The Use of Chirped Pulse Millimeter-Wave ... Dynamics and Kinetics ARCIHNES

advertisement
The Use of Chirped Pulse Millimeter-Wave Spectroscopy in Chemical
Dynamics and Kinetics
by
Rachel Glyn Shaver
B.S. Chemistry
University of Texas at Austin, 2009
SUBMITTED TO THE DEPARTMENT OF CHEMISTRY IN PARTIAL
FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF
MASTER OF SCIENCE IN CHEMISTRY
AT THE
MASSACHUSETTS INSTITUTE OF TECHNOLOGY
FEBRUARY 2013
ARCIHNES
i-SEACHUSETTS INSTNT'
OF TECHNOLOGY
@2013 Massachusetts Institute of Technology
All rights reserved
Signature of Author:
Department of Chemistry
January 22, 2013
Certified by:
Robert W. Field
Haslam and Dewey Professor of Chemistry
Thesis Supervisor
Accepted by:
Robert W. Field
Chairman, Department Committee on Graduate Theses
1
The Use of Chirped Pulse Millimeter-Wave Spectroscopy in Chemical Dynamics and Kinetics
by
Rachel Glyn Shaver
Submitted to the Department of Chemistry
on January 22, 2013 in Partial Fulfillment of the
Requirements for the Degree of Master of Science in
Chemistry
ABSTRACT
Chirped-pulse millimeter wave (CPmmW) spectroscopy is a revolutionary technique that has taken
advantage of advances in electronics to give high signal to noise broadband rotational spectra in a very
short period of time that provides meaningful line intensities. We have implemented this technique in the
58 - 102 GHz range to study the rotational spectra of molecules with two heavy atoms.
Photolysis (at 193 nm) and pyrolysis of vinyl cyanide have produced differing HCN and HNC vibrational
population distributions. The photolysis experiment does not sample a collisional regime and the resulting
spectra show excited states of HCN and HNC, whereas the pyrolysis experiment, which does sample a
collisional regime, results in spectra that are devoid of vibrational satellites. This indicates that the
intensities of vibrational satellite transitions sample the photolysis reaction only and not post-photolysis
collisional effects. Mono-deuterated vinyl cyanide was photolyzed at 193 nm, in which all HCN/HNC are
produced via a four-center mechanism and all DCN/DNC are produced via a three-center mechanism. The
HCN and HNC products dominate, demonstrating the greater importance of the three-center mechanism.
CPmmW spectroscopy is also a valuable tool in studying unimolecular and bimolecular reactions. We
have studied the unimolecular decomposition of deuterated methyl nitrite which produces DNO products
and bimolecular hydrogen abstraction reaction of NO with acetaldehyde resulting in HNO products.
These reactions demonstrate the potential use of nitric oxide radical as a gas-phase catalyst to perform
cracking of hydrocarbons and sugars.
Thesis Supervisor: Robert W. Field
Title: Haslam and Dewey Professor of Chemistry
2
Dedication
This thesis is for the pleasure of my spiritual master, His Holiness Radhanath Swami Maharaja, without
Whose mercy and guidance I would be lost.
3
Acknowledgements
I would like to acknowledge many people for their support, guidance and advice over the past few years, making
this thesis a reality. First of all, I'd like to thank my advisor Bob Field for his encouragement, kind support and
patience as I battled not only with research issues, but with life issues as well. I feel honored to have worked under
such a talented and intelligent, yet sincere and humble, advisor.
Dr. Kirill Prozument was my mentor, friend, and ever well-wisher throughout my time at MIT and is inseparable
from this work. He not only helped me with practical things but he also inspired me by his curiosity and brilliant
method of doing science and he helped me to appreciate and understand the subject beyond what I could have
without him.
I would also like to acknowledge Prof. John Muenter, who was a second advisor to me and who was very kind,
encouraging, and gracious with his knowledge of all things spectroscopy. And of course I'd like to thank of all the
members of the Field group, especially Josh, Tony, David, Barratt, Yan and Bryan, for being always willing to help
or just a source for a laugh; you were all a pleasure to work with.
Apart from the people associated with MIT, there were many other people in my life who were instrumental in
supporting me during my time here:
My mother was physically far away but without her support I don't think I would have survived the past few years;
she was always there for me whether it was to laugh or to cry. I love and appreciate her more than she can know.
I am also very grateful to my temple president, Maha Visnu das, for his support in me pursuing studies at MIT (and
of the erratic schedule I had as a result) while at the same time providing me the utmost privilege of performing
devotional service for Sri Sri Radha Gopivallabha while living at the ISKCON temple in Boston. I'd also like to
thank my spiritual mothers Saranagati devi dasi and Gopa-lila devi dasi, who were always there to give me kind
words and advice (and medicine) when I needed them most. And to the entire ISKCON Boston devotee
community; if it weren't for their love and transcendental spirit I'm not sure I would have endured my time here.
To my best friend Arnika, who is my soul sister: I cannot overestimate your wonderful presence in my life over the
past few months.
And lastly, I'd like to thank my loving, caring and devoted finance, Prem Murti das, who has been a source of
inspiration and balance in my life and who has taken care of me literally in sickness and in health both here and
from Croatia. I am forever grateful and consider myself truly blessed, and I cannot wait to start the new chapter of
my life with you!
4
Table of Contents
A bstract ...................................................................................................................
2
Dedication .............................................................................................................
3
Acknowledgements.................................................................................................4
T able of C ontents.......................................................................................................5
List of Figures and Schemes........................................................................................6
Chapter 1: Introduction to the chirped-pulse millimeter wave technique...............................7
1.1 History of rotational spectroscopy techniques.......................................................7
1.2 The millimeter wave region...........................................................................10
1.3 A truly broadband millimeter wave spectrometer.....................................................11
Chapter 2: HCN/HNC.............................................................................................16
2.1 Importance of the four-center mechanism in photolysis of vinyl cyanide.........................16
2.2 H-D scrambling possibility ..............................................................................
25
2.3 Analysis of hyperfine splitting in HCN................................................................26
2.4 Fitting accuracy........................................................................................29
2.5 Resolution of the spectra...............................................................................32
Chapter 3: HCN and HNC vibrational population distribution in the pyrolysis reaction.............35
3.1 Comparison of the intensities of vibrational satellite transitions in photolysis vs. pyrolysis.. .35
3.2 Relative abundance of HCN and HNC pyrolysis products.......................................36
3.3 Cooling in an adiabatic expansion: comparing the experimental results of pyrolysis vs.
photolysis.......................................................................................................39
3.4 Relative abundance of HCN and HNC pyrolysis products.......................................41
3.5 Vibrational cooling vs. rotational cooling and connection to slit type.........................42
Chapter 4: HNO as a product of thermal unimolecular decomposition of methyl nitrite.......43
4.1 Thermal decomposition of ethyl nitrite.................................................................43
4.2 Nitric oxide radical abstracts hydrogen and deuterium..........................................
44
Appendix ................................................................................................................
50
References............................................................................................................
