most,
a 0.7% difference between the
81
Br gg- and aa-species was determined, which is
just over four and a half times less than the percent difference found between the
iodine gg- and aa-species. The greater differences between the χzz elements of the
1-iodobutane conformers is as expected and can be explained simply by the fact
that iodine is both larger and more polarizable than bromine[41].
Additional comparisons were made between progressively longer haloalkanes, namely iodoethane, t-1-iodopropane, aa-1-iodobutane, and the corresponding
bromine analogs. These NQCCs are presented in Table 3.7. In order to complete
Table 3.7, the NQCCs belonging to t-1-bromopropane were measured by highresolution microwave spectroscopy because these values could not be found in the
previous work on this species by Sarachman[42]. Table 3.7 shows clearly that
the NQC tensors of iodine in each of these progressively longer haloalkanes are
identical, within experimental error. Unsurprisingly, additional terminal carbon
atoms seem to have a very small effect on the electric field gradient at the iodine nucleus. Upon investigating the differences in the bromine analogs, the same
47
conclusion can be realized. Although this does not necessarily present itself as a
surprise, one curious trend did emerge. If closer attention is payed to Table 3.7,
it can be seen that although the tensors amongst alkanes with identical halogen
substituents are the same, within experimental error, the tensors belonging to
iodine,
79
Br, and
81
Br are in better agreement between the ethane and butane
chains than between the ethane and propane chains or the propane and butane
chains. However, this trend remains inconclusive because the NQCCs belonging to
35
Cl and
37
Cl in aa-1-chlorobutane have yet to be obtained. Additionally,
NQC tensors belonging to longer iodine- and bromine-containing alkanes have
also yet to be measured, due to the fact that collecting and measuring spectra for
progressively longer haloalkanes becomes more and more challenging.
3.5
Conclusion
The reinvestigation of 1-iodobutane by high-resolution microwave spec-
troscopy in the frequency range of 7-13 GHz led to the determination of rotational constants, centrifugal distortion constants, nuclear quadrupole coupling
constants, and nuclear-spin rotation constants for three low energy conformations. The full NQC tensor of iodine in each conformation was obtained, which
allowed for comparisons to be made between the chemical environments of these
conformers, as well as other similar haloalkanes.
3.6
Acknowledgments
The authors thank Professor Wallace (Pete) Pringle for many useful dis-
cussions and Professor W. C. Bailey for the calculations that he shared. This work
was supported at Wesleyan University by NSF grant CHE-1565276. The cluster at
Wesleyan University is supported by the NSF under CNS-0619508. The Pacific
Northwest National Laboratory is operated for the United States Department
of Energy by the Battelle Memorial Institute under contract DE-AC05-76RLO
48
1830.
3.7
Supplemental Material
The final fit outputs for gg-, ga-, and aa-1-iodobutane are provided at
doi:10.1016/j.jms.2017.03.014.
References
[1] C. Townes, B. Dailey, J. Chem. Phys. 17 (1949) 782–796.
[2] W. Gordy, R. L. Cook, Microwave Molecular Spectra, 1984, Wiley, New
York, 1984.
[3] S. E. Novick, J. Mol. Spectrosc. 267 (2011) 13–18.
[4] D. A. Obenchain, D. S. Frank, S. E. Novick, W. Klemperer, J. Chem. Phys.
143 (2015) 084301.
[5] H. O. Leung, J. Chem. Phys. 107 (1997) 2232–2241.
[6] H. O. Leung, A. M. Osowski, O. A. Oyeyemi, J. Chem. Phys. 114 (2001)
4829–4836.
[7] H. O. Leung, Chem. Commun. (1996) 2525–2526.
[8] H. O. Leung, D. Gangwani, J.-U. Grabow, J. Mol. Spectrosc. 184 (1997)
106–112.
[9] H. O. Leung, J. Chem. Phys. 110 (1999) 4394–4401. doi:10.1063/1.3517494.
[10] H. O. Leung, O. M. Ibidapo, P. I. Abruña, M. B. Bianchi, J. Mol. Spectrosc.
222 (2003) 3–14.
[11] H. O. Leung, J. Chem. Phys. 108 (1998) 3955–3961.
[12] M. D. Marshall, H. O. Leung, J. Mol. Spectrosc. 196 (1999) 149–153.
49
[13] H. O. Leung, W. T. Cashion, K. K. Duncan, C. L. Hagan, S. Joo, J. Chem.
Phys. 121 (2004) 237–247.
[14] M. S. Ngarı, W. Jäger, J. Mol. Spectrosc. 192 (1998) 320–330.
[15] M. S. Ngarı, W. Jäger, J. Chem. Phys. 111 (1999) 3919–3928.
[16] N. R. Walker, A. J. Minei, S. E. Novick, A. C. Legon, J. Mol. Spectrosc. 251
(2008) 153–158.
[17] P. Pyykkö, Mol. Phys. 106 (2008) 1965–1974.
[18] IUPAC, Quantities, units and symbols in physical chemistry, 2007.
[19] J. L. Neill, S. T. Shipman, L. Alvarez-Valtierra, A. Lesarri, Z. Kisiel, B. H.
Pate, J. Mol. Spectrosc. 269 (2011) 21–29.
[20] E. A. Arsenault, D. A. Obenchain, Y. J. Choi, T. A. Blake, S. A. Cooke,
S. E. Novick, J. Phys. Chem. A 120 (2016) 7145–7151.
[21] G. G. Brown, B. C. Dian, K. O. Douglass, S. M. Geyer, S. T. Shipman, B. H.
Pate, Rev. Sci. Instrum. 79 (2008) 053103.
[22] G. S. Grubbs II, C. T. Dewberry, K. C. Etchison, K. E. Kerr, S. A. Cooke,
Rev. Sci. Instrum. 78 (2007) 096106.
[23] D. Feller, J. Comput. Chem. 17 (1996) 1571–1586.
[24] K. L. Schuchardt, B. T. Didier, T. Elsethagen, L. Sun, V. Gurumoorthi,
J. Chase, J. Li, T. L. Windus, J. Chem. Inf. Model. 47 (2007) 1045–1052.
[25] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb,
J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov,
J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota,
R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao,
H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro,
M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov,
50
R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S.
Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox,
J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski,
R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador,
J. J. Dannenberg, S. Dapprich, A. D. Daniels, Farkas, J. B. Foresman, J. V.
Ortiz, J. Cioslowski, D. J. Fox, Gaussian-09 revision d.01, 2013. Gaussian
Inc. Wallingford, CT 2009.
[26] Z. Kisiel, L. Pszczólkowski, I. R. Medvedev, M. Winnewisser, F. C.
De Lucia, E. Herbst, J. Mol. Spectrosc. 233 (2005) 231–243. URL:
http://www.sciencedirect.com/science/article/pii/S0022285205001554.
doi:10.1016/j.jms.2005.07.006.
[27] H.
M.
Pickett,
J.
Mol.
Spectrosc.
148
(1991)
371–377.
URL:
http://www.sciencedirect.com/science/article/pii/002228529190393O.
doi:10.1016/0022-2852(91)90393-O.
[28] S. E. Novick, J. Mol. Spectrosc. 329 (2016) 1–7.
[29] Z. Kisiel, PROSPE–Programs for ROtational SPEctroscopy, 2000.
[30] W. C. Bailey, Calculation of nuclear quadrupole coupling constants in
gaseous state molecules, http://nqcc.wcbailey.net/, 2010.
[31] W. C. Bailey, Nuclear quadrupole coupling constants in 1-iodobutane,
http://nqcc.wcbailey.net/, 2016.
[32] W. C. Bailey, Private communication, 2016.
[33] P. Thaddeus, L. Krisher, J. Loubser, J. Chem. Phys. 40 (1964) 257–273.
[34] D. Posener, Austr. J. Phys. 11 (1958) 1–17.
[35] D. Boucher, J. Burie, D. Dangoisse, J. Demaison, A. Dubrulle, J. Chem.
Phys. 29 (1978) 323–330.
51
[36] W. E. Steinmetz, F. Hickernell, I. K. Mun, L. H. Scharpen, J. Mol. Spectrosc.
68 (1977) 173–182.
[37] J. Kim, H. Jang, S. Ka, D. A. Obenchain, R. A. Peebles, S. A. Peebles, J. J.
Oh, J. Mol. Spectrosc. 328 (2016) 50–58.
[38] T. Inagusa, M. Fujitake, M. Hayashi, J. Mol. Spectrosc. 128 (1988) 456–468.
[39] M. Fujitake, M. Hayashi, J. Mol. Spectrosc. 127 (1988) 112–124.
[40] E. A. Arsenault, S. E. Novick, Private communication, 2016.
[41] E. V. Anslyn, D. A. Dougherty, Modern Physical Organic Chemistry, University Science Books, 2006.
[42] T. Sarachman, in: International Symposium on Molecular Spectroscopy,
Ohio State University, 1968, p. N7.
52
Chapter 4
Nuclear Quadrupole Coupling in
SiH2I2 due to the Presence of
Two Iodine Nuclei
4.1
Abstract
The rotational spectrum of diiodosilane was measured with a jet-pulsed,
cavity Fourier transform microwave (FTMW) spectrometer over the frequency
range 8.8 GHz to 15 GHz and assigned for the first time. The complete nuclear quadrupole coupling (NQC) tensors for both iodine nuclei were obtained for
the
28
Si,
29
Si, and
30
Si isotopologues of diiodosilane. In addition to the nuclear
quadrupole coupling constants (NQCCs), rotational constants, centrifugal distortion constants, and nuclear spin-rotation constants were determined for each
silicon isotopologue. Subtle, yet unmistakable, changes in the NQCCs of iodine
upon isotopic substitution were observed and will be discussed. A r0 structure of
diiodosilane was also fit via isotopic substitution, which led to the determination
of bond lengths and angles: r0 (Si−I) = 2.4236(19) Å, r0 (Si−H) = 1.475(21) Å,
∠(I−Si−I) = 111.27(13)◦ , and ∠(I−Si−H) = 105.9(19)◦ . These molecular parameters will be compared to the results of a previous gas electron diffraction
53
study.
4.2
Introduction
Halosilanes of the form SiY4−n Xn , with Y = HF ; X = ClBrI ; n = 2,
3, remain vastly understudied by microwave, millimeter wave, and submillimeter
wave spectroscopy, especially when compared to the carbon analogs, CY4−n Xn , of
these species. To date, five halosilanes that meet this criteria have been studied,
none of which include even one iodine atom[1–11]. However, over the same time
period, thirteen halomethanes with at least two quadrupole-containing halogen
substituents (these include: Cl, Br, and I[12]) have been the subject of over 50
studies by similar spectroscopic techniques[13–69].
Additionally, out of all eighteen of the previously mentioned halosilanes
and halomethanes, only a select few were found to be the focus of spectroscopic
studies in the terahertz, far infrared, and infrared frequency regimes[70–76].
Perhaps more notably, fewer than half a dozen molecules that contain
both a silicon and an iodine atom have been studied by microwave spectroscopy
throughout the 64 years previous to the work presented in this paper[77–86].
This work aims to fill that gap. The molecule that is the focus of this
paper, diiodosilane, is the first halosilane with two iodine substituents and the
third molecule with two iodine substituents, the first being diiodomethane[66]
and the second being difluorodiiodomethane[69], to be studied by high-resolution
microwave spectroscopy.
High-resolution microwave studies of iodine-containing silanes allow for
the measurement of complete
127
I nuclear quadrupole coupling (NQC) tensors.
These tensors contain an abundance of information about the Si−I bond, as
will be discussed throughout the course of this paper. One method of deducing
the chemical nature of the Si−I bond through these tensors is the Townes-Dailey
analysis[87–89], the implications of which will be examined for the mono-, di-, and
tetraiodosilane series. Alongside this analysis, twelve spectroscopic parameters for
54
each silicon isotopologue of diiodosilane and a r0 structure for this molecule will
be presented.
4.3
Experiment
R
The diiodosilane sample (Sigma-Aldrich
) was held in a glass U-form tube
R
and 760 torr of dry argon (99.999 %, Airgas
) was bubbled through the liquid.
The resulting mixture was pulsed through a solenoid valve into the chamber of a
Balle-Flygare type spectrometer, which freatures a Fabry-Perot cavity[90, 91]. A
microwave pulse of 0.9 µs duration then polarized the sample. Following a 26 µs
delay, the free induction decay (FID) was collected for 102.4 µs and digitized. A
few hundred to a few thousand averages were collected for each each rotational
transition, in order to achieve a satisfactory signal-to-noise ratio. The average
line width of a transition is 5 kHz and the uncertainty in the center frequency is
|3| kHz. Further details on this instrument can be found elsewhere[92, 93].
4.4
Spectral Assignments
In order to assign the complex hyperfine structure belonging to SiH2 I2 , an
amalgam of predictive techniques, including scaling, quantum chemical calculations, and tensor rotations were employed. Rotational constants, based on the
r0 gas phase structure obtained by Altabef et al.[94], were calculated using the
Gaussian 09 Revision A suite[95] at the MP2 level of theory. A 6-311G* basis
set, imported from the EMSL Basis Set Library[96, 97], was used for the iodine
atom and a 6-311G++(2d,2p) basis set was used for the remaining silicon and
hydrogen atoms.
Scaled NQCCs were predicted using the ratio presented in equation (1).
It is important to note here that scaling was performed with the elements of
the diagonalized
127
I NQC tensors, χ, in each of the three species, CH3 I, CH2 I2 ,
and SiH3 I, to ensure that the tensors were all independent of specific molecular
55
geometries[98]. Table 4.1 presents the values used for scaling, as well as the
resulting prediction of the NQC tensor of iodine in SiH2 I2 .
χCH3 I
χSiH3 I
=
χCH2 I2
χSiH2 I2
(4.1)
Table 4.1: Results obtained from equation (4.1)
CH3 I[99]
CH2 I2 [66]
SiH3 I[80]b
SiH2 I2 a
χzz (MHz)
-1934.1306(51)
-2030.1(5)
-1245.1
-1306.9
χyy (MHz)
967.0635(36)
1036.66(9)
622.55
667.3
χxx (MHz)
967.0635(36)
993.4(10)
622.55
600.9
Parameters
127 I
127 I
127 I
a
Method of prediction described in detail in section on Spectral Assignments. These are not
ab initio predictions.
b
No error reported.
The scaled NQC tensor was then projected into the inertial axis frame of
the r0 gas phase structure by using the rotation matrix obtained through diagonalization of the NQC tensor from the MP2 optimization. Diagonalization was
performed with QDIAG[100]. The results of this prediction can be found in Table
4.2.
Spectral assignments, in the frequency range of 8.8 GHz to 15 GHz, were
completed with the AABS package[101] and Pickett’s programs, SPFIT/SPCAT[102,
103]. The coupling scheme used to assign the hyperfine structure was F1 = I1 + J
and F2 = I2 + F1 where I1 and I2 are both 52 , the nuclear spin of iodine, and
0
0 0
00
00 00
quantum labels took on the form: JK
← JK
0
0F F
00 00 F1 F2 . Rotational cona Kc 1 2
a Kc
stants, centrifugal distortion constants, nuclear quadrupole coupling constants
(χij where i, j = a, b, c), and nuclear-spin rotation constants (Cij where i, j =
a, b, c) were fit for each of the three silicon isotopologues of diiodosilane. An
additional term, χK
aa , was included in the fit to handle K-dependent centrifugal
distortion in χaa . As a result of this small perturbation, the term χaa ef f = χaa
+ K2 χK
aa replaced χaa in the nuclear quadrupole coupling Hamiltonian[66, 104].
Table 4.2 contains the final collection of these spectroscopic parameters. In total,
56
Table 4.2: Spectroscopic parameters of three silicon isotopologues of diiodosilane
Parameters
Prediction
28
SiI2 H2
Experimental
28
29
SiI2 H2
30
SiI2 H2
SiI2 H2
A (MHz)
8550a
8574.52263(20)b
8363.96765(18)
8165.3310(48)
B (MHz)
499
496.1275(5)
496.1410(7)
496.1493(14)
C (MHz)
474
471.42779(34)
470.7846(6)
470.1462(8)
DJ (kHz)
–
0.0362(10)
0.0343(9)
[0.0343]c
DJK (kHz)
–
-3.96(17)
-3.86(28)
[-3.86]
-773e
-693.754(1)
-693.777(1)
-693.791(9)
127
I
χaa d (MHz)
127
I
χbb (MHz)
133
36.871(1)
36.890(1)
36.908(7)
127
I
χcc (MHz)
601
656.883(1)
656.887(2)
656.884(11)
|χab | (MHz)
877
856.2242(19)
856.2266(17)
856.237(11)
–
20.1(7)
19.6(7)
[19.6]
Caa f (kHz)
–
4.60(9)
4.37(9)
[4.37]
Cbb (kHz)
–
0.567(31)
0.43(4)
[0.43]
Ccc (kHz)
–
0.906(21)
0.845(29)
[0.845]
Ng
–
189
147
27
RMSh (kHz)
–
0.5
0.5
0.6
127
I
127
I K
χaa
(kHz)
a
Calculated rotational constants based on GED study[94].
b
Numbers in parentheses give standard errors (1σ, 67% confidence level)
in units of the least significant figure.
c
Brackets denote that the term was held constant to the value measured for
29
d
χaa , χbb , χcc , χab , and χK
aa are identical for each iodine nucleus. However, the sign of χab
SiI2 H2 .
can not be exactly determined. It is only known that χab (127 I1 ) = -χab (127 I2 )[64, 66].
e
Method of prediction described in detail in section on Spectral Assignments.
f
Caa , Cbb , and Ccc are identical for each iodine nucleus.
g
Number of transitions used in the fit. r
P h
i 2
h
Root mean square deviation of the fit,
(obs − calc) /N .
57
363 b-type R-branch rotational transitions were measured for the three silicon isotopologues of diiodosilane. Of the measured transitions, 189 belonged to the
isotopologue (92.21% natural abundance), 147 belonged to the
(4.70% natural abundance) and only 27 belonged to the
30
29
28
Si
Si isotopologue
Si isotopologue (3.09%
natural abundance)[105]. Although fewer transitions were measured for the
30
Si
isotopologue, due mainly to low natural abundance, enough were measured such
that the rotational constants and the 127 I NQC tensor could be determined, which
were necessary for the purposes of structural analysis and NQC tensor comparison. However, not enough transitions were measured to determine the centrifugal
distortion constants or nuclear spin-rotation constants for
30
SiH2 I2 , so in the fit,
these parameters were held constant to those measured for the most similar isotopologue, in terms of mass,
29
SiH2 I2 .
