2011-DMA jp206762j

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

ARTICLE

pubs.acs.org/JPCA

Fourier Transform Infrared Spectroscopy and Theoretical Study of


Dimethylamine Dimer in the Gas Phase
Lin Du and Henrik G. Kjaergaard*
Department of Chemistry, University of Copenhagen, Universitetsparken 5, DK-2100 Copenhagen Ø, Denmark
bS Supporting Information
ABSTRACT: Dimethylamine (DMA) has been studied by gas-phase Fourier
transform infrared (FTIR) spectroscopy. We have identified a spectral
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

transition that is assigned to the DMA dimer. The IR spectra of the dimer
in the gas phase are obtained by spectral subtraction of spectra recorded at
different pressures. The enthalpy of hydrogen bond formation was obtained
Downloaded via ACADEMIA SINICA on September 10, 2020 at 13:54:40 (UTC).

for the DMA dimer by temperature-dependence measurements. We comple-


ment the experimental results with ab initio and anharmonic local mode model
calculations of monomer and dimer. Compared to the monomer, our calcula-
tions show that in the dimer the NH bond is elongated, and the NH-
stretching fundamental shifts to a lower wavenumber. More importantly, the
weak NH-stretching fundamental transition has a pronounced intensity
increase upon complexation. However, the first NH-stretching overtone
transition is not favored by the same intensity enhancement, and we do not
observe the first NH-stretching overtone of the dimer. On the basis of the measured and calculated intensity of the NH-stretching
transition of the dimer, the equilibrium constant for dimerization at room temperature was determined.

’ INTRODUCTION overtone range of the NH-stretching vibration as a function


Dimethylamine [(CH3)2NH, DMA] has been studied both of concentration and temperature in different solutions (CCl4,
experimentally and theoretically. Fang et al. reported the gas- n-hexane) and identified a characteristic band from DMA
phase overtone absorption spectra of DMA by intracavity laser associates.19 The self-association of DMA was also examined in
photoacoustic and FTIR spectroscopy in an attempt to detect Bernard-Houplan and Sandorfy’s low temperature infrared study
aggregates.1 They measured the temperature and pressure de- of dissolved DMA.20 The solvents were 1:1 mixtures of CCl3F
pendence of the NH-stretching overtone intensities of DMA. No and CF2Br-CF2Br and of CCl3F and methylcyclohexane, which
relative intensity changes of the various NH-stretching overtone can form rigid glasses near liquid nitrogen temperature. Odutola
bands were observed when the vapor pressure of DMA changed et al. performed an electric deflection study of (DMA)2 and
from 3 to 100 Torr or when the temperature was increased from found that the dipole moment of (DMA)2 is greater than 0.3 D.21
0 to 80 C. On the basis of these results, they concluded that It is more favorable for the dimer to exhibit a linear hydrogen-
formation of aggregates is unlikely in their experiments. In 2007, bonded conformation, compared with a cyclic structure with two
Marinov et al. reported high-resolution absorption spectrum for hydrogen bonds between N and H atoms, which would have a
gas-phase DMA in the region of the first NH-stretching overtone dipole moment of only a few tenths of a Debye. Bohn and
by cavity-ringdown spectroscopy.2 They measured the spectra Andrews identified bands of (DMA)2 in their measured FTIR
at room temperature and at pressures in the range from 29.4 spectra of NH3DMA complexes in solid argon.22 On the basis
to 204.4 mbar. Very recently, Miller et al. measured absolute of the spectrum, they proposed that the two submolecules are
intensities of the fundamental to third NH-stretching overtone inequivalent in (DMA)2 like in (NH3)2.13 Several years later,
transitions in DMA.3 Tubergen and Kuczkowski measured the rotational spectra for
Several studies have been aimed at understanding the struc- six isotopmers of (DMA)2 using a Fourier-transform microwave
ture and bonding of molecular complexes involving NH3 and spectrometer and determined the most stable structure of the
methylamines.417 However, the DMA dimer [(DMA)2] has DMA dimer.23 In addition, the ultraviolet photoelectron spec-
been the subject of only a few studies. Early in 1950, Lambert trum of (DMA)2 was obtained in the vapor phase by spectrum
and Strong studied the dimerization of ammonia and amines stripping and assigned with the help of ab initio molecular orbital
and pointed out that the dimerization is attributed to hydrogen (MO) calculations.24 The calculated orbital energies based on a
bonding.18 The energy of the NH 3 3 3 N bond in the DMA
dimer was determined to be 10.5 kJ mol1 by measuring the Received: July 15, 2011
compressibility of the vapor. Wolff and Gamer investigated Revised: September 23, 2011
the infrared spectra of DMA in the fundamental and in the first Published: September 27, 2011

r 2011 American Chemical Society 12097 dx.doi.org/10.1021/jp206762j | J. Phys. Chem. A 2011, 115, 12097–12104
The Journal of Physical Chemistry A ARTICLE