51
5
List of Figures and Tables
Figure 1.1a
Classical Balle-Flygare Fourier-transform microwave spectrometer........................8
Figure 1.1b
Fourier transform of molecular free induction decay..........................................9
Figure 1.2
Chirped-pulse millimeter wave (CPmmW) spectromer..........................................12
Figure 2.1a
Photodissociation of acrylonitrile via three-center elimination mechanism..................16
Figure 2.1b
Photodissociation of acrylonitrile via four-center elimination mechanism................17
Figure 2.1c
Isomerization of the HNC molecule............................................................18
Figure 2.2
Schematic potential energy diagrams for paths leading to HCN and HNC..................19
Figure 2.3
Photolysis of vinyl cyanide at 193 nm.............................................................21
Figure 2.4
Schematic for 3- and 4-center reaction products of mono-deuterated vinyl cyanide........22
Figure 2.5
Photolysis of vinyl cyanide and its mono-deuterated isotopologue at 193 nm................23
Figure 2.6
Spectrum of DCN from photolysis of mono-deuterated vinyl cyanide at 193 nm...........24
Figure 2.7
Fitting of HCN photolysis spectrum to an exponential curve................................28
Figure 2.8
Quadrupole hyperfine structure of HCN vibrational ground state.............................30
Figure 2.9
Quadrupole hyperfine structure of HCN vibrational excited state...........................31
Figure 2.10
A close-up fit of vibrationally excited HCN....................................................33
Figure 2.11
Demonstration of 10 kHz resolution via varying the center frequency in Fig. 2.10.........34
Figure 3.1
Schematic diagram of the pyrolysis nozzle....................................................36
Figure 3.2
Pyrolysis of vinyl cyanide (1% in argon).........................................................38
Figure 4.1
Proposed for mechanism for acetaldehyde plus H radical to produce formaldehyde........44
Figure 4.2
CPmmW spectrum of DNO from pyrolysis of 3% acetaldehyde and 5.5% deuterated
m ethyl nitrite...................................................................................................46
Figure 4.3
CPmmW spectrum of HNO from residual H atoms in the chamber..........................48
Table 2.1
Details of quadrupole hyperfine structure of HCN transitions in Fig. 2.3..................27
6
Chapter 1: Introduction to the chirped-pulse millimeter wave technique
1.1 History of rotational spectroscopy techniques
Molecular rotations do not require much energy to excite them and therefore most
molecular rotational transitions occur in the microwave and infrared regions of the
electromagnetic spectrum [1]. A pure rotation spectrum only occurs for molecules that possess a
permanent electric dipole moment. Generally, rotations are about the center of mass of the
molecule in order to conserve angular momentum along the rotation axis [1]. Historically,
microwave and infrared technologies were created in the 1940s, spurred by World War II efforts
to develop microwave oscillators and understand radio frequency (RF) technology [2]. This in
turn allowed for the development and evolution of microwave and infrared spectroscopy to study
molecular rotations.
The molecular beam, Fabry-Perot, Fourier transform microwave (FTMW) spectrometer
by Balle and Flygare in 1979 (Fig. 1.1 a) revolutionized microwave spectroscopy and along with
modifications to improve resolution and sensitivity, still dominates the field today [3-5]. It
performs microwave spectroscopy in the time-domain, providing sensitive detection of the
rotational free-induction decay (FID) that follows the microwave pulse (Fig. 1. 1b). The use of
time-domain spectroscopy increased the frequency resolution by avoiding power broadening or
line-shape distortion that was characteristic of the older modulation spectrometers [7]. Also,
pulsed molecular beam sources contributed other benefits such as avoiding collisional
broadening and providing rotational cooling so that larger molecules could be studied [6].
However, even though this spectrometer is able to operate over a wide range of frequencies (-10
GHz), the cavity must be precisely tuned to resonance at each frequency step and a narrow
7
frequency region is measured (-500 kHz) before it must be tuned again for the next frequency
step. This leads to very long acquisition times and a limitation on many applications due to the
possibility of long-term laser frequency drift. This may include a limitation on performing
rotational spectroscopy of excited vibrational states [7], which is crucial to providing insight into
the mechanisms of fundamental chemical reactions. Information regarding the transition state is
embedded in the vibrational and rotational population distributions of the reaction products and
can be disentangled by using the well-established transition-state theory (TST) to describe the
progression from products to reactants [8-9].
70 cm
---
Pulsed
nozzle
---.------
Cavity mirrors
Figure 1.1a The classical Balle-Flygare Fourier transform microwave spectrometer [10]. Diagram shows
pulsed nozzle and cavity mirrors which must be mechanically adjusted at each frequency step during the
scanning process.
8
I-
oc
S
Ar CaTier
0
0256
S12
Tune [m croseconds]
2319S.24
23196.74
Frequrny [MHz]
23 1".24
Figure 1.1b The free induction decay of the polarized sample is measured in the time domain (top figure)
and then is Fourier-transformed into the frequency domain to give the resulting rotational spectrum
(bottom figure) (Figures adopted from reference [10]).
Pate and co-workers recently pioneered a truly broadband rotational spectroscopy
technique in the microwave region: chirped-pulse Fourier-transform microwave (CP-FTMW)
spectroscopy (over 10 GHz per pulse in the 7.5-18.5 GHz spectral range) [7]. They achieved this
by taking advantage of advances in electronics such as the arbitrary waveform generator (AWG),
which creates a "chirped pulse," a few micro-seconds long phase-reproducible pulse with a
linear-in-time frequency sweep. The frequency range of this pulse depends on the number of
gigasamples per second (GS/s) of the AWG. The FID emitted by molecules at their transition
frequencies is then collected and subsequently digitzed by using a high-speed digitizer in a fast
digital oscilloscope. Both of these features allow for fast acquisition times covering a broad
frequency range in a short period of time. Other major advantages over the traditional BalleFlygare spectrometer are: (1) FID is averaged until the desired signal to noise is reached; (2)
9
there are no mechanical parts to scan because of the purely electronic nature of the spectrometer;
and (3) the relative intensity of spectral lines becomes meaningful within the spectrum as well as
between spectra taken under the same experimental conditions, since the scan over the entire
frequency range is covered practically at once.
1.2 The millimeter wave region
For small molecules with large rotational constants (small moments of inertia), the
millimeter wave (mm-wave) region (30 - 300 GHz) proves to be very important for making
measurements on these molecules [11-13]. Our specific technique, adapted from the CP-FTMW
spectrometer established by Brooks Pate, takes advantage of fast electronics to overcome two
major issues that have plagued mm-wave spectroscopy: problems with direct generation and
detection of mm-waves and absence of truly broadband capabilities. We have developed a
CPmmW spectrometer that can operate within 55-102 GHz, encompassing most rotational
transitions of molecules containing two heavy atoms.
Classic issues with mm-wave generation come from both the long and short wavelength
end of the spectrum. From the long end, microwaves are about 1-3 cm in length, for which
electronic oscillators such as magnetrons and other microwave sources are used to work within
this region because they require that the wavelength be comparable to the dimensions of the tube
[1]. However, for the mm-wave region, this condition is no longer satisfied. From the short end,
hot bodies are used to emit infrared radition and the intensity of the radiation scales with the
inverse square of the wavelength [1]. Therefore, the closer one gets to the mm-wave region, the
less power there is. And even with the advancements in electronics, transitions must be relatively
strong to be observed in a broadband spectrum. The downside of looking at higher frequencies is
10
that because of the lack of power it will be more difficult to polarize transitions with high dipole
moments or small population. One must then surrender to narrowing the bandwidth to increase
the degree of polarization [1].
1.3 A truly broadband millimeter wave spectrometer
Advances in broadband millimeter-wave amplifiers, arbitrary waveform generators and
heterodyne receivers allow the circumvention of difficulties in the mm-wave region [14]. It is
possible because of advancements in amplifiers and active frequency multipliers to up-convert
the pulse generated by the AWG into the mm-wave region as well as to down-convert the
resulting polarized pulse to allow for easy detection. The schematic of the chirped-pulse
millimeter wave (CPmmW) spectrometer designed in the Field lab is shown in Fig. 1.2.
11
i
i
0.2 - 4 GHz
Chirped pulse
iii
ix
xi
v
LO
xv
N+ 0.01 GHz
20 mW
xiv
xx
55-73 GHz
Or
87-102
GHz
xii
Molecular
Beam
Figure 1.2 Schematic diagram of the CPmmW instrument used for the experiments herein. For a full list of parts, see the Appendix.
All components of the experiment are phase-locked to the same 10 MHz Rubidium frequency standard (i). An 8 Gs/s AWG (iv),
operating at the 8.75 GHz rate of an external clock (iii), is used to generate a linearly-chirped pulse, which is up-converted (vi) by
mixing with the output of a fixed-frequency 9.375 GHz oscillator (v). The resultant signal is isolated (vii) and amplified (viii). The
desired sideband is then selected using a bandpass filter (ix) and actively frequency-multiplied by a factor of 8 (x and xi) to produce a
chirped pulse that covers the 55-73 GHz or the 87-102 GHz frequency range with a peak power of 20 mW. The precision attenuator
(xii) is introduced for intensity calibration purposes. The mm-wave pulse is coupled into free space using a 24 dBi standard gain horn
(xiii) and focused into a molecular beam chamber by a pair of teflon lenses (xiv). After the chirped excitation pulse has polarized the
12
molecular sample, the FID is collected and down-converted (xviii) by mixing with the output of a Gunn oscillator (xvi). The resultant
signal is input to a low-noise amplifier (xix) and averaged in the time domain on an 8 GHz oscilloscope (xx). Figure adopted and
modified from Park, et al [14].