In Table 4.2, the results of the predicted NQC tensor of iodine in the
28
Si isotopologue are provided next to the experimentally determined tensor of
this species. The predicted tensor elements, χaa , χcc , and χab , are all within
less than 6% of the experimental values. However, χbb was predicted to be three
and a half times larger in magnitude than the final spectroscopic value. This
discrepancy is attributed to small imperfections in the rotation matrix, which
was used to transform χpredicted into the principal axis system of the predicted
structure (based on molecular parameters from the GED study[94]).
The predicted NQC tensor is in even better agreement with experimental
results when the respective diagonalized NQC tensors are compared. Of course,
this makes sense as, once again, the diagonalized tensors are independent of specific molecular geometries. Table 4.3 contains the experimental NQC tensor,
diagonalized with QDIAG[100], and the predicted NQC tensor of iodine in the
28
Si isotopologue. At most, there is only a 1.8% difference between the elements
of χ in the predicted tensor versus the experimental tensor.
58
Table 4.3: NQC tensor of iodine in diiodosilane
Parameters
Prediction
28
127 I
127 I
127 I
a
SiH2 I2 a
Experimental
28
SiH2 I2
29
SiH2 I2
30
SiH2 I2
χzz (MHz)
-1306.9
-1259.3406(20)b
-1259.3530(20)
-1259.367(7)
χyy (MHz)
667.3
656.8830(10)
656.8870(20)
656.8840(11)
χxx (MHz)
600.9
602.4576(20)
602.4660(20)
602.484(6)
ηχ c
0.05
0.0432174(18)
0.0432134(22)
0.043196(5)
θzb d (◦ )
–
56.55
56.55
56.55
Method of prediction described in detail in section on Spectral Assignments. These are not
ab initio predictions.
b
Numbers in parentheses give standard errors (1σ, 67% confidence level) in units of the least
significant figure.
c
ηχ is a measure of the asymmetry of the nuclear quadrupole coupling tensor, where
ηχ =
d
χxx −χyy
.
χzz
Angle between the quadrupolar z-axis and the inertial b-axis.
59
4.5
4.5.1
Discussion
Structural Determination
Utilizing STRFIT (STRucture FITting program)[106], an r0 structure of
diiodosilane was determined via isotopic substitution. Using the twelve available
experimental rotational constants, four chemically relevant molecular parameters,
namely the Si−I and Si−H bond lengths and the ∠(I−Si−I) and ∠(I−Si−H) bond
angles, were fit. These parameters are presented in Table 4.4 alongside previous
work by Altabef et al.[94] on the r0 gas phase structure of diiodosilane. Upon
comparison, it can be concluded that the results of these two studies are nearly
identical, as most parameters agree within experimental error. The only discrepancy is in the value of the ∠(I−Si−I) angle, but the disparity is reasonable and
only on the order of 0.14◦ . This can be attributed to the fact that the r0 structure
determined from this spectroscopic study only included isotopic substitution data
for the central silicon atom. Overall, the agreement between the two experimentally determined r0 structures is acceptable and serves as further evidence that
the spectral assignments presented in this work are correct. More on the Si−I
bond length will be discussed in the subsequent sections.
4.5.2
Nuclear Quadrupole Coupling Tensor of Iodine
In addition to providing structural information, the observation of all
three silicon isotopologues allowed for comparisons between χ, measured for
each species, to be made. As shown in-depth by Arsenault et al.[107], subtle
changes in χzz between the CH3 12 CHICH2 CH3 and CH3 13 CHICH2 CH3 species of
g-2-iodobutane can be attributed to small changes in the C−I bond length due to
isotopic substitution. Changes in the C−I bond length upon isotopic substitution,
observed through χzz , can also be found in a study of iodobenzene by Neill et al.,
although this difference was not explicitly mentioned in the text[108].
A comparison of χzz (see Table 4.3), the projection of χ nearly along the
60
Table 4.4: r0 structure versus GED of diiodosilane
a
Parameters
This work
GED[94]
Si−I (Å)
2.4236(19)a
2.423(1)b
Si−H (Å)
1.475(21)
1.470(11)
∠(I−Si−I) (◦ )
111.27(13)
110.8(2)
∠(I−Si−H) (◦ )
105.9(19)
107.3(13)
χ2
0.0012
–
σ
0.0016
–
For this work, numbers in parentheses give standard errors (1σ, 67% confidence level) in
units of the least significant figure.
b
Numbers in parentheses for GED results give 1σ standard errors in units of the least
significant figure, which were converted from the provided 3σ standard errors.
Si−I bond (angle between Si−I bond and quadrupolar z-axis is less than 1◦ ),
between the three silicon isotopologues of diiodosilane shows that in fact, small
changes in the electric field gradient at the iodine nucleus result from isotopic
substitution of the central silicon atom. The magnitude of χzz increases by 12
kHz as a direct result of isotopic substitution, namely the replacement of a
nucleus with a
29
28
Si
Si nucleus. This trend continues as a 14 kHz increase in the
magnitude of χzz is observed upon the isotopic substitution of a 29 Si nucleus with
a
30
Si nucleus.
Estimating the change in bond length that corresponds to these changes
in χzz is not as straightforward as in the case of 2-iodobutane. For one thing,
diiodosilane is of C2v symmetry, which requires that both Si−I bond lengths
must change in an identical fashion. Many possible interactions between any
combination of the iodine and silicon atoms leaves an accurate estimation of the
Si−I bond length change very hard to make. In g-2-iodobutane, the C−I bond
was estimated to decrease by 1.8 mÅ upon
13
C substitution, where the percent
increase in the mass of 13 C from 12 C is about 8%[105, 107]. Based on this previous
61
study and the fact that the percent increase in the mass of
of
30
Si from
29
29
Si from
28
Si and
Si is only about 3.5%, it is expected that the Si−I bond should
decrease by less than 1.8 mÅ upon the isotopic substitution of silicon[105]. With
this understood, the Si−I bond length change is most likely on the order of 1 mÅ
or less.
Regardless, it is understood that upon
29
Si or
30
Si substitution, the vibra-
tionally averaged Si−I bond length should decrease. Perhaps information about
the NQC tensors of iodine in iodosilane, specifically in either
29
SiH3 I or
30
SiH3 I,
would help illuminate the trends in Si−I bond length changes resulting from isotopic substitution, but these tensors have yet to be measured.
4.5.3
Chemical Nature of the Si−I Bond
As summarized by Noll[109], Si−X bonds are understood to include ionic,
covalent, and double bond character. Beginning with the Si−I bond in iodosilane,
evidence of ionic and double bond character was shown by Gordy and Cook
through a Townes-Dailey analysis[87–89], where the ionic character of the σ bond,
iσ , was estimated to be 32% and the π character, πc , of the bond was found to
be approximately 28%. These estimations neglect the possibility that the iodine
atom is hybridized and assume that the negative pole, δ − , of the σ bond is on the
halogen. Gordy and Cook also performed this analysis on tetraiodosilane where
iσ and πc were determined to be 29% and 26%, respectively. It was concluded by
Gordy and Cook that the π character of the Si−I bond in these species is evidence
for (p→d)π bonding and the contributions of hyperconjugation to the π character,
although likely present, were assumed to be small and ignored[88].
With the aim of linking that which was found for the Si−I bond in iodosilane and tetraiodosilane, a Townes-Dailey analysis[87–89] was performed for
diiodosilane. Here, it was also assumed that the iodine atom is not hybridized.
For diiodosilane, iσ was estimated to be 31%. The π character, πc , of the bond
was found to be approximately 12% if hyperconjugation effects were taken into
account, as in the case of dihalomethanes[88]. If this possibility was neglected, πc
62
was found to be approximately 28%, a value which is in better agreement with
the trend in πc that was established for iodosilane and tetraiodosilane.
The latter value makes the most sense because it is to be expected that
the Si−I bond in tetraiodosilane would have the least amount of π character,
as the silicon atom in this molecule should have the highest degree of sp3 hybridization. This can be extrapolated from the work by Donald et al.[110] on the
bonding parameters and structure of halosilanes and from the experimental work
by Kolotis et al.[111] on the gas-electron diffraction structure of tetraiodosilane.
The theoretical study of fluorine-, chlorine-, and bromine-containing silanes by
Donald et al. showed maximum and minimum sp3 hybridization at the central
silicon atom occurred in species of the form SiX4 and SiH3 X1 (X = FClBr). Although no in-depth theoretical study of this sort has been done on X = I, it can
be predicted that the same trend should hold. This trend is not surprising, however, as molecules composed of a central atom and four equivalent substituent
atoms can be of Td symmetry, where the four equivalent σ bonds would be sp3
hybridized[112]. Of course, tetraiodosilane is of Td symmetry, which was shown
experimentally by Kolotis et al.[111] where the ∠(I−Si−I), based on a ra structure, was determined to be 109.4(1)◦ . To conclude, the Si−I bond in SiI4 has
the highest degree of sp3 hybridization and therefore the least π character, by
approximately 2-3%, as was supported by a Townes-Dailey analysis. However,
some π character is still present in SiI4 due to (p→d)π bonding.
This increase in sp3 hybridization at the silicon atom from SiH3 I to SiI4 is
due directly to an increase in the number of relatively more electronegative substituents. As more electronegative substituents (i.e. iodine atoms) are added, the
central silicon atom rehybridizes such that the percent of p character in the Si−I
bond increases and the percent of s character in the Si−H bond increases.[110,
113]. This simultaneously causes an increase in hybridization and a decrease in
π character. Incidentally, this is also why the Si−I bond length (r0 ) in diiodosilane (2.4236(19)Å) is 15 mÅ less than the Si−I bond length in iodosilane
(2.43835(59)Å[114]). This phenomenon, although opposite to intuition, has been
63
widely observed and was well summarized long ago by Bent[113].
4.6
Conclusion
Rotational transitions belonging to diiodosilane were measured for the first
time, which allowed for nuclear quadrupole coupling constants, rotational constants, centrifugal distortion constants, and nuclear spin-rotation constants to be
fit for each silicon isotopologue. A r0 structure of diiodosilane was also fit via
isotopic substitution and was found to be in agreement with a previously determined r0 gas phase structure. An in-depth analysis of the NQCCs belonging to
iodine in each isotopologue revealed small changes in the Si−I bond length that
resulted from isotopic substitution. The NQCCs of iodine in this species were
compared to those measured for iodine in similar molecules, allowing for insights
into the chemical nature of the Si−I bond to be made.
4.7
Acknowledgments
The authors thank Professor Wallace (Pete) Pringle for many useful dis-
cussions and Angela Y. Chung and Yoon Jeong Choi for their involvement on this
project. This work was supported at Wesleyan University by NSF grant CHE1565276. The cluster at Wesleyan University is supported by the NSF under
CNS-0619508.
4.8
Supplemental Material
The final fit outputs for
28
SiH2 I2 ,
Chapter 6.
64
29
SiH2 I2 , and
30
SiH2 I2 are presented in
References
[1] M. Mitzlaff, R. Holm, H. Hartmann, Z. Naturforsch. A 23 (1968) 1819.
[2] M. Mitzlaff, R. Holm, H. Hartmann, Z. Naturforsch. A 23 (1968) 65.
[3] R. Holm, M. Mitzlaff, Z. Naturforsch. A 22 (1967) 288–289.
[4] M. Mitzlaff, R. Holm, H. Hartmann, Z. Naturforsch. A 22 (1967) 1415.
[5] R. Holm, M. Mitzlaff, H. Hartmann, Z. Naturforsch. A 22 (1967) 1287.
[6] R. W. Davis, M. Gerry, J. Mol. Spectrosc. 60 (1976) 117–129.
[7] K. D. Hensel, W. Jager, M. Gerry, I. Merke, J. Mol. Spectrosc. 158 (1993)
131–139.
[8] R. Mockler, J. Bailey, W. Gordy, Phys. Rev. 87 (1952) 172–172.
[9] R. C. Mockler, J. H. Bailey, W. Gordy, J. Chem. Phys. 21 (1953) 1710–1713.
[10] J. G. Smith, J. Mol. Spectrosc. 120 (1986) 110–117.
[11] J. H. Carpenter, J. G. Smith, J. Mol. Spectrosc. 121 (1987) 270–277.
[12] IUPAC, Quantities, Units and Symbols in Physical Chemistry, ER Cohen,
T. Cvitas, JG Frey, B. Holmström, K. Kuchitsu, R. marquardt, I. Mills, F.
Pavese, M. Quack, J. Stohner, HL Strauss, M. Takami, and AJ Thor. The
Royal Society of Chemistry Publications, Cambridge, UK, 2007.
[13] Q. Williams, W. Gordy, Phys. Rev. 79 (1950) 225–225.
[14] Q. Williams, J. T. Cox, W. Gordy, J. Chem. Phys. 20 (1952) 1524–1525.
[15] S. Kojima, K. Tsukada, S. Hagiwara, M. Mizushima, T. Ito, J. Chem. Phys.
20 (1952) 804–808.
[16] S. Kojima, K. Tsukada, J. Chem. Phys. 22 (1954) 2093–2094.
[17] G. Herrmann, J. Chem. Phys. 22 (1954) 2093–2093.
65
[18] Z. Kisiel, A. Kraśnicki, L. Pszczólkowski, S. T. Shipman, L. AlvarezValtierra, B. H. Pate, J. Mol. Spectrosc. 257 (2009) 177–186.
[19] W. J. Pietenpol, J. D. Rogers, Phys. Rev. 76 (1949) 690.
[20] R. W. Davis, M. Gerry, J. Mol. Spectrosc. 109 (1985) 269–282.
[21] Y. Niide, H. Tanaka, I. Ohkoshi, J. Mol. Spectrosc. 139 (1990) 11–29.
[22] D. Chadwick, D. Millen, Trans. Faraday Soc. 67 (1971) 1539–1550.
[23] D. Chadwick, D. Millen, Trans. Faraday Soc. 67 (1971) 1551–1568.
[24] D. Chadwick, D. Millen, Chem. Phys. Lett. 5 (1970) 407–409.
[25] S. Srivastava, A. Rajput, Phys. Lett. A 31 (1970) 411–412.
[26] R. Bettens, R. Brown, J. Mol. Spectrosc. 155 (1992) 55–76.
[27] Z. Kisiel, E. Bialkowska-Jaworska, L. Pszczolkowski, J. Mol. Spectrosc. 177
(1996) 240–250.
[28] Z. Kisiel, E. Bialkowska-Jaworska, L. Pszczólkowski, J. Mol. Spectrosc. 185
(1997) 71–78.
[29] Y. Niide, I. Ohkoshi, J. Mol. Spectrosc. 136 (1989) 17–30.
[30] Y. Niide, H. Tanaka, I. Ohkoshi, J. Mol. Spectrosc. 137 (1989) 104–113.
[31] S. Bailleux, D. Duflot, K. Taniguchi, S. Sakai, H. Ozeki, T. Okabayashi,
W. Bailey, J. Phys. Chem. A 118 (2014) 11744–11750.
[32] M. W. Long, Q. Williams, T. Weatherly, J. Chem. Phys. 33 (1960) 508–516.
[33] J. H. Carpenter, P. J. Seo, D. H. Whiffen, J. Mol. Spectrosc. 123 (1987)
187–196.
[34] P. Reinhart, Q. Williams, T. Weatherly, J. Chem. Phys. 53 (1970) 1418–
1421.
[35] A. Wolf, Q. Williams, T. Weatherly, J. Chem. Phys. 47 (1967) 5101–5109.
66
[36] J. Loubser, J. Chem. Phys. 36 (1962) 2808–2809.
[37] P. Favero, M. Mirei, Il Nuovo Cimento (1955-1965) 30 (1963) 502–506.
[38] A. de Luis, J. C. López, A. Guarnieri, J. Alonso, J. Mol. Struct. 413 (1997)
249–253.
[39] J. C. López, A. de Luis, S. Blanco, A. Lesarri, J. L. Alonso, J. Mol. Struct.
612 (2002) 287–303.
[40] I. Ohkoshi, Y. Niide, J. Mol. Spectrosc. 126 (1987) 282–289.
[41] I. Ohkoshi, Y. Niide, M. Takano, J. Mol. Spectrosc. 124 (1987) 118–129.
[42] Y. Niide, I. Ohkoshi, J. Mol. Spectrosc. 128 (1988) 88–97.
[43] S. Bailleux, H. Ozeki, S. Sakai, T. Okabayashi, P. Kania, D. Duflot, J. Mol.
Spectrosc. 270 (2011) 51–55.
[44] W. V. Smith, R. R. Unterberger, J. Chem. Phys. 17 (1949) 1348–1348.
[45] R. Unterberger, R. Trambarulo, W. V. Smith, J. Chem. Phys. 18 (1950)
565–566.
[46] S. Ghosh, R. Trambarulo, W. Gordy, J. Chem. Phys. 20 (1952) 605–607.
[47] P. N. Wolfe, J. Chem. Phys. 25 (1956) 976–981.
[48] D. McLay, G. Winnewisser, J. Mol. Spectrosc. 44 (1972) 32–36.
[49] A. Goertz, Z. Naturforsch. A 24 (1969) 688–689.
[50] M. Jen, D. R. Lide Jr, J. Chem. Phys. 36 (1962) 2525–2526.
[51] G. Cazzoli, G. Cotti, L. Dore, Chem. Phys. Lett. 203 (1993) 227–231.
[52] J. H. Carpenter, P. J. Seo, D. H. Whiffen, J. Mol. Spectrosc. 170 (1995)
215–227.
[53] J.-M. Colmont, D. Priem, P. Dréan, J. Demaison, J. E. Boggs, J. Mol.
Spectrosc. 191 (1998) 158–175.
67
[54] E. Bialkowska-Jaworska, Z. Kisiel, L. Pszczólkowski, J. Mol. Spectrosc. 238
(2006) 72–78.
[55] L. Margulès, J. Demaison, P. Pracna, J. Mol. Struct. 795 (2006) 157–162.
[56] R. W. Davis, M. Gerry, C. Marsden, J. Mol. Spectrosc. 101 (1983) 167–179.
[57] R. A. Booker, F. C. De Lucia, J. Mol. Spectrosc. 118 (1986) 548–549.
[58] C. F. Su, E. Beeson Jr, J. Chem. Phys. 66 (1977) 330–334.
[59] H. Takeo, C. Matsumura, Bull. Chem. Soc. Jpn. 50 (1977) 636–640.
[60] O. Baskakov, S. Dyubko, A. Katrich, V. Ilyushin, E. Alekseev, J. Mol.
Spectrosc. 199 (2000) 26–33.
[61] R. J. Myers, W. D. Gwinn, J. Chem. Phys. 20 (1952) 1420–1427.
[62] R. W. Davis, A. Robiette, M. Gerry, J. Mol. Spectrosc. 85 (1981) 399–415.
[63] M. D. Harmony, S. Mathur, S. J. Merdian, J. Mol. Spectrosc. 75 (1979)
144–149.
[64] W. Flygare, W. D. Gwinn, J. Chem. Phys. 36 (1962) 787–794.
[65] Z. Kisiel, J. Kosarzewski, L. Pszczólkowski, Acta. Phys. Pol. A 92 (1997)
507–516.
[66] Z. Kisiel, L. Pszczól, W. Caminati, P. Favero, et al., J. Chem. Phys. 105
(1996) 1778–1785.
[67] Z. Kisiel, L. Pszczólkowski, L. B. Favero, W. Caminati, J. Mol. Spectrosc.
189 (1998) 283–290.
[68] Z. Kisiel, E. Bialkowska-Jaworska, L. Pszczólkowski, J. Mol. Spectrosc. 199
(2000) 5–12.
[69] G. S. Grubbs, S. A. Cooke, Manuscript in preparation (2017).
[70] J. Park, B. Carli, A. Barbis, Geophys. Res. Lett. 16 (1989) 787–790.
68
[71] A. Bertolini, G. Carelli, A. Moretti, G. Moruzzi, IEEE J. Quant. Electon.
36 (2000) 1232–1236.
[72] F. Tullini, G. Nivellini, M. Carlotti, B. Carli, J. Mol. Spectrosc. 138 (1989)
355–374.
[73] H. Altan, B. Yu, S. Alfano, R. Alfano, Chem. Phys. Lett. 427 (2006) 241–
245.
[74] J. Pietila, V.-M. Horneman, R. Anttila, B. Lemoine, F. Raynaud, J.-M.
Colmont, Mol. Phys. 98 (2000) 549–557.
[75] C. Haase, J. Liu, F. Merkt, J. Mol. Spectrosc. 262 (2010) 61–63.
[76] P. Pracna, A. Ceausu-Velcescu, V.-M. Horneman, J. Quant. Spectrosc. Radia. Transfer 113 (2012) 1220–1225.
[77] H. Rexroad, D. Howgate, R. Gunton, J. Ollom, J. Chem. Phys. 24 (1956)
625–625.
[78] I. Merke, A. Lüchow, W. Stahl, J. Mol. Struct. 780 (2006) 295–299.
[79] A. Sharbaugh, G. Heath, L. Thomas, J. Sheridan, Nature 171 (1953) 87.
[80] R. Kewley, P. McKinney, A. Robiette, J. Mol. Spectrosc. 34 (1970) 390–398.
[81] M. Fujitake, J. Nakagawa, M. Hayashi, J. Mol. Spectrosc. 119 (1986) 367–
376.
[82] J. Nakagawa, M. Hayashi, J. Mol. Spectrosc. 93 (1982) 441–444.
[83] L. C. Sams Jr, A. W. Jache, J. Chem. Phys. 47 (1967) 1314–1315.
[84] A. Cox, T. Gayton, C. Rego, J. Mol. Struct. 190 (1988) 419–434.
[85] L. Kang, M. A. Gharaibeh, D. J. Clouthier, S. E. Novick, J. Mol. Spectrosc.
271 (2012) 33–37.
[86] M. Montejo, S. L. Hinchley, A. B. Altabef, H. E. Robertson, F. P. Ureña,
69
D. W. Rankin, J. J. L. González, Phys. Chem. Chem. Phys. 8 (2006) 477–
485.
[87] C. Townes, B. Dailey, J. Chem. Phys. 17 (1949) 782–796.
[88] W. Gordy, R. L. Cook, Microwave Molecular Spectra, Wiley, New York,
1984.
[89] S. E. Novick, J. Mol. Spectrosc. 267 (2011) 13–18.
[90] T. Balle, W. Flygare, Rev. Sci. Instrum. 52 (1981) 33–45.
[91] F. Lovas, R. Suenram, J. Chem. Phys. 87 (1987) 2010–2020.
[92] A. H. Walker, W. Chen, S. E. Novick, B. D. Bean, M. D. Marshall, J. Chem.
Phys. 102 (1995) 7298–7305.
[93] G. Grubbs II, D. A. Obenchain, H. M. Pickett, S. E. Novick, J. Chem. Phys.
141 (2014) 114306.
[94] A. B. Altabef, H. Oberhammer, J. Mol. Struct. 641 (2002) 259–261.
[95] M. Frisch, G. Trucks, H. Schlegel, G. Scuseria, M. Robb, J. Cheeseman,
G. Scalmani, V. Barone, B. Mennucci, G. Petersson, et al., Inc., Wallingford
CT (2009).
[96] D. Feller, J. Comput. Chem. 17 (1996) 1571–1586.
[97] K. L. Schuchardt, B. T. Didier, T. Elsethagen, L. Sun, V. Gurumoorthi,
J. Chase, J. Li, T. L. Windus, J. Chem. Inf. Model. 47 (2007) 1045–1052.
[98] E. A. Arsenault, D. A. Obenchain, T. A. Blake, S. Cooke, S. E. Novick, J.
Mol. Spectrosc. (2017). doi:10.1016/j.jms.2017.03.014.
[99] S. Carocci, A. Di Lieto, A. De Fanis, P. Minguzzi, S. Alanko, J. Pietilä, J.
Mol. Spectrosc. 191 (1998) 368–373.
[100] Z. Kisiel, PROSPE–Programs for ROtational SPEctroscopy, 2000.
[101] Z. Kisiel, L. Pszczólkowski, I. R. Medvedev, M. Winnewisser, F. C.
70
De Lucia, E. Herbst, J. Mol. Spectrosc. 233 (2005) 231–243. URL:
http://www.sciencedirect.com/science/article/pii/S0022285205001554.
doi:10.1016/j.jms.2005.07.006.
[102] H.
M.
Pickett,
J.
Mol.
Spectrosc.
148
(1991)
371–377.
URL:
http://www.sciencedirect.com/science/article/pii/002228529190393O.
doi:10.1016/0022-2852(91)90393-O.
[103] S. E. Novick, J. Mol. Spectrosc. 329 (2016) 1–7.
[104] S. G. Kukolich, S. C. Wofsy, J. Chem. Phys. 52 (1970) 5477–5481.
[105] J. R. Ehleringer, P. W. Rundel, Stable isotopes: History, Units, and Instrumentation, Springer, 1989.
[106] Z. Kisiel, J. Mol. Spectrosc. 218 (2003) 58–67.
[107] E. A. Arsenault, D. A. Obenchain, Y. J. Choi, T. A. Blake, S. A. Cooke,
S. E. Novick, J. Phys. Chem. A 120 (2016) 7145–7151.
[108] J. L. Neill, S. T. Shipman, L. Alvarez-Valtierra, A. Lesarri, Z. Kisiel, B. H.
Pate, J. Mol. Spectrosc. 269 (2011) 21–29.
[109] W. Noll, Chemistry and Technology of Silicones, Elsevier, 2012.
[110] K. J. Donald, M. C. Böhm, H. J. Lindner, J. Mol. Struct. Theochem 713
(2005) 215–226.
[111] M. Kolonits, M. Hargittai, Struct. Chem. 9 (1998) 349–352.
[112] B. S. Tsukerblat, Group Theory in Chemistry and Spectroscopy: A Simple
Guide to Advanced Usage, Courier Corporation, 2006.
[113] H. A. Bent, Chem. Rev. 61 (1961) 275–311.
[114] J. Duncan, J. Harvie, D. McKean, S. Cradock, J. Mol. Struct. 145 (1986)
225–242.
71
Chapter 5
Assorted Projects
This chapter includes other published projects with which I have been
scientifically involved, yet that do not fall into the theme of species containing quadrupolar nuclei. The publications, in the order they will be presented,
are:
W. Wu, E. A. Arsenault, B. E. Long, S. A. Cooke, J. Mol. Struct. 1107 (2016)
344.
W. Wu, B. E. Long, S. A. Cooke, J. Mol. Struct. 1093 (2015) 77–81.
B. E. Long, E. A. Arsenault, D. A. Obenchain, Y. J. Choi, E. J. Ocola, J.
Laane, W. C. Pringle, S. A. Cooke, J. Phys. Chem. A 120 (2016) 8686–8690.
72
Journal of Molecular Structure 1107 (2016) 344
Contents lists available at ScienceDirect
Journal of Molecular Structure
journal homepage: http://www.elsevier.com/locate/molstruc
Corrigendum
Corrigendum to “Conformations of the semifluorinated n-alkane H(CF2)8-H investigated using Fourier transform microwave
spectroscopy and quantum chemical calculations” [J. Mol. Struct. 1093
(2015) 77e81]
Weixin Wu a, Eric A. Arsenault b, B.E. Long b, S.A. Cooke a, *
a
b
School of Natural and Social Sciences, Purchase College SUNY, 735 Anderson Hill Rd, Purchase, NY 10577, USA
Department of Chemistry, Wesleyan University, Hall-Atwater Laboratories, 52 Lawn Ave, Middletown, CT 06459-0180, USA
The authors regret to inform that the investigator, Eric A. Arsenault, was inadvertently left off the author list of this published work.
S. A. Cooke would like to apologize for the inconvenience caused.
DOI of original article: http://dx.doi.org/10.1016/j.molstruc.2015.03.031.
* Corresponding author.
E-mail address: [email protected] (S.A. Cooke).
http://dx.doi.org/10.1016/j.molstruc.2015.11.028
0022-2860/© 2015 Elsevier B.V. All rights reserved.
73
Journal of Molecular Structure 1093 (2015) 77–81
Contents lists available at ScienceDirect
Journal of Molecular Structure
journal homepage: www.elsevier.com/locate/molstruc
Conformations of the semifluorinated n-alkane H–(CF2)8–H investigated
using Fourier transform microwave spectroscopy and quantum chemical
calculations
Weixin Wu a, B.E. Long b, S.A. Cooke a,⇑
a
b
School of Natural and Social Sciences, Purchase College SUNY, 735 Anderson Hill Rd, Purchase, NY 10577, USA
Department of Chemistry, Wesleyan University, Hall-Atwater Laboratories, 52 Lawn Ave, Middletown, CT 06459-0180, USA
h i g h l i g h t s
g r a p h i c a l a b s t r a c t
209 pure rotational transitions of
1H,8H-perfluorooctane recorded
between 7.8 GHz and 16.2 GHz.
Precise rotational constants have
been obtained for the first time.
Quantum chemical calculations
support the experimental
measurements.
a r t i c l e
i n f o
Article history:
Received 3 February 2015
Received in revised form 13 March 2015
Accepted 21 March 2015
Available online 27 March 2015
Keywords:
C8H2F16
Oligomer
Structure determination
Microwave spectra
a b s t r a c t
The three lowest energy conformations of the title compound have been investigated using quantum
chemical calculations and the lowest energy conformer has been observed using pure rotational
spectroscopy. The lowest energy conformer possesses C 2 symmetry, a helical CF2 backbone, with the
hydrogens nearly eclipsing one another when looking down the long axis of the molecule. The technique
of Fourier transform microwave spectroscopy in conjunction with quantum chemical calculations is
demonstrated as a complimentary method to X-ray diffraction for structural determinations of small
oligomers for which the location of hydrogen atoms may be important.
Ó 2015 Elsevier B.V. All rights reserved.
Introduction
Structural studies on the perfluoroalkane chain (–CF2)n are of
interest due to the simplicity of the repeating unit and the importance of the structural motif to modern materials chemistry. In
contrast to the paraffins it is well known that the most stable
⇑ Corresponding author.
E-mail address: [email protected] (S.A. Cooke).
URL: http://www.openscholar/purchase/cooke (S.A. Cooke).
http://dx.doi.org/10.1016/j.molstruc.2015.03.031
0022-2860/Ó 2015 Elsevier B.V. All rights reserved.
74
conformation for a perfluoroalkane chain is that of a slowly twisting helix in which each C–C–C–C dihedral is approximately 17°
from the planar trans position [1]. The source of this helicity has
been investigated in several studies [2–4], most recently by Jang
et al. [5] who conclude that the dominant source of helicity is electrostatics. Experimental methods pertaining to structure
determination of helical polymers were reviewed in 2010 by
Yashima [6]. Yashima notes that the ‘‘exact helical structures of
most of the already prepared synthetic helical polymers remain
unsolved’’. Current methods for structure determination of helical
78
W. Wu et al. / Journal of Molecular Structure 1093 (2015) 77–81
Fig. 1. The PBE0/6-311++G(d,p) geometry of three low energy conformers of 1H,8H-perfluorooctane. From left to right the panels show the conformers in the ac plane, the bc
plane, the bc plane with fluorines removed to highlight the positions of the hydrogen atoms, and the bc plane in which two lines replace the terminal C–H bonds and the
dihedral angle between these bonds is shown.
Table 1
Experimentala and PBE0/6-311++G(d,p) calculated spectroscopic parameters, and
related quantities, for H–(CF2)8–H.
Parameter
Experimental
A/MHz
B/MHz
C/MHz
695.72346(20)b
117.704735(72)
113.846756(73)
DJ /Hz
DK /Hz
dJ /Hz
0.299(58)
6.2(19)
0.074(12)
P aa /u Å2
P bb /u Å2
P cc /u Å2
4003.1628
435.9539
290.4540
la /D
lb /D
lc /D
Relative energies/cm1c
Nd
209
RMSe
0.994
Calculated
Cis
Transoid
Skew
697.035
117.324
113.504
662.957
120.663
117.916
697.899
115.718
114.166
4017.517
435.023
290.019
3855.986
429.953
332.357
4034.954
391.758
332.385
0.0
0.0
2.9
0
1.3
0.6
0.8
201
0.0
1.6
1.0
318
a
Using a Watson A reduction. The first group of constants are the three rotational
constants, followed by the centrifugal distortion constants, then the three second
moments, and then the components of the dipole moment along the principal axes.
b
Numbers in parentheses give standard errors (1r, 67% confidence level) in units
of the least significant figure.
c
Energies calculated using an MP2/6-311++G(d,p) method using the geometry
optimized at the PBE0/6-311++G(d,p) level.
d
Number of observed transitions used in the fit.
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P h
i
2
e
Root mean square deviation of the fit,
ððobs calcÞ=errorÞ =Nlines .
polymers include liquid crystalline X-ray diffraction, atomic force
microscopy, theoretical/experimental circular dichroism, and notably X-ray analysis of uniform oligomers. In this latter technique the
complex problem of polymer structure determination is simplified
through the preparation and study of small oligomers using, for
example, size exclusion chromatography. Uniform oligomers as
small as pentamers have been used to determine the nature of
the helix for some polymers [7].
Microwave spectroscopy has been demonstrated to have the
resolution and sensitivity to determine the structure of oligomers
possessing similar numbers of repeating units to those examined
with the above mentioned X-ray diffraction method [8–10]. Two
of the few shortcomings of X-ray diffraction concern the solid state
nature of the experiment and the difficulty of locating hydrogen
atoms. Fourier transform microwave spectroscopy coupled with
quantum chemical calculations and/or simple models may serve
as a complimentary technique in this regard as the sample molecules are studied once stabilized within a rapidly expanding puff
of argon gas. This ensures that the target molecule is observed in
the absence of solvent or lattice effects and furthermore the lower
energy conformers are often, but not always, the dominant species
observed.
In this study we apply Fourier transform broadband microwave
spectroscopy to the study of H–(CF2)8–H, also known as 1H,8Hperfluorooctane,
1,1,2,2,3,3,4,4,5,5,6,6,7,7,8,8-hexadecafluorooctane, or a, x-dihydroperfluorooctane. This compound contains
the repeating CF2 unit but has the added challenge that complete
structural determination requires the location of two hydrogens
within a molecule with an atomic mass of 402 Da.
Experimental
The target molecule, H–(CF2)8–H, was purchased from Synquest
Labs (97% purity) and was used without further purification. The
compound is a liquid at room temperature with a boiling point
of 432–436 K and a vapor pressure of 0.01 bar at 298 K [11]. The
liquid sample was placed in a 1/4 in. tube about 40 cm behind a
solenoid valve. Argon held at backing pressures of 1.5 bar was bubbled through the liquid prior to passage through the solenoid valve
and into a vacuum chamber held at approximately 108 bar. This
process resulted in a rotationally cold, 2 K, pulse of the target
molecule stabilized within a matrix of supersonically expanded
argon. Expansions in neon or helium were not attempted.
75
W. Wu et al. / Journal of Molecular Structure 1093 (2015) 77–81
79
Fig. 2. Top: an approximately 20 MHz portion of the experimental spectrum. The spectrum was recorded in 2 GHz sections between 7 GHz and 13 GHz. Transitions are
labeled as J0ka kc
J00ka kc . Bottom: simulated spectrum using the constants determined in Table 1.
Table 2
Helical angles derived from the calculated structures of H–(CF2)8–H. See Fig. 3 for the
dihedral angle labeling scheme.
Dihedral
Cis
Transoid
Skew
/1
/2
/3
/4
/5
/6
/7
62.8
161.4
162.1
162.4
162.1
161.4
62.8
63.1
161.5
162.2
162.1
161.9
163.1
169.6
62.6
161.4
162.5
162.1
161.6
170.8
49.5
A Fourier transform microwave spectrometer utilizing chirps of
radiation was then used to record the spectra of the target molecules between the frequency regions of 7.8 GHz and 16.2 GHz.
This instrument has been described elsewhere [12,13] and is based
upon the chirped experiment previously introduced by Pate and
coworkers [14]. Briefly, the instrument mixes a microwave pulse
of frequency m with a fast (4 ls) linear frequency sweep of DC1 GHz generating a m ± 1 GHz broadband pulse. The pulse was then
amplified (5 W) using a solid state amplifier and broadcast onto
the supersonically expanding gas sample through a horn antenna.
Following a delay of 100 ns a second antenna horn received the
resulting free induction decay (FID). The signal was then passed
through an amplification stage and proceeded to be directly digitized at a rate of 25 picoseconds per point for 800,000 points.
Transition line widths were 80 kHz with line centers possessing
an uncertainty of 8 kHz.
Geometry calculations were performed using quantum chemical
calculations at the PBE0/6-311G++(d,p) level of theory [15–19]. This
hybrid functional has been shown to perform well with perfluorinated systems [20]. The energies of three lowest energy optimized structures, shown in Fig. 1, were recalculated using the
MP2/6-311G++(d,p) level of theory [21]. The Gaussian 03 software
package was used for all quantum chemical calculations [22].
Calculated properties for the three lowest energy conformers are
given in Table 1.
Results
The observed rotational spectrum of H–(CF2)8–H consisted of
c-type transitions in a pattern expected of a nearly prolate, asymmetric rotor. Initially Q-branch progressions were located, an
example of which, near 8.7 GHz, is shown in Fig. 2. These
Q-branches are separated by 1159 MHz from one another, which
was taken to be the value of 2A ðB þ CÞ. Secondly, doublets such
as those shown near 8.68 GHz in Fig. 2 were found to repeat
approximately every 230 MHz, which was taken to be the approximate value of B þ C. From these assumptions a spectrum was
predicted using an A; B; C of 694.5 MHz, 116 MHz and 114 MHz,
respectively. This predicted spectrum produced patterns close
enough to the observed spectrum that transition quantum
Fig. 3. Figure indicating the 7 dihedral angles for 1H,8H-perfluorooctane.