13
Briefly, it operates as follows: An 8 GS/s AWG (Tektronix model AWG7082C)
generates customized chirped pulses that are then mixed with a fixed 9.375 GHz phaselocked local oscillator (LO). This resulting frequency is then isolated, amplified, and
passed through a band-pass filter to select the sum or difference frequency, respectively.
For example, the AWG can generate chirps from 0-4 GHz in theory (the Nyquist
frequency is half the sampling rate, which is 8 Gs/s). This frequency is mixed ("upconversion") with the LO fixed at 9.375 GHz, resulting in two frequency ranges: 5.375 9.375 GHz or 9.375 - 13.375 GHz. We can choose either the lower side band or upper
side band by using band pass filters (BPFs). We have three BPFs that transmit 6.875 -
9.125 GHz, 8.7 - 10.5 GHz, and 10.9 - 12.7 GHz. These ranges refer to the flatness
regions of the BPFs and it is very important that when choosing which BPF to use, any
residual microwave power at the carrier frequency of the LO is blocked (therefore we
cannot use the second BPF mentioned, but will use the first and third). These ranges are
then multiplied by eight in the multiplier chain, at last generating radiation in the mmwave region. Depending on whether we want to choose lower or upper side bands, the
frequency ranges possible are 55 - 73 GHz and 87.2 - 101.6 GHz. We are also limited by
the frequency range on our oscilloscope, which is 8 GHz (Tektronix DP070804C, 25
GS/s Digital Phosphor Oscilloscope). Therefore, we choose a center frequency that will
be appropriate for encompassing the transition frequencies we are interested in. The mmwave frequency is broadcasted from a horn antenna and, with Teflon lenses, is loosely
focused in the center of the chamber where it intersects the molecular beam. The
CPmmW spectrometer is matched with the pulsed supersonic molecular beam setup via a
master 10 MHz clock. Therefore, each shot of the chirped pulse identically polarizes each
14
beam pulse and all molecular transitions within the chirped pulse's sweep range become
polarized. Each of the polarized two-level systems undergoes coherent radiation at their
transition frequency, or FID. The FID signal lasts for about 10 ps and is limited by
Doppler dephasing. It is then mixed with another LO to a lower frequency ("downconversion"), amplified with a low-noise amplifier, then recorded in real time by our fast
8 GHz oscilloscope (Tektronix DP070804C). In this case, the FID is mixed with the
78.76 GHz (9th harmonic of the Miteq 8.75 GHz DLCRO phase-locked oscillator) output
of a phase-locked Gunn oscillator. The last step is Fourier-transforming the signal read by
the scope into the frequency domain to give the rotational spectrum.
A major advantage of recording the entire spectrum in one shot and having it
synchronized with the molecular beam pulse, is that if the signal to noise (S/N) of a
single-shot spectrum is not acceptable, then FID signals from multiple shots can be
averaged phase-coherently in the time domain before being Fourier transformed into the
frequency domain. In this way, both high-resolution and high S/N can be achieved while
uncovering weaker transitions or line characteristics that may not have been apparent in
the single-shot spectrum. Also, as was previously noted, another major benefit of singleshot spectroscopy is that the relative line intensities of the transitions become meaningful
(within 3%). Because of this, both frequency and intensity data can be used as input into
different fit models, an impossible thing to do until chirped pulse spectroscopy.
15
Chapter 2: HCN/HNC vibrational population distribution in the
photolysis reaction
The following work was done in very close collaboration with and guidance from Dr.
Kirill Prozument.
2.1 Importance of the four-center mechanism in photolysis of vinyl cyanide
In the photolysis of vinyl cyanide with 193 nm laser radiation, two qualitatively
distinct reaction paths were proposed by Derecskei-Kovacs and North via their ab initio
molecular orbital calculations [15]. The first mechanism passes through a three-center
transition state to produce hydrogen cyanide (HCN) and singlet vinylidene, and the
second mechanism passes through a four-center transition state to produce hydrogen
isocyanide (HNC) and acetylene (Figs. 2.1 a, b, c).
TS
H2
H3
HI
C1
H2
C2
H2
0
C1 C2
H3
H1
H1
C3
C3
H3
N
C3
N
Figure 2.1a Photodissociation of acrylonitrile molecule via the three-center elimination
process, as calculated by high level ab initio theory [15]. Hydrogen atom H3 embeds
itself between the carbon C2 of the acetylene group and C3 of the cyano- group. The
transition state involves formation of two new bonds shown by dashed lines - those
between C2, C3, and the "traveling" H3 atom. Therefore, this mechanism is usually
called "three-center elimination." The reaction proceeds from the transition state by
16
eliminating the HCN molecule and the vinylidene radical. In most cases, the latter has
been assumed to quickly isomerize to form acetylene (Fig. 2.1c). There is little doubt in
the literature that this process is one of the major channels for the photodissociation of
the acrylonitrile molecule.
TS
H2
C1
C1
C2
C2
C1
H3H2
H3
H2H3H2
H3H2
H3
C2
CC 2
3
U.3
HI
H1
C3
N
N
HI
N
C3
Figure 2.1b Dissociation of acrylonitrile molecule via the four-center elimination
process, as calculated by high level ab initio theory [15]. The cyano-group and the
opposite hydrogen atom, HI, approach each other and form three new bonds in the
transition state. In the transition state a bond is created between four atoms, Cl, C3, Hi,
and N. Thus this decomposition mechanism is called "four-center elimination." The
resulting products are acetylene and HNC molecules. The reaction is commonly believed
to proceed with HNC isomerization to the more stable HCN isomer, as shown in Fig. 1
c).
17
TS
Figure 2.1c Isomerization of the HNC molecule [15]. The HNC -> HCN isomerization is
believed to take place subsequent to the four-center elimination step of acrylonitrile
photodissociation. This is a fundamental isomerization reaction that proceeds via
delocalization of the hydrogen atom during its large amplitude motion around the N-C
core. The assumption is that there are collisions with the carrier atoms, which can
accommodate the energy released during isomerization.
The authors concluded by their RRKM calculations and the predicted HCN/HNC
ratio that the three-center mechanism path should have a yield 124 times greater than that
of the four-center one, thereby predicting that the production of HNC will be irrelevant in
comparison to that of HCN. This initial proposal was later modified by electronic
structure calculations done by Homayoon, et. al [16], who recognized that HCN and
HNC can both be produced by three or four center mechanisms (Figs. 2.2 a, b). Thus they
reduced this ratio to 1.9, still in favor of HCN production, predominantly via a threecenter mechanism, and HNC via a four-center mechanism. (All above figures are taken
from an ACS PRF research proposal by Dr. Kirill Prozument).
18
Path IV
Path I
a.
b.
Tl-
~
100.6
4
A
M NP th
T3 S-
:CCHa + HCN
vc
Path V
Path II
TE -V
CHHC
kW2-V
405A
116.0
39..
M-C HWC
>VC
HCCH + He"
1 rVC
Path III
Paths VI and VI1
T82-U
%o
sT8as
TA4
57A
J*
VC
CCHj+ WNC
3.4
accue iwc
[j
HCCH + HCN
Figure 2.2 These are schematic potential energy diagrams (PEDs) for paths leading to HCN
production (panel a) and for paths leading to HNC production (panel b). calculated by
Hoyamoon, et al [16]. As both Fig. 2.2 a and b show, HNC and HCN can both be produced from
either three or four center mechanisms. Shown are the relative energy values in kcal/mol. Figures
are taken directly from reference [16].
19
To address the reaction mechanism of photolysis of vinyl cyanide (VCN) we
carried out a series of CPmmW experiments. The ArF excimer laser (Lambda Physik
COMPex) operating at 193 nm with output power of about 30 mJ/pulse was used to
photolyze VCN. The resulting CPmmW spectrum of the 87-93 GHz region is shown in
Fig 2.3. HCN, and to a lesser extent HNC, are produced in different vibrationally excited
states that are comparable in relative population to their respective ground states.
20
60-
HCN (0000)
4
*
vibrationally excited HCN
+ vibrationally excited HNC
40
HNC (0000)
*
E
E
*t*
C20
*++
0
87
88
89
91
90
Transition frequency, GHz
93
92
2012-W14005
Figure 2.3 Photolysis of vinyl cyanide at 193 nm. Vibrational satellites of HCN and HNC in this millimeter wave spectrum are
indicated by the * and + labels, respectively. In a triatomic molecule, the bending vibrations are expected to increase the rotational
constant and vice versa with stretches, so lines to the right of the ground state transition are expected to be bending vibrations and
lines to the left to be stretching vibrations. The combination vibrations can go either way. The lines assigned to HCN are assigned due
to the characteristic hyperfine splitting of HCN [17]. The very intense lines are the precursor lines of vinyl cyanide.