76
80
W. Wu et al. / Journal of Molecular Structure 1093 (2015) 77–81
Fig. 4. A simple molecular model for 1H,8H-perfluorooctane.
numbers could be assigned. Three types of transition sequence
were observed these being c Q 1;2 (25 transitions), c Q 1;0 (29 transitions), and c R1;0 (155 transitions). No unassigned lines remained in
the spectrum.
A standard, iterative least squares analysis of all observed transitions led to the spectroscopic constants given in Table 1 and the
residuals Dm ¼ mobs mcalc of the fit reported in the supplementary
data. Pickett’s SPCAT/SPFIT suite of programs [23,24] were used
for the least squares analysis. To simplify the analytical procedure
we have written a program called SpecFitter [25] that serves as a
spectral viewer and graphical user interface to the SPCAT/SPFIT
programs. The Hamiltonian employed was the familiar semi-rigid
asymmetric top in the Watson A reduction [26]. Not all of the quartic centrifugal distortion (CD) constants were required to produce a
satisfactory fit, with the most successful combination being DJ ; DK ,
and dJ . The criterion for success being the least number of CD constants required to produce the lowest root mean square deviation
of the fit while maintaining relatively small uncertainties on the
resulting constants.
Discussion
In regards to the structure of the observed conformer comparison of experimental and calculated properties is most usefully
made through the second moments:
X
1
ðI b þ I c I a Þ ¼
mi a2i
2
i
X
1
2
mi bi
P b ¼ ðI c þ I a I b Þ ¼
2
i
X
1
mi c2i
P c ¼ ðI a þ I b I c Þ ¼
2
i
Pa ¼
ð1Þ
ð2Þ
the higher, C 2 symmetry of this conformer in contrast to the other
two conformers. The conformers differ mainly in the magnitude of
their terminal H–C–C–C dihedral angles.
The location of the hydrogen atoms as being in the designated
cis conformation can also be demonstrated through an appeal to
a very simple moment of inertia model. In the first application of
this model we have used the familiar formula for the moment of
inertia of a solid cylinder:
mR2
2
2
ml
mR2
¼
þ
12
4
Ia ¼
ð4Þ
Ib;c
ð5Þ
in which the radius, R, and length, l, of the cylinder, were treated as
adjustable parameters with the mass, m fixed at 438 Da, i.e. the
mass of the fully fluorinated octane, C8F18. The radius and length
of the cylinder were adjusted to best reproduce the experimentally
observed moments of inertia for H–(CF2)8–H. The radius and
length determined were 1.92 Å and 11.385 Å, respectively. These
values produced second moments for the model cylinder of
Pa = 4731.1 u Å2 and Pb = Pc = 403.7 u Å2. At this point two separate,
negative 18 Da point masses were allowed to ‘‘roam’’ the cylinder
with b-coordinates fixed at zero but adjustable a and c-coordinates.
The best agreement between this models second moments and the
experimental second moments occured when the negative 18 Da
point masses, i.e. the hydrogens, were symmetrically located
±4.5 Å from the cylinder origin, both 1.5 Å from the c axis
corresponding to the cis conformer. This model is illustrated in
Fig. 4 where the angle between the negative 18 point masses
h = 0° produces best agreement between the models second
moments and those experimentally observed for H–(CF2)8–H.
ð3Þ
which are given in Table 1. It is clear that the second moments of
the observed species match very closely the calculated structure
of the cis conformer, differing by, at most 0.4%. Furthermore, the
observed spectrum consisted of only c-type transitions consistent
with the calculated dipole moment of the cis conformer also given
in Table 1. The dipole moments of the transoid and skew conformers are sizable and the lack of observation of transitions from these
conformers is likely the result of the experimental conditions only
populating the lowest energy conformer.
The calculated dihedral angles for all three low energy conformers of H–(CF2)8–H are presented in Table 2. All three conformers
show the typical ‘‘trans minus’’ angle common to perfluoroalkane
chains [5]. The dihedral angles for the cis conformer nicely display
Conclusions
Fourier transform microwave spectroscopy in concert with both
quantum mechanical calculations and a simple molecular model
has been demonstrated to provide useful insights into the
structure of an a, x-dihydrogenated perfluoroalkyl oligomer. The
lowest three conformers of H–(CF2)8–H have been identified an
the lowest conformer experimentally observed through its pure
rotational spectrum. The three conformers differ primarily in the
position of the terminal hydrogens. The rotational spectroscopic
constants, together with the quantum chemical calculations
demonstrate that the lowest energy conformer possesses C 2 symmetry with the terminal hydrogen atoms eclipsing one another
when looking down the long axis of the molecule.
77
W. Wu et al. / Journal of Molecular Structure 1093 (2015) 77–81
Acknowledgements
We gratefully acknowledge financial support from the Petroleum
Research Foundation administered by the American Chemical
Society, award number 53451-UR6.
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.molstruc.2015.03.
031.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
C.W. Bunn, E.R. Howells, Nature 174 (1954) 549–551.
D.A. Dixon, J. Phys. Chem. 96 (1992) 3698–3701.
G.D. Smith, R.L. Jaffe, D.Y. Yoon, Macromolecules 27 (1994) 3166–3173.
U. Rothlisberger, K. Laasonen, M.L. Klein, M.J. Sprik, Chem. Phys. 104 (1996)
3692–3700.
S.S. Jang, M. Blanco, W.A. Goddard III, G. Caldwell, R.B. Ross, Macromolecules
36 (2003) 5331–5341.
E. Yashima, Polym. J. 42 (2010) 3–16.
Y. Ito, T. Ohara, R. Shima, M. Suginome, J. Am. Chem. Soc. 118 (1996) 9188–
9189.
J.A. Fournier, R.K. Bohn, J.A. Montgomery Jr., M. Onda, J. Phys. Chem. A 114
(2010) 1118–1122.
J.A. Fournier, C.L. Phan, R.K. Bohn, Arkivoc (2011) 5–11.
W.C. Bailey, R.K. Bohn, C.T. Dewberry, G.S. Grubbs II, S.A. Cooke, J. Mol.
Spectrosc. 270 (2011) 61–65.
A.M.A. Dias, C.M.B. Gonçalves, J.L. Legido, J.A.P. Coutinho, I.M. Marrucho, Fluid
Phase Equilib. 238 (2005) 7–12.
78
81
[12] G.S. Grubbs II, C.T. Dewberry, K.C. Etchison, K.E. Kerr, S.A. Cooke, Rev. Scient.
Instrum. 78 (2007) 096106.
[13] G.S. Grubbs II, R.A. Powoski, D. Jojola, S.A. Cooke, J. Phys. Chem. A 114 (2010)
8009–8015.
[14] G.G. Brown, B.C. Dian, K.O. Douglass, S.M. Geyer, B.H. Pate, Rev. Scient. Instrum.
79 (2008) 053103.
[15] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865–3868.
[16] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 78 (1997) 1396.
[17] C. Adamo, V. Barone, J. Chem. Phys. 110 (1999) 6158–6169.
[18] R. Krishnan, J.S. Binkley, R. Seeger, J.A. Pople, J. Chem. Phys. 72 (1980) 650.
[19] M.N. Glukhovtsev, A. Pross, M.P. McGrath, L. Radom, J. Chem. Phys. 103 (1995)
1878.
[20] M. D’Amore, G. Talarico, V. Barone, J. Am. Chem. Soc. 128 (2006) 1099–1108.
[21] C. Møller, M. Plesset, Phys. Rev. 46 (1934) 618–622.
[22] Gaussian 03 Revision D.01, M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria,
M.A. Robb, J.R. Cheeseman, J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C.
Burant, J.M. Millam, S.S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G.
Scalmani, N. Rega, G.A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R.
Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M.
Klene, X. Li, J.E. Knox, H.P. Hratchian, J.B. Cross, V. Bakken, C. Adamo, J.
Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C.
Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A. Voth, P. Salvador, J.J.
Dannenberg, V.G. Zakrzewski, S. Dapprich, A.D. Daniels, M.C. Strain, O. Farkas,
D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G.
Baboul, S. Clifford, J. Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I.
Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A.
Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong,
C. Gonzalez, J.A. Pople, Gaussian 03, Gaussian, Inc., 340 Quinnipiac Street,
Building 40, Wallingford, CT, 06492, 2003, copyright Ó 1994–2003.
[23] H.M. Pickett, J. Mol. Spectrosc. 148 (1991) 371–377.
[24] SPFIT/SPCAT package. <http://spec.jpl.nasa.gov>.
[25] S.A. Cooke, SpecFitter. <http://specfitter.droppages.com/>, 2014.
[26] J.K.G. Watson, in: J.R. Durig (Ed.), Vibrational Spectra and Structure, vol. 6,
1977, p. 1.
Article
pubs.acs.org/JPCA
Microwave Spectra, Structure, and Ring-Puckering Vibration of
Octafluorocyclopentene
B.E. Long,§,† Eric A. Arsenault,† Daniel A. Obenchain,⊥,† Yoon Jeong Choi,‡ Esther J. Ocola,¶
Jaan Laane,¶ Wallace C. Pringle,† and S.A. Cooke*,‡
†
Department of Chemistry, Wesleyan University, Hall-Atwater Laboratories, 52 Lawn Avenue, Middletown, Connecticut 06459,
United States
‡
School of Natural and Social Sciences, Purchase College SUNY, 735 Anderson Hill Road, Purchase, New York 10577, United States
¶
Department of Chemistry, Texas A & M University, College Station, Texas 77840, United States
S Supporting Information
*
ABSTRACT: The rotational spectra of octafluorocyclopentene
(C5F8) has been measured for the first time using pulsed jet Fourier
transform microwave spectroscopy in a frequency range of 6 to 16
GHz. As in the molecule cyclopentene, the carbon ring is
nonplanar, and inversion through the plane results in an inversion
pair of ground state vibrational energy levels with an inversion
splitting of 18.4 MHz. This large amplitude motion leads to the
vibration−rotation coupling of energy levels. The symmetric double
minimum ring-puckering potential function was calculated,
resulting in a barrier of 222 cm−1. The rotational constants A0 =
962.9590(1) MHz, B0 = 885.1643(4) MHz, C0 = 616.9523(4) MHz, A1 = 962.9590(1) MHz, B1 = 885.1643(4) MHz, C1 =
616.9528(4) MHz, and two centrifugal distortion constants for each state were determined for the parent species and all 13C
isotopologues. A mixed coordinate molecular structure was determined from a least-squares fit of the ground state rotational
constants of the parent and each 13C isotopologue combined with the equilibrium bond lengths and angles from quantum
chemical calculations.
■
■
INTRODUCTION
The ring-puckering of small ring molecules has been studied for
more than 50 years by both microwave1,2 and far-infrared
spectroscopy.3−5 In many cases, four-membered ring molecules
such as cyclobutane possess two equivalent energy minima
separated by low-lying barriers. Similarly, five-membered rings
containing a double bond (“pseudo-four-membered rings”)
such as cyclopentene (C5H8) have the same type of double
minimum potential energy function for the ring puckering.
Such potential functions give rise to an inversion splitting often
less than 1 cm−1. The microwave spectrum of C5H8 was first
reported by Rathjens, Jr.6 in 1962 and was further analyzed by
Butcher and Costain,7 who determined the ring-puckering
inversion frequency to be 27 GHz. In 1967, Laane and Lord8
reported the far-infrared spectrum of cyclopentene and
determined its ring-puckering potential energy function,
which had a barrier to inversion of 232 cm−1. In the present
paper, we present the microwave spectrum of the fully
fluorinated cyclopentene molecule and investigate its structure
and ring-puckering potential energy function. These will be
compared to cyclopentene results. The infrared and Raman
spectra of this molecule have been previously studied by Harris
and Longshore.9 Microwave spectra have been previously
reported for fluorocyclobutane10 and 1-chloro-cyclopentene,11
and inversion frequencies were determined for both.
© 2016 American Chemical Society
EXPERIMENTAL SECTION
The rotational spectra of octafluorocyclopentene was initially
collected from 6 to 16 GHz on a chirped Fourier transform
microwave spectrometer (FTMW). This spectrometer, which
has been described previously,12 is based on the instrument
designed by Pate et al.13 The desired microwave center
frequency is first mixed with a linear frequency sweep, lasting 6
μs, from DC to 1 GHz. The resultant microwave radiation, ν ±
1 GHz, is then broadcast directly into the vacuum chamber via
a microwave horn antenna. The incident radiation then induces
a polarization in the concurrent supersonic expansion of the gas
sample. After a ≈1 μs delay, a second horn antenna collects the
free induction decay (FID), which is then fast Fourier
transformed and directly digitized on a Tektronix TDS6124C
digital oscilloscope. The FID is collected for 20 μs, resulting in
800 000 points at 25 ps per point. The average line width of a
transition measured on this instrument is approximately 80 kHz
with an uncertainty in the center frequency of ±8 kHz.
Additional rotational transitions for the parent and all three
unique 13C isotopologues were measured with a Balle−Flygare
type spectrometer.14
Received: July 27, 2016
Revised: September 12, 2016
Published: October 4, 2016
8686
79
DOI: 10.1021/acs.jpca.6b07554
J. Phys. Chem. A 2016, 120, 8686−8690
Article
The Journal of Physical Chemistry A
Figure 1. Ab initio structure obtained from MP2/cc-pVTZ calculation.
level. Considering previous work,15−20 a scaling factor of 0.985
was used because all of the approximated frequencies were
below 1800 cm−1. The energy levels for the ring-puckering
potential energy function, shown in Figure 2, were calculated
using the Meinander−Laane Da1OPTN1 program.21
The gas sample was purchased from SynQuest Laboratories
(97% C5F8) and used without further purification. Utilizing a
flow control system, a 1.7% tank of sample in approximately 20
atm of dry argon (99.999%, Airgas) was mixed. Experiments
ran with a final sample concentration of 0.75% in 2 atm argon
carrier gas.
■
■
RESULTS
The spectral assignment for octafluorcyclopentene, an asymmetric top with κ = 0.55, was completed using Pickett’s SPFIT/
CALCULATIONS
Quantum chemical calculations were utilized to obtain an ab
initio structure for C5F8. Shown in Figure 1 is the resultant
Table 1. Fit Comparison of Parent Species
ab
initioa
states fit separately
v=0
A (MHz)
B (MHz)
C (MHz)
DJ (kHz)
dJ (kHz)
Fbc (MHz)
ΔE01 (MHz)
Nd
RMSe (kHz)
965.0
889.0
623.1
−c
−
−
−
states fit
together
v=1
962.9590(1)
962.9590(1)b
885.1643(4)
885.1643(4)
616.9523(4)
616.9528(4)
0.0229(3)
0.0223(3)
0.0052(1)
0.0050(1)
12.508(5)
18.404(5)
176
4.1
962.9590(1)
885.1666(9)
616.9503(9)
0.0226(4)
0.0051(2)
12.483(9)
18.401(6)
176
5.7
a
From MP2 level calculation. Results given are re values. bNumbers in
parentheses give standard errors (1σ, 67% confidence level) in units of
the least significant figure. cAb initio value not predicted. dNumber of
transitions used in the fit. eRoot mean square deviation of the fit,
(∑ [(obs − calc)2 ]/Nlines) fNumbers in square brackets indicate
values held constant to those obtained for the parent.
Figure 2. Theoretically determined ring-puckering potential function
of C5F8. On the x-axis, Z is the puckering coordinate of the vibrational
motion. Symmetric minima occur when the carbon ring assumes a
puckered angle of ±22.9°.
SPCAT22 programs in tandem with the AABS package.23 Table
1 presents a comparison of fits when the vibrational states are
either fit individually or together. The comparison with the ab
initio constants is very good, and deviations, to a large part, may
be attributed to the calculation of equilibrium constants
compared to the experimental determination of constants
within the stated vibrational quanta. There is a small increase in
RMS error when the states are fit together. This is perhaps an
artifact of a small difference between the two states or because
an increase in parameters causes a decrease in the statistical fit
structure calculated at the MP2/cc-pVTZ level of theory using
the GAUSSIAN09 Revision A.02 suite.15 Computation predicts
that the carbon ring backbone assumes a puckered geometry
with a puckering angle of ±22.9°. The barrier to ring inversion,
which occurs when the puckering angle is 0°, was calculated to
be 222 cm−1 at the same level of theory. Predicted vibrational
frequencies for the molecule were computed using density
functional theory (DFT) calculations at the B3LYP/cc-pVTZ
8687
80
DOI: 10.1021/acs.jpca.6b07554
J. Phys. Chem. A 2016, 120, 8686−8690
Article
The Journal of Physical Chemistry A
Table 2. Spectroscopic Parameters for Octafluorocyclopentene
predicteda
parent
A0 (MHz)
B0 (MHz)
C0 (MHz)
DJ,0 (kHz)
dJ,0 (kHz)
965.0
889.0
623.1
−
−
962.9590(1)b
885.1643(4)
616.9523(4)
0.0229(3)
0.0052(1)
960.2972(1)
884.465(1)
615.529(1)
0.026(1)
0.0079(5)
962.9430(1)
882.8039(2)
615.8092(1)
0.022(1)
[0.0052]e
960.8169(1)
885.2221(7)
616.0695(8)
0.023(1)
0.0051(5)
A1 (MHz)
B1 (MHz)
C1 (MHz)
DJ,1 (kHz)
dJ,1 (kHz)
−
−
−
−
−
962.9590(1)
885.1643(4)
616.9528(4)
0.0223(3)
0.0050(1)
960.2977(1)
884.465(1)
615.530(1)
0.025(1)
0.0069(5)
962.9445(1)
885.8041(1)
615.8098(1)
0.022(1)
[0.0050]
960.8151(1)
885.2231(9)
616.0703(8)
0.027(1)
0.0070(5)
Fbc (MHz)
ΔE01 (MHz)
Nc
RMSd (kHz)
−
−
−
−
12.508(5)
18.404(5)
176
4.1
12.58(1)
18.53(1)
38
1.0
[12.508]
[18.404]
28
1.3
12.084(9)
[18.404]
31
1.5
parameters
a
13
C1,2
13
C3,5
13
C4
From MP2 level calculation. Results given are re values. bNumbers in parentheses give standard errors (1σ, 67% confidence level) in units of the
least significant figure. cNumber of transitions used in the fit. dRoot mean square deviation of the fit,
square brackets indicate values held constant to those obtained for the parent.
(∑ [(obs − calc)2 ]/Nlines) eNumbers in
Figure 3. Ring inversion accompanied by a c-dipole moment inversion.