21
In order to get additional insight into the transition states of the photolysis
reaction, we have used isotopically substituted precursor species. In this way, we are able
to determine which mechanism is responsible for the observed products. Due to the sitespecific labeling, as shown in Fig. 2.4, we can assume that most of the HCN and HNC
produced is via the 4-center mechanism while most DCN and DNC produced is via the 3center mechanism.
4-center
HCN or HNC by 4-center
N
H
+
3-center
C==C
H
mechanism
DCN or DNC by 3-center
mechanism
D
Figure 2.4 Shown as the reactant is a di isotopologue of vinyl cyanide. After photolysis
of this isotopologue, all HCN and HNC will be necessarily produced by the 4-center
mechanism while all DCN and DNC will be from the 3-center mechanism.
Fig. 2.5 shows the CPmmW spectra of photolysis products of normal vinyl
cyanide CH2=CH-CN and its isotopologue CH 2=CD-CN in panels a) and b) respectively.
The spectra were recorded under the same conditions; therefore the relative line
intensities are meaningful within each spectrum as well as between the two spectra.
Interestingly, the two spectra demonstrate very similar vibrational population
distributions (VPD) and overall intensities of the lines. If the three-center mechanism is
the dominant reaction path, then hardly any HCN would be expected to be produced and
DCN and DNC would be the primary reaction products. However, not only was HCN
22
directly observed in the same vibrational states as in the previous spectra, but the
intensities are very comparable between the two spectra; the HCN lines from deuterated
vinyl cyanide are about 80% as intense as the HCN lines from normal vinyl cyanide.
These results imply that the four-center mechanism is the major path for production of
HCN, the three-center one being much less important unless H-D scrambling is
considered (see next section on H-D scrambling possibility).
30
20
> 10
0
E
E
0-30
20
10
07.
87.5
88.0
88.5
89.0
89.5
90.0
Transition frequency, GHz
90.5
91.0
Figure 2.5 The 193 nm photolysis of vinyl cyanide and its mono-deuterated
isotopologue, CH 2=CD-CN, shown in panels a and b, respectively. In panel a, HCN and
HNC are produced by both the three and four-center mechanisms. In panel b, due to the
isotope substitution at the three-center position, HCN and HNC are exclusively produced
via the four-center mechanism, assuming no scrambling effects. Due to the identical
23
conditions under which the spectra were recorded, the intensities can be directly
compared. The parent lines of HCN and HNC in the substituted spectrum are roughly
80% as intense as in panel a, revealing the dominance of the four-center mechanism in
both HNC and HCN production.
This conclusion was further supported when the spectrum was taken of the DCN
produced from deuterated VCN (Fig. 2.6). Indeed, the ground-state DCN had a relative
intensity of 16% that of the vibrational ground state HCN line. According to Fig. 2.4,
DCN can only be expected to be created via the three-center mechanism and the lack of
DCN product directly supports the four-center over the three-center mechanism, and
effectively rules out scrambling. Despite the uncertainty regarding rotational cooling, we
do not expect a significant difference in rotational cooling between HCN and DCN. We
thus regard the intensities to be comparable under similar conditions.
DCN (040)7
DCN (000)
DCN (02*0)
0-
5
0 72.5
73.5
73.0
Frequency, GHz
74.0
2012-W21
Figure 2.6 193 nm photolysis of mono-deuterated vinyl cyanide (top). Bottom spectrum
is background.
24
To the best of our knowledge, the only other experiments that have been done to
measure the ratio of three-center versus four-center mechanisms was the work done by
Wilhelm, et al. They measured, via Fourier transform time-resolved infrared emission
spectroscopy, HCN + vinylidene, which they assumed came from a three-center
elimination channel versus HNC + acetylene, which they assumed came from a fourcenter elimination channel [18]. Our experiments show that it is the four-center
mechanism that is the dominant one in producing both HCN and HNC in the photolysis
reaction of VCN. It would be interesting to explore the geometry of the four-center
transition state that results in both HCN and HNC products. Homayoon, et. al were able
to also arrive at their predicted ratio of 3.3 by summing over the three-center paths and
the four-center paths they derived from their electronic structure calculations. (Fig. 2.2)
2.2 H-D scrambling possibility
There have been reports on the scrambling of H and D atoms in laser-excited
molecules [19-21]. The possibility of H-D scrambling in vinyl cyanide prior to its
dissociation will be discussed here as this effect could provide an alternative explanation
for our results. In the gas phase, H-D scrambling may be initiated, resulting in
unimolecular H-D exchange of the intermediates before they fragment into products.
However, we do not believe that this effect is relevant in the present study. If it were, we
would expect to observe much more DCN and DNC as scrambling of H and D would
mean that mono-deuterated vinyl cyanide would make DCN and DNC by the three center
mechanism at least one-third of the time if one assumes complete scrambling, but this is
not even remotely what is observed. As previously stated, the production of DCN and
DNC is about 10% of the total products, not near enough to 33% to suggest significant H25
D exchange. Experiments that have observed scrambling via laser excitation, including
193 nm [20], which provides an excitation energy significantly higher than the energy
threshold for breaking of C-H and C-D bonds, the scrambling effect was modeled by one
or two 1,2- or 1,3-H/D shifts between near-lying isomers on the potential energy surface.
This implies that a special combination of vibrational energy levels is required.
Therefore, we can conclude that our results reveal the four-center mechanism to be the
dominant pathway for HNC and HCN production in vinyl cyanide.
2.3 Analysis of hyperfine splitting in HCN
Bechtel, et al [20] demonstrated how the quadrupole hyperfine structure (hfs)
could be used as a sensitive electronic reporter to reveal the degree of bending excitation
in HCN +-+HNC isomerization. In other words, by measuring the hfs, it should be
possible to explore how the chemical bond has been transformed when we excite a
vibration that is coupled to the isomerization reaction coordinate. In general, the
quadrupole effect arises from the non-uniformity of the electronic environment of the
nucleus, in which the energy of non-symmetric I> 1 nucleus is orientation-dependent.
When this nuclear electric quadrupole moment interacts with the non-uniform electric
field at that nucleus (specifically, the gradient of the E-field), rotational levels split into
several components. The hyperfine splittings were analyzed by fitting the peaks with the
Lorentz formula and the table of results is shown in Table 2.1.
26
Peak
Assignment
1
2
3
4
Transition frequency
col, MHz
co2, MHz
F=1-1
F=2-1
87418.24
87843.29
87870.03
87882.71
87419.55
87844.72
87871.53
87884.14
88005.5
88006.88
88026.07
88466.9
88487.52
88027.46
88468.35
88488.98
88631.85
88933.72
o3, MHz
F=0-1
Hperfine splitting
m2-col,
o3-co2,
MHz
MHz
Peak area
Al
A2
1.31
1.43
1.5
1.43
0.334
0.207
0.083
0.385
0.532
0.445
0.502
0.73
1.38
0.287
Relative peak area
A1/A2,
A3/A2,
A3
62.7
46.5
16.6
52.7
100 (C-H
5
7
stretch)
001 (C-N
stretch)
021
8
120
9
10
11
12
13
000 (vibr. g. s.)
6
020 (bend)
14
15
16
17
040 (bend)
060 (bend)?
18
080 (bend) ?
19
20
21
22
23
24
,
88630.43
88932.22
88996.58
89086.44
89426.77
89537.1
89567.88
89958.59
90080.03
90115.62
90628.34
90741.92
91221.97
91430.47
91873.83
, 92603.42
88029.53
88470.48
88491.12
88633.93
1.39
1.45
1.46
1.42
1.5
1.54
88998.12
89087.92
89428.32
89538.65
89569.43
89960.23
90081.6
90117.22
90629.98
90743.59
91223.64
91432.17
91875.53
92605.12
89090.09
1.48
1.55
1.55
1.55
89571.68
2.07
2.13
2.14
2.08
90632.33
91226.08
1.7
10.82
0.203
0.469
2.17
3.99
0.444
2.25
1.64
1.57
1.6
1.64
1.67
1.67
1.7
1.7
0.875
3.44
3.29
2.35
2.44
0.539
1.95
0.277
0.514
0.22
1.51
0.304
0.724
0.198
0.169
0.171
1.64
5.39
4.87
18.12
0.494
1.01
0.272
1.01
0.903
3.36
1.42
0.445
16.6
18.7
18.5
18.5
41.0
46.4
6.94 0.839
1.03
0.992
3.6 0.502
0.416
0.411
2.67
0.395
1.29
0.302
0.431
53.3
63.8
67.6
59.7
0.436
0.224
57.5
43.1
54.3
54.2
66.6
36.2
53.4
56.6
76.9
56.1
65.7
39.2
38.5
12.1
13.9
16.3
17.4
Table 2.1 Details for the hfs of the HCN transitions indicated in Fig. 2.3. For 06Q and 080, assignments are tentative, see Fig. 2.7.