To fit the b- and c-type transitions, a rotation-vibration term,
FaPApZ, was required, where Fa is the Coriolis parameter, PA is
the angular momentum about the a-axis, and pZ is the linear
momentum resulting from the inversion motion. As described
by Pickett,24 a better fit of the vibration−rotation interaction
results from using the equivalent operator Fbc, (PBPC+PCPB)pZ.
Three rotational constants and two centrifugal distortion
constants were fit spectroscopically for the v = 0 and v = 1
vibrational states.
As seen in Figure 3, the ring-puckering motion resulted in
the inversion of the c-dipole moment. As a direct result of this,
c-type cross state transitions were observed. The J′ ← J″
transitions are shown in Figure 4. An additional parameter,
ΔE01, was necessary to account for the inversion frequency of
the ring-puckering motion. The experimental result is ΔE01 =
18.404(5) MHz.
Figure 4. Roughly a 50 MHz portion of spectra obtained from 10 000
averages on the chirped FTMW. Intrastate b-type inversion pairs are
shown in blue. Cross state c-type transitions are depicted in red. Labels
″ a k″c.
for transitions are Jk′′ a k′c- Jk″
■
DISCUSSION
Ring-Puckering Potential Energy Function. As is wellknown,1−5 the ring-puckering potential energy functions for
molecules such as cyclopentene and octafluorocyclopentene are
determined by the competing forces of angle strain and
torsional interactions. The former strives to maintain the planar
rings, while the torsions prefer the rings to pucker. The barrier
to planarity of cyclopentene is 232 cm−1, as determined from
far-infrared spectra by Laane and Lord.8 The barrier for the
error. The three 13C isotopologues were fit separately because
the RMS error is better when the vibrational states are treated
in this way. Spectroscopic constants for the parent species and
three 13C isotopologues are presented in Table 2. The ab initio
values agree to within 1% of the experimental rotational
constants. Line listings for all of the fits performed are given in
the supplementary data tables.
8688
81
DOI: 10.1021/acs.jpca.6b07554
J. Phys. Chem. A 2016, 120, 8686−8690
Article
The Journal of Physical Chemistry A
Table 3. Experimental r0 Structure for Octafluorocyclopentene
a
Carbons C1 and C5 are symmetrically equivalent, as well as C2 and C4. bNumbers in parentheses give standard Costain errors27 (1σ) in units of
the least significant figure.
the fluorine substitution also decreases the angle strain in the
C5F8, leading to a similar barrier to planarity. A plausible
rational for this observation in C5F8 may hinge on the electron
withdrawing properties of the fluorines which, in turn, result in
a reduced repulsion in the eclipsed torsion potential between
the two C−C bonds in the ring when compared to the
nonfluorinated C−C bonds in C5H8.26
Although the barriers are very similar for the two molecules,
because of the much higher reduced mass for the C5H8, the
ground state splitting for C5F8 is much smaller (18.404 MHz vs
27 GHz), and the 1 → 2 far-infrared transition is predicted to
be 77 vs 127.1 cm−1 for cyclopentene.8
Structural Determination. The carbon skeleton was
determined via isotopic substitution. All 13C isotopologues
were observed in natural abundance. A Schwendeman
analysis28 was preformed to gather structural details for the
molecule. Results from this analysis can be found in Table 3.
Due to the Cs symmetry of the molecule, the carbons labeled 1
and 5 are equivalent, as are those labeled 2 and 4. The dihedral
angle presented, 21.6°, refers to the puckering angle of the
carbon ring in its minimum energy conformation. This value
was calculated to be 22.9° from the MP2/cc-pVTZ structural
optimization. These dihedral angles are all presented in Table 4.
A Kraitchman analysis27,29 was also performed to confirm that
the assigned carbon positions are correct. However, the
Kraitchman errors are much larger than those from the
Schwendeman fit. As a result, these values are not used for the
structural determination of the molecule. The carbon ring is
very close to planar, so the b- and c-coordinate values for
carbons labeled 1, 2, 4, and 5 are very small. It then becomes
rather difficult to determine the structure and at best these
coordinates are found to be very small and nonzero.
Kraitchman coordinates are presented in Table 5.
Table 4. Dihedral Angle Determination
dihedrala (deg)
experimental, r0
MP2/cc-pVTZ
C5H8 experimental,6 r0
a
21.6
22.9
22.3
Labeled as φ on the structure above.
Table 5. Kraitchman Coordinates (Å) for
Octafluorocyclopentene
coordinates
atoms
|a|
|b|
|c|
C1, C5
C2, C4
C3
0.6578(24)a
1.2334(12)
0.128(12)i
1.2105(13)
0.037(4)
1.0917(14)
0.073(21)
0.087(17)
0.144(11)i
a
Values in parentheses give absolute Costain errors27 of the least
significant figure. Imaginary values are indicative that the coordinate is
very close to zero.
■
analogous fluoro compound determined in the present work is
222 cm−1. This is somewhat surprising because the CF2−CF2
torsional interactions are expected to be greater than the
CH2−CH2 interactions. The barrier to internal rotation of
ethane is 1040 cm−1, while that of hexafluoroethane (C2F6) is
1564 cm−1.25 Thus, the C5F8 barrier would be higher than that
for C5H8 based on the torsional interactions alone. Apparently,
CONCLUSION
Twelve unique spectroscopic constants were determined from
the rotational spectra of C5F8 between 6 and 16 GHz. Isotopic
substitution allowed for an experimentally derived carbon ring
structure. The ring-puckering potential function was also fit,
and both were explored in detail and compared to cyclopentene.
8689
82
DOI: 10.1021/acs.jpca.6b07554
J. Phys. Chem. A 2016, 120, 8686−8690
Article
The Journal of Physical Chemistry A
■
Spectroscopy; Laane, J., Ed.; Elsevier: Amsterdam, The Netherlands,
2009; pp 25−32.
(11) Caminati, W.; Danieli, R.; Fantoni, A. C.; Lopez, J. C. RingPuckering Motion in 1-Chloropentene: Rotational Spectrum and ab
initio Calculations. J. Mol. Spectrosc. 1997, 181, 91−98.
(12) Grubbs, G. S., II; Dewberry, C. T.; Etchison, K. C.; Kerr, K. E.;
Cooke, S. A. A Search Accelerated Correct Intensity Fourier
Transform Microwave Spectrometer with Pulsed Laser Ablation
Source. Rev. Sci. Instrum. 2007, 78, 096106.
(13) Brown, G. G.; Dian, B. C.; Douglass, K. O.; Geyer, S. M.;
Shipman, S. T.; Pate, B. H. A Broadband Fourier Transform
Microwave Spectrometer Based on Chirped Pulse Excitation. Rev.
Sci. Instrum. 2008, 79, 053103.
(14) Etchison, K. C.; Dewberry, C. T.; Cooke, S. A. BornOppenheimer Breakdown Effects and Hyperfine Structure in the
Rotational Spectrum of Strontium Monosulfide, Srs. Chem. Phys. 2007,
342, 71.
(15) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
B.; Petersson, G. A., et al. Gaussian 09, Revision A.02; Gaussian Inc.:
Wallingford, CT, 2009.
(16) Ocola, E. J.; Bauman, L. E.; Laane, J. Vibrational Spectra and
Structure of Cyclopentane and its Isotopomers. J. Phys. Chem. A 2011,
115, 6531−6542.
(17) Sheu, H.-L.; Kim, S.; Laane, J. Infrared, Raman, and Ultraviolet
Absorption Spectra and Theoretical Calculations and Structure of 2,6Difluoropyridine in Its Ground and Excited Electronic States. J. Phys.
Chem. A 2013, 117, 13596−13604.
(18) Ocola, E. J.; Medders, C.; Cooke, J. M.; Laane, J. Vibrational
Spectra, Theoretical Calculations, and Structure of 4-Silaspiro(3,3)heptane. Spectrochim. Acta, Part A 2014, 130, 397−401.
(19) Egawa, T.; Shinashi, K.; Ueda, T.; Ocola, E. J.; Chiang, W.-Y.;
Laane, J. Vapor-Phase Raman Spectra, Theoretical Calculations, and
the Vibrational and Structural Properties of cis- and trans-Stilbene. J.
Phys. Chem. A 2014, 118, 1103−1112.
(20) Ocola, E. J.; Shin, H. W.; Laane, J. Infrared and Raman Spectra
and Theoretical Calculations for Benzocyclobutane in its Electronic
Ground State. Spectrochim. Acta, Part A 2015, 136, 58−63.
(21) Meinander, N.; Laane, J. Computation of The Energy Levels of
Large-Amplitude Low-Frequency Vibrations. Comparison of the
Prediagonalized Harmonic Basis and the Prediagonalized Distributed
Gaussian Basis. J. Mol. Struct. 2001, 569, 1−24.
(22) Pickett, H. M. The Fitting and Prediction of Vibration-Rotation
Spectra with Spin Interactions. J. Mol. Spectrosc. 1991, 148, 371.
(23) Kisiel, Z.; Pszczolkowski, L.; Medvedev, I. R.; Winnewisser, M.;
De Lucia, F. C.; Herbst, C. E. Rotational Spectrum of Trans-Trans
Diethyl Ether in the Ground and Three Excited Vibrational States. J.
Mol. Spectrosc. 2005, 233, 231.
(24) Pickett, H. M. Vibration-Rotation Interactions and the Choice
of Rotating Axes for Polyatomic Molecules. J. Chem. Phys. 1972, 56,
1715−1723.
(25) Nagy, B.; Csontos, B.; Csontos, J.; Szakács, P.; Kallay, M. HighAccuracy Theoretical Thermochemistry of Fluoroethanes. J. Phys.
Chem. A 2014, 118, 4824−4836.
(26) Pringle, W. C.; Meinzer, A. L. Far Infrared Ring Puckering
Vibration of 3,3-Difluoroxetane: Effect of Fluorine Substitution on the
Ring Puckering Potential. J. Chem. Phys. 1974, 61, 2071−2076.
(27) Costain, C. C. Determination of Molecular Structure from
Ground State Rotational Constants. J. Chem. Phys. 1958, 29, 864−874.
(28) Schwendeman, R. H. Structural Parameters from Rotational
Spectra. In Critical Evaluation of Chemical and Physical Structural
Information; Lide, D. R., Jr., Paul, M. A., Ed.; National Academy of
Science, 1974; pp 94−115.
(29) Kraitchman, J. Determination of Molecular Structure from
Microwave Spectroscopic Data. Am. J. Phys. 1953, 21, 17−24.
ASSOCIATED CONTENT
S Supporting Information
*
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.jpca.6b07554.
■
Formatted output of SPFIT file for data fits (PDF)
AUTHOR INFORMATION
Corresponding Author
*E-mail: [email protected], Phone: 914-251-6675.
Present Addresses
§
B.E.L.: Department of Chemistry, Trinity University, San
Antonio, Texas 78212-7200, United States.
⊥
D.A.O.: Institut fur Physikalische Chemie and Elektrochemie,
Callinstr. 3A, 30167 Hannover, Germany.
Notes
The authors declare no competing financial interest.
■
ACKNOWLEDGMENTS
The authors thank Stew Novick for many useful discussions. An
internal grant from SUNY Purchase funded the chemicals used
in these experiments. The cluster at Wesleyan University is
supported by the NSF under Grant CNS-0619508. J.L. wishes
to thank the Robert A. Welch Foundation for financial support
under Grant A-0396. S.A.C. acknowledges financial support
from the Petroleum Research Foundation administered by the
American Chemical Society, Award 53451-UR6. Computations
were carried out on the Texas A&M University Department of
Chemistry Medusa computer system funded by the National
Science Foundation, Grant No. CHE-0541587.
■
REFERENCES
(1) Caminati, W.; Grabow, J.-U. Microwave Spectroscopy: Molecular
Systems. In Frontiers of Molecular Spectroscopy; Laane, J., Ed.; Elsevier:
Amsterdam, The Netherlands, 2009; pp 455−552, and references
therein.
(2) Caminati, W.; Grabow, J.-U. Microwave Spectroscopy: Molecular
Systems. In Frontiers of Molecular Spectroscopy; Laane, J., Ed.; Elsevier:
Amsterdam, The Netherlands, 2009; pp 383−454, and references
therein.
(3) Laane, J. Vibrational Potential Energy Surfaces in Electronic Excited
States. In Frontiers of Molecular Spectroscopy; Laane, J., Ed.; Elsevier:
Amsterdam: The Netherlands, 2009; pp 63−132, and references
therein.
(4) Laane, J. Experimental Determination of Vibrational Potential
Energy Surfaces and Molecular Structures in Electronic Excited States.
J. Phys. Chem. A 2000, 104, 7715−7733. and references therein.
(5) Laane, J. Spectroscopic Determination of Ground and Excited
State Vibrational Potential Energy Surfaces. Int. Rev. Phys. Chem. 1999,
18, 301−341. and references therein.
(6) Rathjens, G. W., Jr. Microwave Investigation of Cyclopentene. J.
Chem. Phys. 1962, 36, 2401−2406.
(7) Butcher, S. S.; Costain, C. C. Vibration-Rotation Interaction in
the Microwave Spectrum of Cyclopentene. J. Mol. Spectrosc. 1965, 15,
40−50.
(8) Laane, J.; Lord, R. C. Far-Infrared Spectra of Ring Compounds.
II. The Spectrum and Ring-Puckering Potential Function of
Cyclopentene. J. Chem. Phys. 1967, 47, 4941−4945.
(9) Harris, W. C.; Longshore, C. T. Interpretation of the Infrared and
Laser-Raman Spectra of Cyclopentene and Perfluorocyclopentene. J.
Mol. Struct. 1973, 16, 187−204.
(10) Caminati, W.; Favero, L. B.; Maris, A.; Favero, P. G. Microwave
Spectrum of the Axial Conformer and Potential Energy Function of the
Ring Puckering Motion in Fluorocyclobutane. In Frontiers of Molecular
8690
83
DOI: 10.1021/acs.jpca.6b07554
J. Phys. Chem. A 2016, 120, 8686−8690
Chapter 6
Appendix A
84
28_DIS
Thu Apr 13 13:05:00 2017
-------------------------------------------------------------------------------------=========
obs
o-c error blends Notes
o-c
wt
/ instead of : below denotes (o-c)>3*err
-------------------------------------------------------------------------------------=========
1: 5 1 5 6 7 4 0 4 5 6
12704.5896 -0.0003 0.003
2: 5 1 5 6 6 4 0 4 5 5
12705.2715 -0.0003 0.003
3: 5 1 5 6 5 4 0 4 5 4
12707.0747 0.0000 0.003
4: 5 1 5 5 6 4 0 4 4 5
12707.3164 0.0002 0.003
5: 5 1 5 6 4 4 0 4 5 3
12713.9846 0.0000 0.003
6: 5 1 5 7 8 4 0 4 6 7
12719.4873 0.0003 0.003
7: 5 1 5 4 5 4 0 4 3 4
12720.8934 -0.0003 0.003
8: 5 1 5 7 7 4 0 4 6 6
12723.9957 0.0000 0.003
9: 5 1 5 4 3 4 0 4 3 3
12724.8016 0.0000 0.003
10: 5 1 5 5 5 4 0 4 4 4
12726.1536 -0.0004 0.003
11: 5 1 5 5 3 4 0 4 4 2
12731.6211 0.0002 0.003
12: 5 1 5 5 4 4 0 4 4 3
12737.9413 -0.0006 0.003
13: 5 1 5 3 3 4 0 4 2 2
12738.4275 -0.0002 0.003
14: 5 1 5 3 4 4 0 4 6 3
12740.6469 0.0000 0.003
15: 5 1 5 7 5 4 0 4 6 4
12741.7150 0.0000 0.003
16: 5 1 5 4 4 4 0 4 3 4
12744.4589 0.0000 0.003
17: 5 1 5 8 8 4 0 4 7 7
12747.0235 0.0000 0.003
18: 5 1 5 4 4 4 0 4 3 3
12748.1912 -0.0002 0.003
19: 5 1 5 7 9 4 0 4 6 8
12749.1477 0.0002 0.003
20: 5 1 5 6 8 4 0 4 5 7
12754.4892 0.0000 0.003
21: 5 1 5 8 7 4 0 4 7 6
12754.7795 -0.0001 0.003
22: 5 1 5 4 3 4 0 4 3 2
12756.2451 0.0000 0.003
23: 5 1 5 5 2 4 0 4 4 1
12758.7377 -0.0004 0.003
24: 5 1 5 5 7 4 0 4 4 6
12759.0230 -0.0002 0.003
25: 5 1 5 8 9 4 0 4 7 8
12767.8070 0.0004 0.003
26: 5 1 5 7 4 4 0 4 2 3
12769.6420 0.0003 0.003
27: 5 1 5 4 2 4 0 4 3 1
12772.6902 0.0001 0.003
28: 5 1 5 8 6 4 0 4 7 5
12777.8249 0.0000 0.003
29: 5 1 5 4 6 4 0 4 3 5
12779.9351 0.0000 0.003
30: 5 1 5 8 10 4 0 4 7 9
12793.6087 0.0004 0.003
31: 5 1 5 3 5 4 0 4 2 4
12808.0834 0.0003 0.003
32: 5 1 5 8 5 4 0 4 7 4
12808.2025 0.0002 0.003
33: 6 1 6 6 6 5 0 5 5 5
13551.5292 -0.0006 0.003
34: 6 1 6 7 7 5 0 5 6 6
13557.1766 -0.0004 0.003
35: 6 1 6 7 6 5 0 5 6 5
13564.8330 -0.0002 0.003
36: 6 1 6 8 7 5 0 5 7 6
13568.0551 -0.0001 0.003
37: 6 1 6 7 8 5 0 5 6 7
13568.8305 -0.0002 0.003
38: 6 1 6 6 5 5 0 5 6 4
13575.8812 -0.0006 0.003
39: 6 1 6 7 5 5 0 5 5 4
13579.2094 -0.0006 0.003
40: 6 1 6 5 5 5 0 5 4 4
13592.3185 -0.0004 0.003
41: 6 1 6 6 7 5 0 5 5 6
13593.3875 0.0000 0.003
42: 6 1 6 8 8 5 0 5 7 7
13597.9169 -0.0004 0.003
43: 6 1 6 6 4 5 0 5 5 3
13599.3301 -0.0007 0.003
44: 6 1 6 5 6 5 0 5 4 5
13605.9188 -0.0005 0.003
45: 6 1 6 8 6 5 0 5 7 5
13606.9210 -0.0009 0.003
46: 6 1 6 5 4 5 0 5 4 3
13609.1579 -0.0005 0.003
47: 6 1 6 8 9 5 0 5 7 8
13615.4132 -0.0004 0.003
48: 6 1 6 4 4 5 0 5 6 3
13619.3685 0.0004 0.003
49: 6 1 6 6 8 5 0 5 5 7
13621.2529 -0.0003 0.003
85
50: 6 1 6 6 3 5 0 5 5 2
51: 6 1 6 9 8 5 0 5 8 7
52: 6 1 6 9 9 5 0 5 8 8
53: 6 1 6 8 5 5 0 5 7 4
54: 6 1 6 5 7 5 2 3 8 6
55: 6 1 6 9 7 5 0 5 8 6
56: 6 1 6 9 10 5 0 5 8 9
57: 6 1 6 9 11 5 0 5 8 10
58: 6 1 6 9 6 5 0 5 8 5
59: 7 1 7 8 8 6 0 6 7 8
60: 7 1 7 8 8 6 0 6 7 7
61: 7 1 7 7 7 6 0 6 6 6
62: 7 1 7 8 7 6 0 6 7 6
63: 7 1 7 8 9 6 0 6 7 8
64: 7 1 7 7 6 6 0 6 6 5
65: 7 1 7 7 8 6 0 6 8 7
66: 7 1 7 6 6 6 0 6 5 5
67: 7 1 7 9 8 6 0 6 6 7
68: 7 1 7 9 9 6 0 6 8 8
69: 7 1 7 7 5 6 0 6 6 4
70: 7 1 7 9 10 6 0 6 8 9
71: 7 1 7 6 7 6 0 6 5 6
72: 7 1 7 6 5 6 0 6 5 4
73: 7 1 7 8 10 6 0 6 7 9
74: 7 1 7 9 7 6 0 6 8 6
75: 7 1 7 8 5 6 0 6 7 4
76: 7 1 7 7 9 6 0 6 6 8
77: 7 1 7 5 5 6 0 6 4 4
78: 7 1 7 10 9 6 0 6 9 8
79: 7 1 7 9 11 6 0 6 8 10
80: 7 1 7 10 10 6 0 6 9 9
81: 7 1 7 6 4 6 0 6 6 3
82: 7 1 7 7 4 6 0 6 5 3
83: 7 1 7 5 4 6 0 6 4 3
84: 7 1 7 5 6 6 0 6 4 5
85: 7 1 7 9 6 6 0 6 8 5
86: 7 1 7 10 11 6 0 6 9 10
87: 7 1 7 10 8 6 0 6 9 7
88: 7 1 7 10 12 6 0 6 9 11
89: 7 1 7 5 7 6 0 6 9 6
90: 8 1 8 9 9 7 0 7 8 8
91: 8 1 8 8 8 7 0 7 7 7
92: 8 1 8 9 8 7 0 7 8 7
93: 8 1 8 8 7 7 0 7 7 6
94: 8 1 8 9 10 7 0 7 8 9
95: 8 1 8 7 7 7 0 7 6 6
96: 8 1 8 8 9 7 0 7 9 8
97: 8 1 8 9 7 7 0 7 8 6
98: 8 1 8 10 9 7 0 7 7 8
99: 8 1 8 8 6 7 0 7 7 5
100: 8 1 8 10 10 7 0 7 9 9
101: 8 1 8 10 11 7 0 7 9 10
102: 8 1 8 7 6 7 0 7 6 5
103: 8 1 8 9 11 7 0 7 8 10
104: 8 1 8 7 8 7 0 7 6 7
105: 8 1 8 11 12 7 0 7 10 11
13635.1161 0.0002 0.003
13635.4831 -0.0005 0.003
13643.6325 0.0000 0.003
13648.2388 0.0000 0.003
13661.5410 -0.0001 0.003
13663.0146 -0.0003 0.003
13666.9546 -0.0003 0.003
13692.9489 0.0000 0.003
13704.4371 -0.0001 0.003
14454.2586 -0.0002 0.003
14456.8196 0.0005 0.003
14459.3828 0.0000 0.003
14462.6604 0.0004 0.003
14471.8247 -0.0012 0.003
14471.8769 0.0024 0.003
14481.4661 0.0000 0.003
14486.8386 0.0000 0.003
14490.9100 0.0001 0.003
14493.2014 0.0000 0.003
14495.1007 -0.0001 0.003
14501.2843 0.0002 0.003
14509.6251 0.0000 0.003
14510.5048 -0.0002 0.003
14511.0545 0.0000 0.003
14511.7693 0.0000 0.003
14516.1558 -0.0002 0.003
14522.8469 -0.0003 0.003
14524.2769 0.0000 0.003
14525.9489 -0.0002 0.003
14527.1325 -0.0002 0.003
14527.9855 -0.0005 0.003
14530.1194 -0.0007 0.003
14540.8674 -0.0001 0.003
14541.9434 -0.0005 0.003
14545.5793 -0.0002 0.003
14548.1477 -0.0001 0.003
14554.2957 -0.0003 0.003
14555.2383 -0.0003 0.003
14578.5793 -0.0008 0.003
14594.6305 0.0027 0.003
15332.6888 0.0003 0.003
15334.2860 0.0005 0.003
15337.1315 0.0009 0.003
15344.9894 0.0006 0.003
15349.6643 0.0004 0.003
15360.3589 0.0000 0.003
15362.3482 0.0006 0.003
15362.5560 0.0001 0.003
15368.4731 0.0003 0.003
15370.9553 0.0001 0.003
15372.8132 0.0008 0.003
15379.4299 0.0006 0.003
15385.2579 0.0000 0.003
15385.9926 -0.0002 0.003
15390.7478 0.0005 0.003
15432.8127 0.0001 0.003
86
106:
107:
108:
109:
110:
111:
112:
113:
114:
115:
116:
117:
118:
119:
120:
121:
122:
123:
124:
125:
126:
127:
128:
129:
130:
131:
132:
133:
134:
135:
136:
137:
138:
139:
140:
141:
142:
143:
144:
145:
146:
147:
148:
149:
150:
151:
152:
153:
154:
155:
156:
157:
158:
159:
160:
161:
4
4
4
4
4
8
8
8
8
8
8
8
8
8
8
8
8
8
8
8
8
8
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
2
2
2
2
2
2
2
2
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
3
1
1
1
1
1
1
1
1
1
1
1
1
4 4 5
4 5 6
4 4 4
4 6 7
4 5 6
8 10 8
8 9 6
8 8 10
8 11 10
8 10 12
8 11 11
8 6 6
8 7 5
8 6 5
8 6 7
8 10 7
8 7 9
8 11 9
8 6 4
8 7 4
8 6 8
8 6 3
4 3 4
4 5 5
4 5 4
4 6 5
4 5 3
4 6 6
4 6 4
4 4 3
4 4 2
4 7 7
4 5 7
4 7 6
4 4 6
4 6 8
4 3 3
4 7 8
4 3 5
4 7 9
4 4 5
4 3 4
4 4 4
2 2 4
4 2 4
2 3 1
2 3 0
2 4 2
2 3 4
2 4 1
2 3 3
2 3 2
2 4 6
1 4 2
1 2 2
1 4 2
3 0 3 3 4
3 0 3 4 5
3 0 3 3 3
3 0 3 5 6
3 0 3 5 6
7 0 7 9 7
7 0 7 8 5
7 0 7 7 9
7 0 7 10 9
7 0 7 9 11
7 0 7 10 10
7 0 7 5 5
7 0 7 6 4
7 0 7 7 4
7 0 7 5 6
7 0 7 9 6
7 2 5 10 8
7 0 7 10 8
7 0 7 5 3
7 0 7 6 3
7 0 7 10 7
7 0 7 5 2
3 0 3 2 3
3 0 3 5 4
3 0 3 4 3
3 0 3 4 4
3 0 3 4 2
3 0 3 5 5
3 0 3 5 3
3 0 3 3 2
3 0 3 4 1
3 0 3 6 6
3 0 3 4 6
3 0 3 6 5
3 0 3 3 5
3 0 3 5 7
3 0 3 2 2
3 0 3 6 7
3 0 3 6 4
3 0 3 6 8
3 0 3 4 5
3 0 3 3 4
3 0 3 5 4
3 2 1 1 3
3 0 3 1 3
1 0 1 4 1
1 0 1 2 1
1 0 1 4 2
1 0 1 2 4
1 0 1 2 1
1 0 1 2 3
1 0 1 2 2
1 0 1 4 6
0 0 0 3 3
0 0 0 3 2
0 0 0 3 1
11775.1200 0.0001 0.003
11781.3631 0.0002 0.003
11810.9765 0.0001 0.003
11811.0709 0.0002 0.003
11811.3036 0.0000 0.003
15392.2770 0.0001 0.003
15392.5908 -0.0006 0.003
15396.8794 0.0001 0.003
15398.8240 -0.0006 0.003
15399.3088 0.0001 0.003
15401.3983 -0.0001 0.003
15401.8726 -0.0001 0.003
15410.5467 0.0003 0.003
15418.9800 -0.0001 0.003
15427.3614 0.0004 0.003
15428.6473 0.0003 0.003
15431.0549 -0.0005 0.003
15431.2205 -0.0012 0.003
15458.3585 -0.0003 0.003
15458.8431 0.0003 0.003
15472.9416 -0.0012 0.003
15504.9780 -0.0001 0.003
11771.2023 0.0001 0.003
11796.2285 0.0000 0.003
11797.2635 0.0000 0.003
11803.8890 0.0001 0.003
11806.4168 0.0003 0.003
11812.6965 0.0001 0.003
11817.3375 0.0001 0.003
11822.8055 0.0005 0.003
11825.0506 0.0003 0.003
11835.7762 0.0005 0.003
11837.9801 0.0000 0.003
11839.4114 0.0000 0.003
11840.9462 0.0002 0.003
11844.7830 0.0003 0.003
11850.2713 0.0006 0.003
11853.6346 0.0005 0.003
11861.0686 0.0004 0.003
11878.0633 0.0006 0.003
11879.4192 0.0004 0.003
11880.4051 0.0003 0.003
11880.7757 0.0010 0.003
11888.4430 0.0007 0.003
11889.7283 0.0006 0.003
9770.9903 0.0008 0.003
9768.3847 0.0008 0.003
9808.8755 0.0000 0.003
9810.3398 -0.0002 0.003
9830.2520 0.0004 0.003
9831.2957 -0.0003 0.003
9833.8900 -0.0001 0.003
9840.5910 -0.0003 0.003
9034.0270 -0.0006 0.003
9038.6840 -0.0002 0.003
9040.4394 -0.0003 0.003
87
162: 1 1 1 4 3 0 0 0 3 4
9048.8638 0.0000 0.003
163: 1 1 1 4 3 0 0 0 3 2
9052.4708 0.0000 0.003
164: 1 1 1 2 3 0 0 0 3 3
9053.2430 -0.0005 0.003
165: 1 1 1 3 3 0 0 0 3 4
9062.3068 -0.0004 0.003
166: 1 1 1 3 4 0 0 0 3 4
9065.6225 -0.0003 0.003
167: 1 1 1 3 3 0 0 0 3 2
9065.9144 0.0001 0.003
168: 1 1 1 3 5 0 0 0 3 5
9070.3216 -0.0002 0.003
169: 1 1 1 3 2 0 0 0 3 3
9071.4422 0.0000 0.003
170: 1 1 1 4 6 0 0 0 3 5
9040.6087 -0.0001 0.003
171: 1 1 1 2 4 0 0 0 3 5
9031.8653 -0.0004 0.003
172: 1 1 1 4 5 0 0 0 3 4
9061.6310 -0.0002 0.003
173: 1 1 1 4 4 0 0 0 3 3
9063.9686 -0.0001 0.003
174: 2 1 2 4 5 1 0 1 4 6
9807.3710 -0.0001 0.003
175: 2 1 2 3 2 1 0 1 4 1
9859.4540 -0.0006 0.003
176: 2 1 2 5 5 1 0 1 2 4
9868.8689 0.0001 0.003
177: 2 1 2 2 1 1 0 1 2 2
9874.8862 0.0010 0.003
178: 2 1 2 4 3 1 0 1 4 3
9882.4455 -0.0003 0.003
179: 2 1 2 4 1 1 0 1 4 2
9896.0758 0.0002 0.003
180: 2 1 2 2 4 1 0 1 2 4
9898.0457 0.0001 0.003
181: 2 1 2 3 3 1 0 1 4 2
9906.4258 -0.0004 0.003
182: 2 1 2 5 4 1 0 1 2 3
9909.2522 -0.0001 0.003
183: 2 1 2 4 5 1 0 1 4 4
9917.5864 -0.0008 0.003
184: 2 1 2 2 2 1 0 1 2 3
9927.1941 0.0006 0.003
185: 2 1 2 3 5 1 0 1 4 5
9948.9017 -0.0010 0.003
186: 2 1 2 2 3 1 0 1 2 2
9951.7173 0.0010 0.003
187: 2 1 2 3 4 1 0 1 4 3
9965.0600 0.0000 0.003
188: 2 1 2 3 2 1 0 1 4 3
9966.5606 -0.0005 0.003
189: 2 1 2 5 3 1 0 1 2 3
9968.2242 0.0002 0.003
-------------------------------------------------------------------------------PARAMETERS IN FIT (values truncated):
10000
A /MHz
8574.5226(11)
1
20000
B /MHz
496.1276(29)
2
30000
C /MHz
471.4278(19)
3
110010000 3/2Chi_aa /MHz
-1040.6316(70)
4
-220010000
3 /MHz
-1040.6316(70)
= 1.00000 * 4
110040000 1/4(Chi_bb /MHz
-155.0030(18)
5
-220040000
1 /MHz
-155.0030(18)
= 1.00000 * 5
110610000
Chi_ab /MHz
-856.223(11)
6
-220610000
Chi_ab /MHz
856.223(11)
= -1.00000 * 6
110011000 3/2Chi_aa_ /MHz
0.0301(61)
7
-220011000
3 /MHz
0.0301(61)
= 1.00000 * 7
200
DJ /kHz
0.0358(62)
8
1100
DJK /kHz
-3.9(10)
9
10010000
Caa /MHz
0.00459(51)
10
-20010000
Caa /MHz
0.00459(51)
= 1.00000 * 10
10020000
Cbb /MHz
0.00056(17)
11
-20020000
Cbb /MHz
0.00056(17)
= 1.00000 * 11
10030000
Ccc /MHz
0.00090(12)
12
-20030000
Ccc /MHz
0.00090(12)
= 1.00000 * 12
MICROWAVE AVG =
-0.000001 MHz, IR AVG =
MICROWAVE RMS =
0.000510 MHz, IR RMS =
END OF ITERATION 1 OLD, NEW RMS ERROR=
88
0.00000
0.00000
0.16988
0.16988
distinct frequency lines in fit: 189
distinct parameters of fit: 12
for standard parameter errors previous errors are multiplied by: 0.175544
MICROWAVE
lines fitted
lines lines
RMS
RMS ERROR J range Ka range
freq. range
total dv=0 dv.ne.0 UNFITTD e>900
v"= 1
2
0
2
0
0
0.000652 0.21731 3 4 0 3 11888 11890
v"= 2
17
4 13
0
0
0.000483 0.16088 1 5 0 1 9768 12808
v"= 3
27
6 21
0
0
0.000258 0.08607 0 5 0 1 9032 12780
v"= 4
31
8 23
0
0
0.000418 0.13938 1 7 0 1 9771 14546
v"= 5
26
2 24
0
0
0.000339 0.11304 3 8 0 1 11796 15505
v"= 6
26
2 24
0
0
0.000578 0.19282 3 8 0 1 11836 15459
v"= 7
23
0 23
0
0
0.000426 0.14193 4 8 0 1 12747 15419
v"= 8
19
0 19
0
0
0.000325 0.10842 5 8 0 2 13635 15393
v"= 9
12
0 12
0
0
0.000907 0.30231 6 8 0 1 14526 15429
v"=10
6
0
6
0
0
0.000765 0.25495 7 8 0 2 15399 15473
-------------------------------------------------------------------------------------------total:
189 22 167
0
0
0.000479 0.15951
PARAMETERS IN FIT WITH STANDARD ERRORS ON THOSE THAT ARE FITTED:
(values rounded)
10000
A /MHz
8574.52262(20)
20000
B /MHz
496.12760(51)
30000
C /MHz
471.42783(34)
110010000 3/2Chi_aa /MHz
-1040.6316(12)
-220010000
3 /MHz
-1040.6316(12)
110040000 1/4(Chi_bb /MHz
-155.00309(33)
-220040000
1 /MHz
-155.00309(33)
110610000
Chi_ab /MHz
-856.2237(19)
-220610000
Chi_ab /MHz
856.2237(19)
110011000 3/2Chi_aa_ /MHz
0.0301(10)
-220011000
3 /MHz
0.0301(10)
=
200
DJ /kHz
0.0358(10)
1100
DJK /kHz
-3.94(17)
10010000