27
The stretches are on the order of about 3000 cm-1 (C-H stretch) and 2000 cm-1 (CN stretch) while the bends are around 700 cm-1. As Fig. 2.7 shows, the exponential curve
fitting the bending modes implies that the bending modes are nearly thermally
distributed. If the spectrum reflected a purely thermal population, then we would expect
less intensity for higher energy stretches and for bends with higher quanta. However, the
assignment of the lines 7 and 8 from Table 2.1 to be the combinations of 0200 (bending
vibration) with the two stretches 0001 or 1000, to result in 0201 and 1200, respectively,
suggests that this is indicative of a unique detail of the reaction mechanism. The
combination bending and stretching levels makes these transitions the most energetic, and
yet the rotational transitions are 2-4 times as intense as they are in the pure stretching
vibrations.
193 nm photolysis of VCN, 1.5 cm laser - slit jet distance
.............. .........
......... ......... -rr....
HCN vibrational modes
HCN (000)
30 - v, -C-H stretching
V-. bend
Y -C-N stretching
20
0
HON
HCN (100)
88.0
88.5
(
HNC (000)
2*0)
HCN (0800) ?
HCN (04*0)
HCN*
10
87.5
HCN
89.0
89.5
HCN (060) ?
90.0
90.5
Transition frequency, GHz
91.0
m-2
Figure 2.7 An exponential curve has been arbitrarily fitted to the vibrational ground state, 020,
and 040 in order to evaluate other assignment possibilities. The tentative 060 assignment does not
seem as likely as that for 080 if the bending modes are indeed in a thermal distribution, which
appears to be the case.
28
2.4 Fitting accuracy
Figure 2.8 shows the hyperfine structure of the vibrational ground state of HCN
and the resulting multiple-peaks fit. The fittings for less intense lines (Fig. 2.9) seem to
be slightly better than those for the most intense lines because of the valley between
peaks is not so high and as is apparent in the figure, some shoulders of the most intense
peak are left out of the fitting.
29
Model
Equation
60
I
I
I
I
I
I
i
I
I
I
I
I
I
I
I
I
I
I
I
I
I1I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
40
30
20
E
E
0
Reduced Chi-S
qr
Adj. R-Square
115739
0 9613:
Value
50
C
Lorentz
y - yO + (2*A/PI)*(w/(4*(x-xc)
24.wA2))
10
eaki ()
Peak1 (B)
Peak1<B)
Peakl(B)
Peaki(B)
Pek2(B)
Peak2(B)
yO
xc
w
A
H
y0
xc
.68785
883043
1.97265E-4
0.01082
34.90588
0.68785
88.63185|
:ek2(B)
Paak2(B)
Peak2(B)
Peak3(B)
Peak3(8)
P--k3(B)
Peak3(B)
Pak(B)
w
A
H
y0
xc
w
A
H
2.0232BE-4
0.01812
57.01621
0.68785
8&.63393:
1.84362E-4
0.00336
11.616
Standard Error
006666
323882E-6
8.48931E-8
3.44881E-4
0.0668
1.87216E-6
5.78291 E-6
3.57215E-4
0.06666
8.42644E-6
2.89066E-5
3.45082E-4
0
I . I I1 I
. I I
88.620
I * I I I I I I
I
a I a A ,
88.625
,
I I I I I I I I I I
88.630
,
, -
I
I
I , ,
88.635
Transition frequency, GHz
I I I I
88.640
,
II
88.645
2012-00-00_000
Figure 2.8 The quadrupole hyperfine structure of the HCN vibrational ground state. The fitting underestimates the valleys as well as
the shoulders of the individual peaks. This occurs in varying degrees for each peak that has been fitted.
30
3.0
> 2.5
-Ad.
-
Rsquer
Peak1(.
Poak1(6
LPask1(6)
-
cu
2.0
15
Peak1 (B)
Pk2B
Pea)
-Peak2(B)
2
Peak2(B)
0.
Aa()
y
:c
wA
H
y0.382
xc
w
H
Value
Standard Error
0.01213
88.93222
1.71929E5
1.02242E-4
4.991E-5
2052.
.98
1.29253
0.01213
88.93372
8.98854E6
1.27599E.4
1.53&33E5
5.19273E-5
4.40E4
2.49164
O.38852
E
E 1.0-
0.5 -
0.5
0 .0
-
A
I
I
I
A
1.
-1
88.930
1
88.935
Transition frequency, GHz
-
#
I -
I
I
88.940
2012-00-00_000
Figure 2.9 The quadrupole hyperfine structure of HCN vibrational excited state (transition no. 10 in Table 2.1). This transition is
much weaker than that of the vibrational ground state (see Fig. 2.8) and the fit seems to be much more accurate. It is observed as a
general trend that the weaker lines tend to be more accurately fitted than the more intense lines.
31
2.5 Resolution of the spectra
The resolution of the spectra seems to be in agreement with the +/- 10 kHz from
the fit as demonstrated in Figs. 2.10 and 2.11. Fig 2.10 shows the original spectrum
zoomed in on line 12 (from Table 2.1) and Fig. 2.11 shows the same line but with the
center frequency parameter fixed 10 kHz from the original fit. It is clear that there is a
significant difference in the fit agreement between the two spectra.
32
1.5
C
1.0
E
E
.. 0.5
0 .0
1-
89.9575
, , , , , , , , I , , ,
89.9580
I , ,
89.9585
, I ,I ,
i
, , .,,
89.9590
Transition frequency, GHz
I
I
I
89.9595
2012-00-00_000
Figure 2.10 A close up of the original fit of vibrationally excited HCN (the first splitting from peak 12 from Table 1/Fig. 2.3).
33
2.0
Re d
i"
qr
Adj. R-Square
Peakl(b)
>Peaki(B)
1..5 -....
0)
1.0
5
E
E
-
..
Peakl(B)
Peakl(B)
eak(B)
Peak2(B)
.ak2(S)
Poak2(8)
Pea2(B)
0.0413
?
I*I
I
I
I
I
I
I
I
I
0.458
Standard Error
Value
0.00879
0.29699
0
89.85858
. 6
. ........
4.25341E-5
2.71 979E-4'
A
1.20841
H
0.00879
yO0.29699
6,24892E-6
89.96023
xc
2.49879E-5
1.30157E-4
w
4,16003E-4:4153A
2.03475
H
yd
xc
S0.5
89.9575
89.9585
89.9580
Transition frequency, GHz
89.9590
89.9595
2012-00-00_000
Figure 2.11 The same line as in Fig. 2.10, but now the center frequency was fixed 10 kHz away from the original fit at 89.95858 GHz
instead of 89.95 857. Compared to Fig. 2.10 there is a noticeable difference in the quality of the fit to the observed transition line.
Therefore we conclude our resolution precision is +/- 10 kHz.
34
Chapter 3: HCN and HNC vibrational population distribution in
the pyrolysis reaction
The following work was done in very close collaboration with and guidance from Dr.
Kirill Prozument.
3.1 Pyrolysis reactions studied by CPmmW spectroscopy
Pyrolysis is the thermal decomposition of molecules in the absence of oxygen
[22]. Characterization of the transient molecules in pyrolysis by spectroscopic methods
has important implications for modeling of combustion and atmospheric chemistry. In
pyrolysis, the molecules pass through a heated nozzle, (Fig. 3.1) and subsequently
undergo supersonic expansion which thereby cools the pyrolysis products so that they
may be studied spectroscopically.
Barney Ellison and coworkers is one of the groups which advance understanding
of basic chemistry with pyrolysis experiments. They have been studying pyrolysis of
nitrites by using a Chen nozzle in conjunction with photo-ionization mass spectrometry
and matrix isolation FTIR spectroscopy to disentangle the pyrolysis mechanisms of
organic molecules that are significant to biomass decomposition [23-24].