Caa /MHz
0.004590(89)
-20010000
Caa /MHz
0.004590(89)
10020000
Cbb /MHz
0.000568(30)
-20020000
Cbb /MHz
0.000568(30)
10030000
Ccc /MHz
0.000909(21)
-20030000
Ccc /MHz
0.000909(21)
1
2
3
4
= 1.00000 * 4
5
= 1.00000 * 5
6
= -1.00000 * 6
7
1.00000 * 7
8
9
10
= 1.00000 * 10
11
= 1.00000 * 11
12
= 1.00000 * 12
CORRELATION COEFFICIENTS, C.ij:
A
B
C
3/2Chi_a 1/4(Chi_ Chi_ab 3/2Chi_a -DJ
A
1.0000
B
-0.5237 1.0000
C
0.2478 -0.8102 1.0000
3/2Chi_aa -0.0902 -0.3493 0.1826 1.0000
1/4(Chi_bb 0.2135 -0.3206 0.1644 0.3188 1.0000
Chi_ab
0.6316 -0.5194 0.4070 0.2203 0.2753 1.0000
3/2Chi_aa_ -0.1063 -0.0526 0.1544 -0.1351 -0.7918 -0.1436 1.0000
-DJ
-0.3301 0.6452 -0.1193 -0.4222 -0.3018 -0.2301 0.0649 1.0000
89
-DJK
Caa
Cbb
Ccc
-0.3399
0.1055
0.0486
-0.0254
-DJK
-DJK
Caa
Cbb
Ccc
0.8562
0.0484
0.0817
-0.0236
Caa
-0.9941
-0.1200
-0.0586
0.1364
Cbb
-0.1926
0.0972
-0.0545
-0.0197
-0.1939 -0.4575 -0.1366 0.1785
0.1832 -0.0245 -0.1777 -0.0521
0.0870 -0.0050 -0.0548 0.1120
0.0431 -0.0087 0.0108 0.1563
Ccc
1.0000
0.1088 1.0000
0.0516 -0.1695 1.0000
-0.1298 -0.4883 0.8105 1.0000
Mean value of |C.ij|, i.ne.j = 0.2366
Mean value of C.ij, i.ne.j = -0.0353
No correlations with absolute value greater than 0.9950
Worst fitted lines (obs-calc/error):
89:
63:
186:
152:
43:
96:
101:
145:
159:
135:
164:
51:
30:
0.9
-0.4
0.3
0.3
-0.2
0.2
0.2
0.2
-0.2
0.2
-0.2
-0.2
0.1
64: 0.8
185: -0.3
45: -0.3
151: 0.3
149: 0.2
38: -0.2
150: 0.2
175: -0.2
142: 0.2
104: 0.2
44: -0.2
60: 0.2
62: 0.1
126: -0.4
148: 0.3
92: 0.3
88: -0.3
81: -0.2
184: 0.2
93: 0.2
114: -0.2
83: -0.2
122: -0.2
143: 0.2
46: -0.2
123: -0.4
177: 0.3
100: 0.3
183: -0.3
112: -0.2
33: -0.2
12: -0.2
39: -0.2
188: -0.2
91: 0.2
80: -0.2
137: 0.2
89: 7 1 7 5 7
64: 7 1 7 7 6
126: 8 1 8 6 8
123: 8 1 8 11 9
63: 7 1 7 8 9
185: 2 1 2 3 5
148: 4 1 4 4 4
177: 2 1 2 2 1
186: 2 1 2 2 3
45: 6 1 6 8 6
6 0 6 9 6
14594.6305 0.0027 0.003
6 0 6 6 5
14471.8769 0.0024 0.003
7 0 7 10 7
15472.9416 -0.0012 0.003
7 0 7 10 8
15431.2205 -0.0012 0.003
6 0 6 7 8
14471.8247 -0.0012 0.003
1 0 1 4 5
9948.9017 -0.0010 0.003
3 0 3 5 4
11880.7757 0.0010 0.003
1 0 1 2 2
9874.8862 0.0010 0.003
1 0 1 2 2
9951.7173 0.0010 0.003
5 0 5 7 5
13606.9210 -0.0009 0.003
_____________________________________
__________________________________________/ SPFIT output reformatted with
PIFORM
90
29_DIS
Thu Apr 13 13:05:42 2017
-------------------------------------------------------------------------------------=========
obs
o-c error blends Notes
o-c
wt
/ instead of : below denotes (o-c)>3*err
-------------------------------------------------------------------------------------=========
1: 4 1 4 7 9 3 0 3 6 8
11662.6836 0.0003 0.003
2: 4 1 4 6 7 3 0 3 5 6
11595.9342 0.0000 0.003
3: 4 1 4 7 8 3 0 3 6 7
11638.6578 0.0005 0.003
4: 4 1 4 6 8 3 0 3 5 7
11629.3248 0.0008 0.003
5: 4 1 4 4 6 3 0 3 3 5
11625.8288 -0.0003 0.003
6: 4 1 4 7 6 3 0 3 6 5
11623.9751 0.0000 0.003
7: 4 1 4 5 7 3 0 3 4 6
11623.2856 0.0004 0.003
8: 4 1 4 7 7 3 0 3 6 6
11620.3362 0.0005 0.003
9: 6 1 6 9 9 5 0 5 8 8
13423.0841 0.0001 0.003
10: 6 1 6 8 10 5 0 5 7 9
13422.9386 -0.0003 0.003
11: 6 1 6 9 8 5 0 5 8 7
13417.1411 -0.0002 0.003
12: 5 1 5 8 9 4 0 4 7 8
12550.1064 0.0004 0.003
13: 5 1 5 8 10 4 0 4 7 9
12575.5439 0.0004 0.003
14: 4 1 4 6 6 3 0 3 5 5
11598.1310 0.0002 0.003
15: 4 1 4 5 3 3 0 3 4 2
11590.7710 0.0005 0.003
16: 4 1 4 6 5 3 0 3 4 4
11589.3765 0.0000 0.003
17: 4 1 4 5 4 3 0 3 4 3
11581.9113 0.0002 0.003
18: 4 1 4 5 5 3 0 3 5 4
11581.2855 0.0001 0.003
19: 4 1 4 5 6 3 0 3 4 5
11566.2432 -0.0002 0.003
20: 4 1 4 4 5 3 0 3 3 4
11559.5108 0.0000 0.003
21: 4 1 4 2 4 3 0 3 1 3
11674.1603 0.0003 0.003
22: 4 1 4 3 4 3 0 3 3 4
11664.7741 0.0015 0.003
23: 4 1 4 3 5 3 0 3 6 4
11645.7732 0.0002 0.003
24: 5 1 5 6 7 4 0 4 5 6
12488.6550 -0.0001 0.003
25: 5 1 5 6 6 4 0 4 5 5
12489.8694 -0.0002 0.003
26: 5 1 5 5 6 4 0 4 4 5
12490.2596 0.0001 0.003
27: 5 1 5 6 5 4 0 4 5 4
12491.1294 -0.0001 0.003
28: 5 1 5 6 4 4 0 4 5 3
12497.4417 -0.0004 0.003
29: 5 1 5 7 8 4 0 4 6 7
12501.8745 0.0002 0.003
30: 5 1 5 4 5 4 0 4 3 4
12504.5934 -0.0001 0.003
31: 5 1 5 7 7 4 0 4 6 6
12507.9342 0.0001 0.003
32: 5 1 5 5 5 4 0 4 4 4
12512.5888 -0.0006 0.003
33: 5 1 5 5 3 4 0 4 4 2
12514.7924 -0.0002 0.003
34: 5 1 5 7 6 4 0 4 6 5
12516.3838 -0.0001 0.003
35: 5 1 5 5 4 4 0 4 4 3
12523.5643 0.0006 0.003
36: 5 1 5 7 5 4 0 4 6 4
12525.7282 0.0006 0.003
37: 5 1 5 8 8 4 0 4 7 7
12529.0405 0.0003 0.003
38: 5 1 5 7 9 4 0 4 6 8
12531.0982 0.0003 0.003
39: 5 1 5 8 7 4 0 4 7 6
12537.7610 0.0000 0.003
40: 5 1 5 6 8 4 0 4 5 7
12538.7924 -0.0001 0.003
41: 5 1 5 5 7 4 0 4 4 6
12542.5832 -0.0001 0.003
42: 5 1 5 7 4 4 0 4 2 3
12552.9662 0.0007 0.003
43: 5 1 5 8 6 4 0 4 7 5
12560.6268 0.0000 0.003
44: 5 1 5 4 6 4 0 4 3 5
12563.1557 -0.0004 0.003
45: 5 1 5 3 5 4 0 4 2 4
12590.4004 0.0005 0.003
46: 5 1 5 8 5 4 0 4 7 4
12590.5270 -0.0006 0.003
47: 7 1 7 10 11 6 0 6 9 10
14331.6714 -0.0009 0.003
48: 7 1 7 9 11 6 0 6 8 10
14303.5701 -0.0001 0.003
49: 7 1 7 10 12 6 0 6 9 11
14354.9464 -0.0007 0.003
91
50: 7 1 7 7 6
51: 7 1 7 8 9
52: 3 1 3 6 8
53: 3 1 3 6 7
54: 3 1 3 4 5
55: 3 1 3 5 6
56: 3 1 3 5 6
57: 3 1 3 5 4
58: 3 1 3 5 5
59: 3 1 3 4 4
60: 3 1 3 3 5
61: 3 1 3 4 3
62: 3 1 3 6 6
63: 3 1 3 6 5
64: 3 1 3 4 6
65: 3 1 3 4 5
66: 3 1 3 2 4
67: 3 1 3 5 7
68: 3 1 3 6 3
69: 3 1 3 1 3
70: 3 1 3 1 2
71: 6 1 6 7 7
72: 6 1 6 7 6
73: 6 1 6 7 8
74: 6 1 6 8 7
75: 6 1 6 6 5
76: 6 1 6 7 5
77: 6 1 6 6 7
78: 6 1 6 5 5
79: 6 1 6 8 8
80: 6 1 6 6 4
81: 6 1 6 5 6
82: 6 1 6 7 9
83: 6 1 6 8 6
84: 6 1 6 8 9
85: 6 1 6 4 4
86: 6 1 6 6 8
87: 6 1 6 8 5
88: 6 1 6 9 7
89: 6 1 6 9 10
90: 6 1 6 9 11
91: 6 1 6 9 6
92: 7 1 7 8 8
93: 7 1 7 7 7
94: 7 1 7 8 7
95: 7 1 7 7 8
96: 7 1 7 8 6
97: 7 1 7 6 6
98: 7 1 7 9 8
99: 7 1 7 9 9
100: 7 1 7 7 5
101: 7 1 7 9 10
102: 7 1 7 6 7
103: 7 1 7 8 10
104: 7 1 7 9 7
105: 7 1 7 8 5
6 0 6 6 5
6 0 6 7 8
2 0 2 5 7
2 0 2 5 6
2 0 2 3 4
2 0 2 4 6
2 0 2 4 5
2 0 2 4 3
2 0 2 4 4
2 0 2 3 3
2 0 2 2 4
2 0 2 4 2
2 0 2 5 5
2 0 2 5 4
2 0 2 3 5
2 0 2 4 5
2 0 2 5 3
2 0 2 4 6
2 0 2 1 2
2 0 2 5 2
2 0 2 1 3
5 0 5 6 6
5 0 5 6 5
5 0 5 6 7
5 0 5 7 6
5 0 5 6 4
5 0 5 5 4
5 0 5 5 6
5 0 5 4 4
5 0 5 7 7
5 0 5 5 3
5 0 5 4 5
5 0 5 6 8
5 0 5 7 5
5 0 5 7 8
5 0 5 6 3
5 0 5 5 7
5 0 5 7 4
5 0 5 8 6
5 0 5 8 9
5 0 5 8 10
5 0 5 8 5
6 0 6 7 7
6 0 6 6 6
6 0 6 7 6
6 0 6 8 7
6 0 6 7 5
6 0 6 5 5
6 0 6 6 7
6 0 6 8 8
6 0 6 6 4
6 0 6 8 9
6 0 6 5 6
6 0 6 7 9
6 0 6 8 6
6 0 6 7 4
14249.0812 0.0000 0.003
14249.2691 0.0004 0.003
10718.4819 0.0000 0.003
10693.1120 -0.0002 0.003
10608.2921 -0.0005 0.003
10634.8502 -0.0001 0.003
10659.9003 0.0001 0.003
10668.2589 -0.0001 0.003
10668.3911 -0.0004 0.003
10634.5456 -0.0006 0.003
10671.3921 0.0004 0.003
10673.0757 0.0000 0.003
10682.5186 0.0001 0.003
10684.4968 0.0003 0.003
10687.4269 0.0000 0.003
10695.4596 0.0005 0.003
10698.9806 0.0000 0.003
10716.7465 0.0001 0.003
10731.2659 0.0011 0.003
10739.3862 0.0002 0.003
10766.3669 0.0008 0.003
13341.9368 0.0004 0.003
13349.2527 -0.0001 0.003
13353.8280 -0.0001 0.003
13356.3144 -0.0006 0.003
13360.3384 -0.0008 0.003
13365.7052 -0.0004 0.003
13376.2430 0.0003 0.003
13376.8583 -0.0005 0.003
13381.1838 -0.0002 0.003
13383.5938 -0.0006 0.003
13389.0404 0.0001 0.003
13389.8491 -0.0006 0.003
13392.2110 -0.0003 0.003
13395.3809 0.0000 0.003
13405.2182 -0.0001 0.003
13406.8618 -0.0006 0.003
13431.4258 0.0004 0.003
13444.8749 -0.0002 0.003
13446.7580 0.0000 0.003
13472.2111 -0.0002 0.003
13484.8327 -0.0006 0.003
14234.3519 0.0000 0.003
14236.9420 0.0001 0.003
14239.9459 0.0004 0.003
14259.6920 0.0004 0.003
14262.9232 0.0000 0.003
14264.0292 -0.0002 0.003
14268.1403 0.0002 0.003
14271.3685 0.0004 0.003
14272.6532 0.0000 0.003
14278.8875 0.0004 0.003
14286.9954 -0.0002 0.003
14288.1449 0.0002 0.003
14289.5529 0.0001 0.003
14293.0038 0.0000 0.003
92
106: 7 1 7 7 9 6 0 6 6 8
14299.9232 -0.0006 0.003
107: 7 1 7 10 9 6 0 6 9 8
14302.7686 0.0000 0.003
108: 7 1 7 10 10 6 0 6 9 9
14304.7204 -0.0003 0.003
109: 7 1 7 5 6 6 0 6 4 5
14322.6835 0.0000 0.003
110: 7 1 7 9 6 6 0 6 8 5
14325.4943 -0.0003 0.003
111: 7 1 7 10 8 6 0 6 9 7
14332.5252 -0.0008 0.003
112: 7 1 7 6 4 6 0 6 6 3
14308.3424 -0.0004 0.003
113: 2 1 2 5 7 1 0 1 4 6
9813.5359 0.0000 0.003
114: 2 1 2 5 6 1 0 1 4 5
9817.8772 -0.0002 0.003
115: 2 1 2 4 6 1 0 1 4 6
9630.4840 -0.0005 0.003
116: 2 1 2 5 5 1 0 1 4 4
9860.4421 -0.0009 0.003
117: 2 1 2 4 6 1 0 1 3 5
9880.0099 -0.0007 0.003
118: 2 1 2 4 5 1 0 1 3 5
9846.7150 -0.0002 0.003
119: 2 1 2 1 3 1 0 1 2 4
9791.4551 0.0005 0.003
120: 2 1 2 2 3 1 0 1 2 2
9741.0436 -0.0001 0.003
121: 2 1 2 5 4 1 0 1 2 3
9698.7910 -0.0001 0.003
122: 2 1 2 4 4 1 0 1 3 3
9803.4204 -0.0008 0.003
123: 1 1 1 2 1 0 0 0 3 1
8815.3291 -0.0005 0.003
124: 1 1 1 4 1 0 0 0 3 0
8816.6627 0.0007 0.003
125: 1 1 1 2 4 0 0 0 3 5
8820.9370 -0.0010 0.003
126: 1 1 1 2 2 0 0 0 3 2
8827.9837 -0.0006 0.003
127: 1 1 1 4 2 0 0 0 3 1
8829.4056 0.0002 0.003
128: 1 1 1 4 6 0 0 0 3 5
8829.7363 0.0002 0.003
129: 1 1 1 4 3 0 0 0 3 4
8838.0505 -0.0001 0.003
130: 1 1 1 4 3 0 0 0 3 2
8841.5895 -0.0003 0.003
131: 1 1 1 2 3 0 0 0 3 3
8842.8264 0.0001 0.003
132: 1 1 1 4 5 0 0 0 3 4
8851.1109 -0.0004 0.003
133: 1 1 1 4 4 0 0 0 3 3
8853.4289 0.0002 0.003
134: 1 1 1 3 4 0 0 0 3 4
8855.2453 0.0006 0.003
135: 1 1 1 3 5 0 0 0 3 5
8859.4731 -0.0004 0.003
136: 1 1 1 3 2 0 0 0 3 3
8860.8221 0.0008 0.003
137: 1 1 1 3 0 0 0 0 3 1
8864.3228 -0.0001 0.003
138: 1 1 1 3 2 0 0 0 3 1
8867.1129 0.0000 0.003
139: 8 1 8 9 9 7 0 7 8 8
15107.2397 0.0004 0.003
140: 8 1 8 8 8 7 0 7 7 7
15108.2355 0.0000 0.003
141: 8 1 8 9 8 7 0 7 8 7
15111.4244 0.0006 0.003
142: 8 1 8 8 7 7 0 7 7 6
15119.2324 0.0004 0.003
143: 8 1 8 8 6 7 0 7 7 5
15146.2147 -0.0012 0.003
144: 8 1 8 9 11 7 0 7 8 10
15159.5241 -0.0002 0.003
145: 8 1 8 11 12 7 0 7 10 11
15208.5465 -0.0001 0.003
146: 8 1 8 10 10 7 0 7 9 9
15149.3807 0.0017 0.003
147: 8 1 8 9 10 7 0 7 8 9
15124.6138 0.0006 0.003
-------------------------------------------------------------------------------PARAMETERS IN FIT (values truncated):
10000
A /MHz
8363.9676(10)
20000
B /MHz
496.1410(40)
30000
C /MHz
470.7846(34)
200
DJ /kHz
0.0343(53)
1100
DJK /kHz
-3.8(17)
110010000 3/2*Chi_aa /MHz
-1040.664(11)
-220010000
3 /MHz
-1040.664(11)
110040000 1/4(Chi_bb /MHz
-154.9992(22)
-220040000
1 /MHz
-154.9992(22)
110610000
Chi_ab /MHz
-856.226(10)
93
1
2
3
4
5
6
= 1.00000 * 6
7
= 1.00000 * 7
8
-220610000
Chi_ab /MHz
110011000 3/2*Chi_aa /MHz
-220011000
3 /MHz
10010000
Caa /MHz
-20010000
Caa /MHz
10020000
Cbb /MHz
-20020000
Cbb /MHz
10030000
Ccc /MHz
-20030000
Ccc /MHz
856.226(10)
0.0294(67)
0.0294(67)
0.00437(51)
0.00437(51)
0.00043(25)
0.00043(25)
0.00084(17)
0.00084(17)
= -1.00000 * 8
9
= 1.00000 * 9
10
= 1.00000 * 10
11
= 1.00000 * 11
12
= 1.00000 * 12
MICROWAVE AVG =
-0.000001 MHz, IR AVG =
MICROWAVE RMS =
0.000483 MHz, IR RMS =
END OF ITERATION 1 OLD, NEW RMS ERROR=
0.00000
0.00000
0.16094
0.16094
distinct frequency lines in fit: 147
distinct parameters of fit: 12
for standard parameter errors previous errors are multiplied by: 0.167941
MICROWAVE
lines fitted
lines lines
RMS
RMS ERROR J range Ka range
freq. range
total dv=0 dv.ne.0 UNFITTD e>900
v"= 1
3
1
2
0
0
0.000804 0.26805 2 4 0 1 10731 11674
v"= 2
6
1
5
0
0
0.000442 0.14720 1 5 0 1 9699 12590
v"= 3
27
6 21
0
0
0.000543 0.18088 0 5 0 1 8815 12563
v"= 4
24
3 21
0
0
0.000358 0.11941 1 7 0 1 9630 14323
v"= 5
21
1 20
0
0
0.000320 0.10666 2 7 0 1 10683 14287
v"= 6
22
2 20
0
0
0.000363 0.12102 3 7 0 1 11620 14308
v"= 7
21
0 21
0
0
0.000406 0.13550 4 8 0 1 12529 15146
v"= 8
16
0 16
0
0
0.000354 0.11785 5 8 0 1 13417 15160
v"= 9
6
0
6
0
0
0.000906 0.30185 6 8 0 1 14303 15149
v"=10
1
0
1
0
0
0.000100 0.03333 7 8 0 1 15209 15209
-------------------------------------------------------------------------------------------total:
147 14 133
0
0
0.000449 0.14955
PARAMETERS IN FIT WITH STANDARD ERRORS ON THOSE THAT ARE FITTED:
(values rounded)
10000
A /MHz
8363.96765(18)
20000
B /MHz
496.14100(67)
30000
C /MHz
470.78460(57)
200
DJ /kHz
0.03430(89)
1100
DJK /kHz
-3.86(28)
110010000 3/2*Chi_aa /MHz
-1040.6649(18)
-220010000
3 /MHz
-1040.6649(18)
110040000 1/4(Chi_bb /MHz
-154.99926(37)
-220040000
1 /MHz
-154.99926(37)
110610000
Chi_ab /MHz
-856.2266(17)
-220610000
Chi_ab /MHz
856.2266(17)
110011000 3/2*Chi_aa /MHz
0.0294(11)
-220011000
3 /MHz
0.0294(11)
=
10010000
Caa /MHz
0.004370(85)
-20010000
Caa /MHz
0.004370(85)
10020000
Cbb /MHz
0.000434(43)
-20020000
Cbb /MHz
0.000434(43)
10030000
Ccc /MHz
0.000845(29)
94
1
2
3
4
5
6
= 1.00000 * 6
7
= 1.00000 * 7
8
= -1.00000 * 8
9
1.00000 * 9
10
= 1.00000 * 10
11
= 1.00000 * 11
12
-20030000
Ccc /MHz
0.000845(29)
= 1.00000 * 12
CORRELATION COEFFICIENTS, C.ij:
A
B
C
-DJ
-DJK
3/2*Chi_ 1/4(Chi_ Chi_ab
A
1.0000
B
-0.5099 1.0000
C
0.2721 -0.9319 1.0000
-DJ
-0.3980 0.2654 0.0625 1.0000
-DJK
-0.3266 0.9474 -0.9981 -0.0372 1.0000
3/2*Chi_aa 0.1151 -0.1082 0.0663 -0.0843 -0.0750 1.0000
1/4(Chi_bb 0.2063 -0.1218 0.0482 -0.1500 -0.0620 0.7589 1.0000
Chi_ab
0.5393 -0.3475 0.1910 -0.3356 -0.2197 0.0489 0.1070 1.0000
3/2*Chi_aa -0.0914 -0.1408 0.1783 0.0060 -0.1687 -0.5815 -0.8157 0.0007
Caa
0.1766 -0.1067 0.0503 -0.1515 -0.0573 0.0222 0.1257 0.0774
Cbb
0.0604 -0.0222 -0.0063 0.0258 -0.0061 0.0359 -0.0035 0.0388
Ccc
-0.0211 0.0509 -0.0626 0.0606 0.0537 0.1182 0.0466 -0.0034
3/2*Chi_ Caa
Cbb
Ccc
3/2*Chi_aa 1.0000
Caa
-0.0925 1.0000
Cbb
-0.0389 -0.2753 1.0000
Ccc
-0.0424 -0.5428 0.8450 1.0000
Mean value of |C.ij|, i.ne.j = 0.2051
Mean value of C.ij, i.ne.j = -0.0354
Worst correlations, with absolute value greater than 0.9950:
1100 -DJK
<->
30000 C
-0.998071
Worst fitted lines (obs-calc/error):
146:
125:
75:
4:
124:
35:
86:
36:
115:
8:
119:
142:
51:
0.6
-0.3
-0.3
0.3
0.2
0.2
-0.2
0.2
-0.2
0.2
0.2
0.1
0.1
22: 0.5
116: -0.3
111: -0.3
49: -0.2
147: 0.2
141: 0.2
106: -0.2
74: -0.2
45: 0.2
123: -0.2
3: 0.2
12: 0.1
135: -0.1
143: -0.4
47: -0.3
136: 0.3
42: 0.2
46: -0.2
126: -0.2
32: -0.2
82: -0.2
15: 0.2
54: -0.2
112: -0.1
139: 0.1
146: 8 1 8 10 10 7 0 7 9 9
22: 4 1 4 3 4 3 0 3 3 4
143: 8 1 8 8 6 7 0 7 7 5
68: 3 1 3 6 3 2 0 2 1 2
125: 1 1 1 2 4 0 0 0 3 5
68: 0.4
122: -0.3
70: 0.3
117: -0.2
80: -0.2
59: -0.2
134: 0.2
91: -0.2
78: -0.2
65: 0.2
94: 0.1
76: -0.1
15149.3807 0.0017 0.003
11664.7741 0.0015 0.003
15146.2147 -0.0012 0.003
10731.2659 0.0011 0.003
8820.9370 -0.0010 0.003
95
116: 2 1 2 5 5 1 0 1 4 4
9860.4421 -0.0009 0.003
47: 7 1 7 10 11 6 0 6 9 10
14331.6714 -0.0009 0.003
122: 2 1 2 4 4 1 0 1 3 3
9803.4204 -0.0008 0.003
75: 6 1 6 6 5 5 0 5 6 4
13360.3384 -0.0008 0.003
111: 7 1 7 10 8 6 0 6 9 7
14332.5252 -0.0008 0.003
_____________________________________
__________________________________________/ SPFIT output reformatted with
PIFORM
96
30_DIS
Thu Apr 13 13:06:27 2017
-------------------------------------------------------------------------------------=========
obs
o-c error blends Notes
o-c
wt
/ instead of : below denotes (o-c)>3*err
-------------------------------------------------------------------------------------=========
1: 5 1 5 8 10 4 0 4 7 9
12369.4447 0.0010 0.003
2: 5 1 5 4 6 4 0 4 3 5
12358.7678 -0.0010 0.003
3: 5 1 5 8 9 4 0 4 7 8
12344.3718 -0.0001 0.003
4: 5 1 5 5 7 4 0 4 4 6
12338.5530 -0.0001 0.003
5: 5 1 5 6 8 4 0 4 5 7
12335.7161 0.0002 0.003
6: 5 1 5 6 7 4 0 4 5 6
12285.3090 -0.0004 0.003
7: 5 1 5 5 6 4 0 4 4 5
12285.4822 0.0001 0.003
8: 5 1 5 6 6 4 0 4 5 5
12287.2590 0.0000 0.003
9: 5 1 5 6 4 4 0 4 5 3
12293.4061 0.0007 0.003
10: 5 1 5 7 8 4 0 4 6 7
12296.2462 0.0002 0.003
11: 5 1 5 4 5 4 0 4 3 4
12300.9302 0.0004 0.003
12: 5 1 5 7 7 4 0 4 6 6
12304.3179 -0.0003 0.003
13: 5 1 5 5 3 4 0 4 4 2
12310.3467 -0.0005 0.003
14: 5 1 5 5 5 4 0 4 4 4
12312.5163 0.0005 0.003
15: 5 1 5 7 6 4 0 4 6 5
12315.6631 -0.0007 0.003
16: 5 1 5 3 4 4 0 4 2 3
12320.7392 -0.0003 0.003
17: 5 1 5 7 5 4 0 4 6 4
12322.2885 0.0008 0.003
18: 5 1 5 5 4 4 0 4 4 3
12322.4681 -0.0002 0.003
19: 5 1 5 8 8 4 0 4 7 7
12323.0305 -0.0002 0.003
20: 5 1 5 7 9 4 0 4 6 8
12325.0168 0.0002 0.003
21: 5 1 5 4 4 4 0 4 3 3
12328.8735 0.0002 0.003
22: 5 1 5 8 7 4 0 4 7 6
12333.1126 -0.0005 0.003
23: 4 1 4 7 8 3 0 3 6 7
11435.4934 -0.0009 0.003
24: 4 1 4 7 9 3 0 3 6 8
11459.1547 0.0009 0.003
25: 8 1 8 9 11 7 0 7 8 10
14945.1307 -0.0008 0.003
26: 8 1 8 10 10 7 0 7 9 9
14938.6617 0.0007 0.003
27: 8 1 8 9 8 7 0 7 8 7
14897.9638 0.0001 0.003
-------------------------------------------------------------------------------PARAMETERS IN FIT (values truncated):
10000
A /MHz
8165.331(22)
1
20000
B /MHz
496.1493(65)
2
30000
C /MHz
470.1462(35)
3
200
DJ /kHz
[ 0.034273534]
4
1100
DJK /kHz
[-3.863214938]
5
110010000 3/2*Chi_aa /MHz
-1040.687(63)
6
-220010000
3 /MHz
-1040.687(63)
= 1.00000 * 6
110040000 1/4*(Chi_b /MHz
-154.9940(43)
7
-220040000
1 /MHz
-154.9940(43)
= 1.00000 * 7
110610000
Chi_ab /MHz
-856.237(55)
8
-220610000
Chi_ab /MHz
856.237(55)
= -1.00000 * 8
110011000 3/2*Chi_aa /MHz
[ 0.029390982]
9
-220011000
3 /MHz
[ 0.029390982]
= 1.00000 * 9
10010000
Caa /MHz
[ 0.004374655873]
10
-20010000
Caa /MHz
[ 0.004374655873]
= 1.00000 * 10
10020000
Cbb /MHz
[ 0.000434372859]
11
-20020000
Cbb /MHz
[ 0.000434372859]
= 1.00000 * 11
10030000
Ccc /MHz
[ 0.000845412496]
12
97
-20030000
Ccc /MHz
[ 0.000845412496]
MICROWAVE AVG =
0.000000 MHz, IR AVG =
MICROWAVE RMS =
0.000576 MHz, IR RMS =
END OF ITERATION 1 OLD, NEW RMS ERROR=
distinct frequency lines in fit:
distinct parameters of fit:
= 1.00000 * 12
0.00000
0.00000
0.19184
0.19184
27
6
for standard parameter errors previous errors are multiplied by: 0.217526
MICROWAVE
lines fitted
lines lines
RMS
RMS ERROR
freq. range
total dv=0 dv.ne.0 UNFITTD e>900
v"= 2
1
0
1
0
0
0.000300 0.10000 4 5 0
v"= 3
3
0
3
0
0
0.000632 0.21082 4 5 0
v"= 4
5
0
5
0
0
0.000335 0.11155 4 5 0
v"= 5
4
0
4
0
0
0.000415 0.13844 4 5 0
v"= 6
7
0
7
0
0
0.000646 0.21529 3 5 0
v"= 7
4
0
4
0
0
0.000570 0.19003 4 5 0
v"= 8
2
0
2
0
0
0.000570 0.19003 7 8 0
v"= 9
1
0
1
0
0
0.000700 0.23333 7 8 0
-------------------------------------------------------------------------------------------total:
27
0 27
0
0
0.000541 0.18031
J range Ka range
1
1
1
1
1
1
1
1
12321
12301
12285
12285
11435
12323
14898
14939
12321
12359
12339
12336
12325
12369
14945
14939
PARAMETERS IN FIT WITH STANDARD ERRORS ON THOSE THAT ARE FITTED:
(values rounded)
10000
A /MHz
8165.3310(48)
1
20000
B /MHz
496.1493(14)
2
30000
C /MHz
470.14620(76)
3
200
DJ /kHz
[ 0.034273534]
4
1100
DJK /kHz
[-3.863214938]
5
110010000 3/2*Chi_aa /MHz
-1040.687(13)
6
-220010000
3 /MHz
-1040.687(13)
= 1.00000 * 6
110040000 1/4*(Chi_b /MHz
-154.99400(93)
7
-220040000
1 /MHz
-154.99400(93)
= 1.00000 * 7
110610000
Chi_ab /MHz
-856.237(11)
8
-220610000
Chi_ab /MHz
856.237(11)
= -1.00000 * 8
110011000 3/2*Chi_aa /MHz
[ 0.029390982]
9
-220011000
3 /MHz
[ 0.029390982]
= 1.00000 * 9
10010000
Caa /MHz
[ 0.004374655873]
10
-20010000
Caa /MHz
[ 0.004374655873]
= 1.00000 * 10
10020000
Cbb /MHz
[ 0.000434372859]
11
-20020000
Cbb /MHz
[ 0.000434372859]
= 1.00000 * 11
10030000
Ccc /MHz
[ 0.000845412496]
12
-20030000
Ccc /MHz
[ 0.000845412496]
= 1.00000 * 12
CORRELATION COEFFICIENTS, C.ij:
A
B
C
3/2*Chi_ 1/4*(Chi Chi_ab
A
1.0000
B
-0.9949 1.0000
C
-0.9979 0.9991 1.0000
3/2*Chi_aa 0.0851 -0.0701 -0.0757 1.0000
98
1/4*(Chi_b 0.1551 -0.1138 -0.1299 0.6499 1.0000
Chi_ab -0.8048 0.8369 0.8300 0.2019 0.2696 1.0000
Mean value of |C.ij|, i.ne.j = 0.4810
Mean value of C.ij, i.ne.j = 0.0560
Worst correlations, with absolute value greater than 0.9950:
30000 C
30000 C
<->
<->
10000 A
20000 B
-0.997920
0.999121
Worst fitted lines (obs-calc/error):
1:
25:
15:
6:
5:
18:
7:
0.3
-0.3
-0.2
-0.1
0.1
-0.1
0.0
2: -0.3
17: 0.3
14: 0.2
11: 0.1
21: 0.1
10: 0.1
4: 0.0
23: -0.3
9: 0.2
22: -0.2
12: -0.1
20: 0.1
27: 0.0
8: 0.0
24:
26:
13:
16:
19:
3:
0.3
0.2
-0.2
-0.1
-0.1
0.0
1: 5 1 5 8 10 4 0 4 7 9
12369.4447 0.0010 0.003
2: 5 1 5 4 6 4 0 4 3 5
12358.7678 -0.0010 0.003
23: 4 1 4 7 8 3 0 3 6 7
11435.4934 -0.0009 0.003
24: 4 1 4 7 9 3 0 3 6 8
11459.1547 0.0009 0.003
25: 8 1 8 9 11 7 0 7 8 10
14945.1307 -0.0008 0.003
17: 5 1 5 7 5 4 0 4 6 4
12322.2885 0.0008 0.003
9: 5 1 5 6 4 4 0 4 5 3
12293.4061 0.0007 0.003
26: 8 1 8 10 10 7 0 7 9 9
14938.6617 0.0007 0.003
15: 5 1 5 7 6 4 0 4 6 5
12315.6631 -0.0007 0.003
14: 5 1 5 5 5 4 0 4 4 4
12312.5163 0.0005 0.003
_____________________________________
__________________________________________/ SPFIT output reformatted with
PIFORM
99
List of Figures
2.1
The experimental rotational spectrum of 2-iodobutane is shown
above the baseline and the simulated spectra of the various conformers are shown below. This portion of spectrum illustrates the
heavy overlap of the hyperfine structure due to the iodine nucleus
in each of the three conformers. The predicted gauche-, anti-, and
gauche0 -2-iodobutane transitions are represented by purple, red,
and green, respectively. Theoretical transition intensities have been
scaled using the relative ab initio energies. . . . . . . . . . . . . .
2.2
14
Illustrations of the three ab initio structures. The C1 −C2 −C3 −C4
dihedral angles for the a-, g0 -, and g-conformers are 64◦ , -62◦ , and
171◦ , respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3
The Doppler doublet of the a-type S-branch transition, 404
3
2
14
←
221 21 , of the parent conformer of g-2-iodobutane is shown, where
the average of each peak is taken to be the transition frequency,
11011.621 MHz. Labels on the y-axis of the plot were omitted,
since the intensity of the transition is arbitrary. The transition was
measured on a Balle-Flygare type spectrometer with 100 averages,
with a signal-to-noise ratio of 76. . . . . . . . . . . . . . . . . . .
100
18
3.1
Calculated structures of the anti-anti (aa)-, gauche-anti (ga)-, and
gauche-gauche (gg)-conformers of 1-iodobutane from an ab initio
optimization. The C−C−C−C and C−C−C−I dihedral angles for
the aa-, ga-, and gg-species were calculated to be 180◦ , 180◦ ; 179◦ ,
66◦ ; and -65◦ , -63◦ . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2
37
A small portion of the experimental spectrum of 1-iodobutane is
shown in black, with the final simulated spectra for the aa-, ga-, and
gg-conformers shown below in green, purple, and red, respectively.
The quantum numbers associated with each rotational transition
are also presented above the experimental spectrum, in the form
0
0
00
00
JK
0
0 F ← JK 00 K 00 F . . . . . . . . . . . . . . . . . . . . . . . . . .
a Kc
a c
3.3
38
A 100 MHz portion of three different predictions of the rotational
spectrum of aa-1-iodobutane. The top orange spectrum is based on
a hybrid tensor (discussed in Spectral Assignments) and ab initio
rotational constants, the middle blue spectrum is the prediction
using the experimental results belonging to the aa-conformer, and
the bottom orange spectrum is based completely on ab initio results. 38
101
List of Tables
1.1
Classes of rigid rotors . . . . . . . . . . . . . . . . . . . . . . . . .
3
2.1
Ab initio results of 2-iodobutane at the MP2 level of theory . . .
15
2.2
Spectroscopic parameters of three conformations of 2-iodobutane .
20
2.3
Spectroscopic parameters for g-2-iodobutane . . . . . . . . . . . .
21
2.4
STRFIT r0 structural parameters of gauche-2-iodobutane . . . . .
23
2.5
Kraitchman versus ab initio coordinates for gauche-2-iodobutane .
24
2.6
Conformational comparison of the diagonalized NQC tensor of iodine in 2-iodobutane . . . . . . . . . . . . . . . . . . . . . . . . .
2.7
Comparison of the diagonalized NQC tensor of iodine 2-iodobutane
with other iodoalkanes . . . . . . . . . . . . . . . . . . . . . . . .
2.8
25
27
Rotation of the diagonalized NQC tensor of iodine into the principal axis system of g-2-iodobutane . . . . . . . . . . . . . . . . . .
28
2.9
Changes in the ab initio NQC tensor due to isotopic substitution
29
3.1
Rotational constants and NQCCs for three conformers of 1-iodobutane
as determined from ab initio optimization at the MP2 level of theory 37
3.2
Spectroscopic parameters of 1-iodobutane . . . . . . . . . . . . . .
3.3
Comparison of the two methods of prediction for the NQC tensors
39
in 1-iodobutane . . . . . . . . . . . . . . . . . . . . . . . . . . . .
42
3.4
Comparison of this work with the previous study of 1-iodobutane
43
3.5
NQC tensor of iodine in 1-iodobutane . . . . . . . . . . . . . . . .
44
3.6
NQC tensor of bromine in 1-bromobutane[37] . . . . . . . . . . .
45
3.7
A comparison of this work with similar haloalkanes . . . . . . . .
46
102
4.1
Results obtained from equation (4.1) . . . . . . . . . . . . . . . .
56
4.2
Spectroscopic parameters of three silicon isotopologues of diiodosilane 57
4.3
NQC tensor of iodine in diiodosilane . . . . . . . . . . . . . . . .
59
4.4
r0 structure versus GED of diiodosilane . . . . . . . . . . . . . . .
61