With our implementation of CPmmW spectroscopy, we are in a unique position to
look at two-heavy-atom molecules such as formaldehyde, HNO, HCN, HNC, DCN,
DNC, C 2H, and others. In a cold molecular beam, such molecules have their most intense
J = 0 - 1 rotational transitions within the range of our spectrometer of 58 - 102 GHz.
35
Horn antenna
(receiving)
Electrodes
Focusing lens Supersonic jet
ePrecursor
molecules
in Ar inlet
Hot reactor tube
Horn antenna
(source)
mm-wave beam
Focusing lens
Figure 3.1 This is a schematic diagram of the widely used pyrolysis nozzle designed by
Chen et al [25-26]. The tube is 4 cm long and 1 mm inner diameter, made of silicon
carbide, SiC (Saint-Gobain Ceramics, Hexoloy SE). Figure by Dr. Donald David,
University of Colorado.
3.2 Comparison of the intensities of vibrational satellite transitions in
photolysis vs. pyrolysis
HCN and HNC were produced by pyrolysis of vinyl cyanide (Fig. 3.2). In contrast
to the CPmmW photolysis spectrum of vinyl cyanide, we do not detect any vibrational
satellites for HCN or HNC. The 000 line of HCN is very intense and there are no satellite
transitions with a signal-to-noise ratio within a factor of 20 weaker than the vibrational
ground state transition, whereas the intensities of the satellite lines in the photolysis
spectrum were comparable to that of the 000 line within a factor of three in intensity.
36
When a molecule breaks, there is a change in geometry for the fragments, and these
fragment molecules can be left vibrating in states that indicate the geometry changes that
occur as the molecule traverses the fragmentation transition state. Thus we hope to gain
insight into the pyrolysis mechanism by observation of the relative intensities of the
vibrational satellites. For example, formaldehyde was observed in an out-of-plane
bending vibrational level, which we initially believed was indicative of the fragmentation
mechanism. However, when pure formaldehyde was put through the hot pyrolysis nozzle,
a similar vibrational population distribution was observed, which therefore revealed that
the population in the out of plane bending vibration is determined primarily by the
collisional cooling (or lack thereof for this particular mode) during the supersonic
expansion from the nozzle.
37
I
200
I
I
I
I
I
I
~
I
I
I
IlIllIjI
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1% V-CN in Ar, 1600 C, 3 bar
unassigned lines belong to V-CN
=L 150
C
HCN
100
E
E
0..
HNC
50
U
87
[rI
88
89
Transition frequency, GHz
-
1- -0
91
90
2012-03-16_006
Figure 3.2 Pyrolysis of vinyl cyanide (1%in argon). This spectrum does not show any vibrational satellites of HCN in contrast to the
photolysis spectrum of vinyl cyanide under the same conditions (Fig.2. 3).
38
The difference between the pyrolysis and photolysis experiments described here that
affects the vibrational population distribution of products is that pyrolysis reactions occur in the
hot tube prior to expansion. In photolysis, the expansion from the slit jet occurs first and
afterwards the photofragmentation reaction occurs, thereby allowing the probing laser to not
sample the most important effects of post-reaction collisional cooling. Since we are only able to
sample populations in low-J levels, we require some collisions to rotationally cool the product
molecules but we want to minimize vibrational relaxation. To explore this further, these
photolysis experiments could be repeated in an axial jet. The slit jet is designed to be a 1dimensional expansion in which the number density of the gas decreases as l/d, where d is the
distance from the slit [27-28] whereas in the axial jet the number density follows the 1/d2 law
[29]. It is difficult to place the laser precisely within the collisional region in an axial jet
expansion to obtain rotational cooling because the collisional region is limited to a length of only
a few nozzle diameters, i.e. a few millimeters.
3.3 Cooling in an adiabatic expansion: comparing the experimental results of
pyrolysis vs. photolysis
Buffer gas is adiabatically expanded, which means that there is no exchange of energy
between the buffer gas molecules and the surroundings. To illustrate, consider a container of gas
molecules. The molecules collide in the container to equilibrate at a certain temperature. When this
gas mixture is contained, molecules collide, fly away from each other, bounce off the wall, then
return and exchange energy with other molecules, thereby continually exchanging energy by heating
up and cooling down via these collisions. But in an expansion, the "wall" is always getting further
away, so there is less chance for the molecules to return and re-heat colder molecules via collision.
The molecules are gone for good.
39
Behind the nozzle, the molecules start out having 3/2 kT of translational energy. As the
gas expands from the nozzle, all collisions are converted into forward motion (Boltzmann
velocity distribution) as the beam speeds up and consequently unidirectional binary collisions
provide extensive cooling of the translational degrees of freedom [30]. This cooling happens
since while still in collisional region, faster molecules catch up and collide with slower atoms in
front, resulting in all the fast atoms getting slower and vice versa. So what was a Boltzmann
velocity distribution in the beginning is now much cooler. The rotational degrees of freedom are
likewise cooled by rare gas molecules since the kinetic energy of the argon atoms approximately
match the rotational energy level spacing (typical B constants are a few cm 1 ).
With the slit jet it is much easier to locate the photolysis laser beam within the extended
collisional region of the expansion in order to achieve sufficient rotational cooling without
subsequent vibrational cooling. Therefore the pyrolysis measurement, where no vibrational
satellite transitions are observed, may be telling us that collisional cooling, which is severe in the
pyrolysis jet, is not responsible for the vibrational satellites in the photolysis spectrum. In other
words, in the pyrolysis jet, collisional cooling completely relaxes both stretching and bending
modes in HCN, effectively depopulating all vibrationally excited states of HCN. This is in
contrast to our observations for molecules such as formaldehyde and OCS, which are resistant to
complete vibrational cooling. For example, in formaldehyde, mode 4 is resistant to vibrational
cooling but other modes are not [to be published]. But the fact that we detect the vibrational
satellites in the photolysis spectrum implies that collisional cooling of vibrational populations is
minimal. Thus, we can be assured that we are in a low collisional regime. This is important to
note because then we can assume that the intensities of vibrational satellite transitions sample the
photolysis reaction only and not post-photolysis collisional effects.
40
3.4 Relative abundance of HCN and HNC pyrolysis products
The pyrolysis spectrum (Fig. 3.2) reflects a distribution of HCN and HNC products
resulting from a pyrolysis reaction at
Tnozzie =
1873 (± 100) K. The energy difference between
HCN and HNC vibrational ground states was calculated to be -5400 cm-1 [31]. The integrated
areas of HCN and HNC transitions in our pyrolysis spectrum, taking the splitting of HCN into
account are 0.0991 and 0.00293 (in arbitrary units), giving a ratio of [HCN]/[HNC] = 33.8. The
areas are themselves proportional to the square of their dipole moments, which are 2.94 D for
HCN and 3.05 D for HNC. Thus, [HCN] = A1 /2.94 2 and [HNC] = A2/3.05 2 . Let us assume that
the areas themselves are each proportional to the Boltzmann distribution Exp[-En/kT], such that
A1/A2
=
(3.05/2.94) 2Exp[E 2-E1/kTff], where
determine the
Teff
Teff
is some effective temperature. Calculation to
follows below:
5400 cm-1
A1
0.0991
3.05
-2= 0.00293=33.8= (9
A2
2.94)
0.00293
2
e
_
0.69 cm-1
T
ke
= 1.076*e
7769.78 KIT
IT
33.8
7769.78 K
In1
= 3.45 =
1.076
T
Thus Teff = 2252.1 K for the HCN and HNC products of the photolysis reaction, which is 379 K
higher than the temperature measured by the optical pyrometer of Tnozzie = 1600 (± 100) C, or
1873 (±100) K. One may expect some cooling of HNC molecules in the supersonic expansion,
which would result in HCN and HNC populations distributed with
Teff
<
Tnozzie.
One can argue
that the HCN population is frozen in the -11 500 cm I HNC potential well [25] at the very
beginning of the supersonic expansion, and the expected distribution is the one with Teff
Tnozzie.
We observe that the [HCN]/[HNC] ratio is slightly less than it would be expected if the HCN and
HNC states were thermally equilibrated at Tnozzie = 1873 K pyrolysis nozzle and about an order
41
of magnitude greater than it was in the photolysis measurements described in this work. The
results are suggesting that either thermal decomposition is proceeding differently via the
transition states, or that a significant amount of HNC is converted to HCN in the hot pyrolysis
tube. More pyrolysis experiments, at different temperatures, are desirable. We hope in the future
to explore the pyrolysis studies of the system to gain more insight.
3.5 Vibrational cooling vs. rotational cooling and connection to jet type
A common generalization made about cooling in a supersonic jet is that vibrations cool
about 1000 times more slowly than rotations [29]. However, such cooling effects cannot be taken
for granted and there are many additional complex factors that have to be taken into account,
such as the degree of coupling between vibrations and rotations. It is a good question about why
some molecules vibrationally cool and others do not. For example, the similarity in geometry of
HCN and OCS might make one at first assume that they would behave the same way with
respect to vibrational cooling, but we see that whereas OSC is resistant to vibrational cooling,
HCN is not. More detailed and complete studies of vinyl cyanide photolysis need to be done in
the future to check for populations in other vibrational levels and elucidate more information.
42
Chapter 4: HNO as a product of thermal decomposition of organic nitrite
molecules
What follows is a preliminary discussion of a project that deserves more work. The following
work was done in very close collaboration with and guidance from Dr. Kirill Prozument.
4.1 Thermal decomposition of ethyl nitrite
Thermal decomposition of ethyl nitrite is known to proceed through the cleavage of the
relatively weak (Ea = 42 kcal/mol) O-NO bond:
A
CH 3CH 2ONO -+ [CH 3CH20] + NO
[CH 3CH 20]
-+
CH 3CHO + H
(1)
[CH 3CH 20]
--
CH 2 CO + -CH 3
(2)
Both acetaldehyde and formaldehyde as produced by paths (1) and (2), respectively, can be
detected by our CPmmW spectrometer. The analysis of the formaldehyde vibrational population
distribution (VPD) via path (2) will be discussed in a future publication by Dr. Kirill Prozument
et al. Also it was an accidental finding by K. Prozument that H atoms may play an important role
in bimolecular pyrolysis reactions in addition to unimolecular decomposition, e.g. the proposed
mechanism for acetaldehyde to formaldehyde:
43
0i
0OH
C3 O
CH2
HCIICH
Proposed Mechanism:
0
0
II
II
~
H3 C
H
H
_
H3
1
CH
+
H
H3C~C
0
+
CH3
CH 2
Figure 4.1 Proposed mechanism for acetaldehyde to formaldehyde making use of the hydrogen
radical as a reactant.
4.2 Nitric oxide radical abstracts deuterium and hydrogen via unimolecular and
bimolecular reactions
Most of the kinetic data is known for ethyl nitrite but for simplicity sake we have
conducted our experiments primarily using methyl nitrite. From the above observed reactions
and the proposed role played by the hydrogen radical, it is possible to think that H-atom
chemistry and possibly other radicals could be of importance in the pyrolysis of biomass (plantderived materials, e.g. cellulose). In the experiments described in this section we studied the
reaction of NO radical with acetaldehyde. HNO molecule was detected, confirming the hydrogen
abstraction reaction by NO radical in a hot pyrolysis reactor. We also qualitatively compared the
rate of that bimolecular reaction with the rate of unimolecular reaction in methyl nitrite.
In order to distinguish between the bimolecular and unimolecular reactions, we
synthesized fully deuterated methyl nitrite, CD 30NO. The unimolecular pyrolysis of CD 30NO
can result in DNO product, which could be confirmation of the below reactions scheme:
44
CD 3 -O-N=0 - CD 20-+DNO
DNO has rotational transitions (from the vibrational ground state) at 73084.5745 and
73085.4600 MHz. The chirp is made to cover 73084 - 73086 MHz, or when backtracking from
the up-conversion (divide by 8 and subtract 10.7 GHz): from 1564.25 - 1564.50 MHz. Therefore
we were looking for lines on the oscilloscope (after down-conversion: 84.385 GHz (9th
harmonic of the Gunn oscillator) - transition frequency) at 11.299 and 11.300 GHz. Fig. 4.2
shows the observed transition line that appears to be two unresolved peaks at the abovementioned frequencies.
45
12
>
10
C
8
6
E 4
E
02
11.296
11.298
11.300
11.302
Transition frequency, GHz
11.304
2012-00-00_000
Figure 4.2 CPmmW spectrum of DNO upon pyrolysis of the mixture 3 %acetaldehyde and 5.5% deuterated methyl nitrite. Spectrum
was taken at long FID duration for higher resolution (2.0 ps/division). The peaks did not completely resolve, but it appears that both
transitions at 11.299 and 11.300 GHz are present in the spectrum. Absolute stagnation pressure in the pyrolysis nozzle was 3.5 bars
and pyrolysis temperature was 800 *C.
46
The bimolecular reaction of NO radical with H atoms of acetaldehyde would result in
HNO molecules in the chamber. We therefore searched for HNO under the same conditions.
HNO has a ground state transition with quadrupole hyperfine splitting at 81477.49 MHz, and
referring to the above up-conversion and down-conversion discussion, we made a single-
frequency chirp at 515.31 MHz (81.47749 GHz/8 - 10.7 GHz) and looked for the transition on
our oscilloscope at 2.9075 GHz (84.385 GHz - 81.47749 GHz). Figure 4.3 shows the CPmmW
spectrum of HNO.
47
7
>6
~u 5
E 3
E
-
0
2.86
2.88
2.92
2.90
Transition frequency, GHz
2.94
2.96
2012-00-00_000
Figure 4.3 CPmmW spectrum of HNO at short FID or lower resolution (400 ns/division). The spectrum is only indicative of H atoms
that were already present in the chamber, but still we clearly see a peak at 2.908 GHz that corresponds to the 3 degenerate groundstate rotational transitions at 81477.49 GHz. Absolute stagnation pressure in the pyrolysis nozzle was 3.5 bars and temperature was
800 0 C.
48
Comparing the DNO and HNO signals in Figs. 4.2 and 4.3, respectively, we can conclude
that the unimolecular decomposition of methyl nitrite with DNO products and bimolecular
hydrogen abstraction reaction of NO with acetaldehyde resulting in HNO product will result in a
similar yield of DNO and HNO. More quantitative studies and modeling will be required to
obtain the rates of the reactions.
In general, to thermally crack molecules such as H 2, acetaldehyde, or methanol, just to
name a few examples, very high temperatures and pressures are required and results in a mixture
of products containing a high proportion of alkene products [32]. Modem fluid catalytic cracking
uses a solid catalyst [33]. These cracking experiments by nitric oxide radical demonstrate that
-NO can be used as a gas-phase catalyst to perform the cracking at much lower temperatures and
with higher efficiency than thermal cracking.
Concluding this section, the key features of CPmmW spectroscopy, which are the ability
to detect a broad spectral range simultaneously, high sensitivity and meaningful relative
intensities, were used to detect multiple pyrolysis products. In this work we have demonstrated
the use of CPmmW spectroscopy for studying unimolecular and bimolecular reactions.
49
Appendix: Parts list of CPmmW spectrometer
i.
10 MHz Rubidium frequency standard (Stanford Research Systems FS725)
ii.
iii.
90 MHz phase-locked crystal oscillator (Miteq PLD-10-90-15P)
8.75 GHz DLCRO phase-locked oscillator (Miteq DLCRO-010-08750-3-15P)
iv.
8 GS/s arbitrary waveform generator (Tektronix AWG7082C)
9.375 GHz DLCRO (dual loop phase-locked coaxial resonator oscillator) (Miteq
v.
DLCRO-010-09375-3--15P)
vi.
Triple-balanced mixer (Macom MY51C)
vii.
8-16 GHz circulator (Hitachi R3113110)
viii.
ix.
6-18 GHz Amplifier (ALC Microwave ALSO3-0283, 20 dB gain)
a. 6.875-9.125 GHz bandpass filter (Spectrum Microwave 311-307389-001)
b. 10.9 - 12.7 GHz bandpass filter (Spectrum Microwave C 11800-1951-1355)
x.
Active frequency quadrupler (Phase One SX50-416)
xi.
Active frequency doubler (Millitech AMC-12-110546)
xii.
Direct Reading Precision Attenuator (Cernex CDA 12609009)
xiii.
WR1O gain horn, 24 dBi gain mid-band (TRG 861W/387)
xiv.
xv.
xvi.
xviii.
Teflon circular lenses, 30 cm and 40 cm focal length at 80 GHz (custom)
Gunn phase lock module (LX Microwave model 800A)
W-Band Gunn oscillator (J. E. Carlstrom Co.)
W-Band subharmonic downconverter (Pacific Millimeter Products)
E-Band balanced mixer (Ducommun Technologies FDB-12-01
xix.
Low-noise amplifier (Miteq AMF-7D00101800-24-1OP, 55 dB gain)
xx.
8 GHz fast digital storage oscilloscope (Tektronix DP070804C, 25 GS/s)
xvii.
50
References
[1]
C. H. Townes and A. L. Schawlow, Microwave Spectroscopy. Dover: New York. (1975).
[2]
D. L. Bryce and R. E. Wasylishen. Microwave spectroscopy and nuclear magnetic
resonance spectroscopy - what is the connection? Accounts of Chemical Research, 36,
327 (2003).
[3]
T. J. Balle and W. H. Flygare. Fabry-Perot Cavity Pulsed Fourier-Transform Microwave
Spectrometer with a Pulsed Nozzle Particle Source. Review of Scientific Instruments, 52,
1, 33 (1981).
[4]
J. U. Grabow, W. Stahl, and H. Dreizler. A multioctave coaxially oriented beamresonator arrangement Fourier-transform microwave spectrometer. Review of Scientific
Instruments, 67, 4072 (1996).
[5]
E. Arunan, S. Dev, and P. K. Mandal. Pulsed nozzle Fourier-transform microwave
spectrometer: advances and applications. Applied Spectroscopy Reviews. 39, 131
(2004).
[6]
[7]
K. B. McAfee, Jr., R. H. Hughes, and E. B. Wilson, Jr. A Stark-Effect Microwave
Spectrograph of High Sensitivity. St Review of Scientific Instruments, 20, 821 (1949).
G. G. Brown, B. C. Dian, K. 0. Douglass, S. M. Geyer, S. T. Shipman, and B. H. Pate. A
broadband Fourier transform microwave spectrometer based on chirped pulse excitation.
Review of Scientific Instruments, 79, 5, 053103 (2008).
[8]
D. N. Newmark. Transition State Spectroscopy. Science, 272, 5267, 1446 (1996).
[9]
G. L. Barnes and W. L. Hase. Transition state analysis bent out of shape. Nature
Chemistry, 1, 2, 103 (2009).
[10]
R. D. Suenram, J. U. Grabow, A. Zuban and I. Leonov. A portable, pulsed-molecularbeam, Fourier-transform microwave spectrometer designed for chemical analysis. Review
of Scientific Instruments, 70, 2127 (1999).
[11]
J. M. Brown and A. Carrington. RotationalSpectroscopy of Diatomic Molecules,
Cambridge Molecular Science Series. Cambridge University Press: Cambridge, New
England (2003).
[12]
F. C. De Lucia. The submillimeter: A spectroscopist's view. Journalof Molecular
Spectroscopy, 261, 1, 1 (2010).
[13]
W. Gordy and R. L. Cook. Microwave MolecularSpectra. Wiley: New York (1984).
51
[14]
G. B. Park, A. H. Steeves, K. Kuyanov-Prozument, J. L. Neill, and R. W. Field. Design
and evaluation of a pulsed-jet chirped-pulse millimeter-wave spectrometer for the 70-102
GHz region. The Journalof Chemical Physics, 135, 024202 (2011).
[15]
A. Derecskei-Kovacs and S. W. North. The unimolecular dissociation of vinylcyanide: A
theoretical investigation of a complex multichannel reaction. Journalof Chemical
Physics, 110, 2862-2871 (1999).
[16]
Z. Homayoon, S. A. Vazquez, R. Rodriquez-Fernandez, and E. Martinez-Nunez. Ab
initio and RRKM study of the HCN/HNC elimination channels from vinyl cyanide.
Journalof Physcial ChemistryA, 115, 979-985 (2011).
[17]
H. A. Betchel, A. H. Steeves, B. M. Wong, and R. W. Field. Evolution of chemical
bonding during HCN<-+HNC isomerization as revealed through nuclear quadrupole
hyperfine structure. Angewandte Chemie InternationalEdition, 47, 2969 (2008).
[18]
J. W. Wilhelm, M. Nikow, L. Letendre, and H.-L. Dai. Photodissociation of vinyl
cyanide at 193 nm: Nascent product distributions of the molecular elimination channels.
Journalof ChemicalPhysics, 130, 44307 (2009).
[19]
B. R. Heazlewood, A. T. Maccarone, D. U. Andrews, D. L. Osborn, L. B. Harding, S. J.
Klippenstein, M. J. T. Jordan, and S. H. Kable. Near-threshold H/D exchange in
CD 3CHO photodissociation. Nature Chemistry, 3, 443 (2011).
[20]
R. H. Qadiri, E. J. Feltham, N. H. Nahler, R. P. Garcia, and M. N. R. Ashfold. Propyne
and allene photolysis at 193.3 nm and at 126.6 nm. Journalof Chemical Physics, 119, 24,
12842 (2003).
[21]
Y. Ganot, S. Rosenwaks, and I. Bar. H and D release in -243.1 nm photolysis of
vibrationally excited 3vi, 4vi, and 4 vCD overtones of propyne-d3. Journalof Chemical
Physics, 120, 18, 8600 (2004).
[22]
D. Mohan, C. U. Pittman, Jr., and P. H. Steele. Pyrolysis of wood/biomass for bio-oil: A
critical review. Energy Fuels, 20, 3, 848 (2006).
[23]
A. Vasiliou, M. R. Nimlos, J. W. Daily, and G. B. Ellison. Thermal decomposition of
furan generates propargyl radicals. Journalof Physical Chemistry A, 113, 30, 8540
(2009).
[24]
A. Vasiliou, K. M. Piech, X. Zhang, M. R. Nimlos, M. Ahmed, A. Golan, 0. Kostko, D.
L. Osborn, J. W. Daily, J. F. Stanton, and G. B. Ellison. The products of the thermal
decomposition of CH 3CHO. Journal of Chemical Physics, 135, 1, 014306 (2011).
52
[27]
P. Chen, S. D. Colson, W. A. Chupka, and J. A. Berson. Flash pyrolytic production of
rotationally cold free radicals in a supersonic jet: resonant multiphoton spectrum of the
3p 2 a2 +- x 2 a2 origin band of methyl. Journalof Physical Chemistry, 90, 11, 2319 (1986).
[28]
D. W. Kohn, H. Clauberg, P. Chen. Flash pyrolysis nozzle for generation of radicals in a
supersonic jet expansion. Review of Scientific Instruments, 63, 4003 (1992).
[29]
C. M. Lovejoy and D. J. Nesbitt. Slit pulsed valve for generation of long-path-length
supersonic expansions. Review of Scientific Instruments, 58, 807 (1987).
[30]
C. M. Lovejoy. Intramoleculardynamics of Van der Waals complexes as elucidatedby
infraredspectroscopy in slit supersonic expansions. Thesis, University of Colorado at
Boulder. Ann Arbor: ProQuest/UMI, 1990. (Publication No. AAT 9032853.)
[31]
M. D. Morse. "Supersonic beam sources," in Methods of ExperimentalPhysics: Atomic,
Molecular, and Optical Physics, Vol II. Atoms and Molecules, edited by F. B. Dunning
and R. Hulet, Academic Press, Inc: Orlando, Florida, 21-47 (1996).
[32]
R. E. Smalley, D. H. Levy, and L. Wharton. The fluorescence excitation spectrum of the
HeI 2 van der Waals complex. Journalof Chemical Physics, 64, 3266 (1976).
[33]
B. M. Wong. Nuclear quadrupole hyperfine structure in HC' 4N/H 14NC and
DC15 N/D1 5 NC isomerization: a diagnostic tool for characterizing vibrational localization.
Physical Chemistry Chemical Physics, 10, 5599 (2008).
[34]
B. S. Greensfelder, H. H. Voge, and G. M. Good. Catalytic and thermal cracking of pure
hydrocarbons: mechanisms of reactions. Industrialand Engineering Chemistry, 41, 11, 2573
(1949).
[35]
Petrolum refining. Encyclopedia Britannica,Encyclopedia BritannicaOnline Academic
Edition. Encyclopedia Britannica Inc. (2013). Web. 16 Jan. 2013.
http://www.britannica.com/EBchecked/topic/454440/etroleum-refinin.
53
Download