nearly linear hydrogen bonded dimer structure show reasonably step size = 1  105 au., and energy = 1  107 au., and the global
good agreement with the experimental values. Recently, thresholds for single-point calculations were set to energy = 1 
Cabaleiro-Lago and Ríos studied the interaction in the DMA 108 au. The harmonic frequencies were calculated with the
dimer and trimer with the HF, DFT/B3LYP, and MP2 ab initio B3LYP/aug-cc-pVTZ and QCISD/aug-cc-pVDZ methods with
methods in conjunction with the 6-31+G* and aug-cc-pVDZ/ MOLPRO. The integration grid-size used for the B3LYP density
cc-pVDZ (i.e., the aug-cc-pVDZ set for carbon and nitrogen functional calculations was set with an overall target accuracy of
atoms and the cc-pVDZ set for hydrogen atoms) basis sets.25 1  108 au. Unless otherwise stated, QCISD calculations were
They found the most stable conformer for DMA dimer to have Cs performed with MOLPRO and B3LYP calculations were carried
symmetry and a close to linear hydrogen bond. out with Gaussian 03. We have used the Boys and Bernardi
Vibrational spectroscopy is extensively applied to the study of counterpoise correction procedure (CP) to reduce the basis
hydrogen bonded complexes; however, measurements in solu- set superposition error (BSSE) in the dimer calculations.34 We
tion and in matrices are affected by solvent effects. Gas-phase apply the CP correction only to the energy of the non-CP
spectroscopy can be used to quantify the complexes and the optimized geometry. The CP-correction method has been shown
equilibrium constant of their formation. On the basis of previous to improve interaction energies of weakly bound complexes with
experience with FTIR and quantum chemical calculations of some methods.810,3537
weakly bound methanol clusters in the gas phase,26,27 we have Furthermore, the behavior of the NH-stretching vibration is a
recorded the FTIR spectra of the DMA dimer and supported very sensitive probe for the local environment. An anharmonic
these experiments with ab initio calculations of the monomer and oscillator local mode model was used to calculate the NH-
dimer. The NH-stretching wavenumbers and oscillator strengths stretching frequencies and intensities.30,31 Previously, the NH-
for DMA and its dimer were calculated with an anharmonic stretching modes in a range of molecules have been shown to
oscillator local mode model.2831 Furthermore, by measuring be isolated from other vibrational modes, and hence are well
the dimer spectra over a range of temperatures, the dimerization described by the local mode model of vibration.3840 We assume
enthalpy ΔH was obtained and compared with the theoretical that the NH-stretching vibrations can be described by a Morse
results. The equilibrium constant of the DMA dimerization oscillator, with the vibrational energy levels given by
reaction was also estimated based on experimental and calculated    
NH-stretching intensities. 1 1 2
EðvÞ=ðhcÞ ¼ v þ ω
~ v þ ω
~x ð1Þ
2 2
’ EXPERIMENTAL AND THEORETICAL METHODS The Morse oscillator frequency ω~ and anharmonicity ω ~x are
The IR spectra were recorded at 1.0 cm1 resolution with a found from the second, third, and fourth-order derivatives of the
VERTEX 70 (Bruker) FTIR spectrometer fitted with a CaF2 potential energy curve.31 These derivatives are found by fitting an
beam splitter and an MCT detector. During the experiments, the eighth-order polynomial to a symmetric 9-point ab initio calcu-
FTIR spectrometer was purged with dry nitrogen gas to mini- lated potential-energy curve, obtained by displacing the NH
mize the interference by water and CO2. Several gas cells with bond from 0.20 to 0.20 Å in 0.05 Å steps around equilibrium
different path lengths were used, including a 10-cm cell equipped bond length. This range and step size of the potential-energy
with KBr windows and a 2.4 m and a 4.8 m path length curve ensure converged energy derivatives.41 The dimensionless
multireflection gas cell (Infrared Analysis, Inc.) fitted with Infrasil oscillator strength f of a transition from the ground vibrational
quartz windows. The 2.4 m cell was equipped with a heating state 0 to an excited vibrational state v is given by42
jacket, and the temperature was controlled by a Digi-Sense
fv, 0 ¼ 4:702  107 ½cmD2 ~vv, 0 jμ
~v, 0 j2 ð2Þ
electronic temperature controller (Eutech Instruments Pte
1
Ltd., Model 6890003). The gas samples were prepared on a where ~vv,0 is the transition frequency in cm and μBv,0 = Æv|μB|0æ
glass vacuum line (base pressure less than 1  104 Torr) is the transition dipole moment in Debye (D). The transition
equipped with a Varian cold cathode vacuum gauge (1  107 to dipole moment matrix element can be expanded as a Taylor
1  102 Torr, Model 860A). Sample pressures were measured series in the NH-stretching displacement coordinate q, and we
with a Varian diaphragm pressure gauge (11500 Torr, DV100). limit the expansion to fifth-order terms. The dipole moment
In the variable temperature experiment, the DMA spectra were coefficients are found by fitting an eighth-order polynomial to a
measured with the 2.4 m cell in the temperature range from 9-point dipole moment curve calculated at the same points as the
296 to 368 K. The pressure was measured at room temperature, potential. The integrals Æv|qn|0æ needed for the transition dipole
and the ideal gas law used to calculate the pressure at the elevated moment were evaluated numerically.
temperature. Before each measurement, we waited for at least
20 min to let the temperature of the gas in the cell stabilize. The ’ RESULTS AND DISCUSSION
adsorption of DMA to the cell walls was observed, however the
rate of adsorption was slow and did not affect our measurement Geometries. The optimized geometries of DMA and two
significantly. DMA (anhydrous, 99+%) was purchased from different DMA dimer conformers are shown in Figure 1. DMA
Aldrich and used without any further purification. has Cs symmetry, and the QCISD/aug-cc-pVTZ calculated
Gaussian 03 (revision E.01)32 and MOLPRO (version NH bond length is 1.011 Å, which agrees well with the 1.022 Å
2009.1)33 were used to perform the calculations. The geometries bond length derived from the microwave spectrum.43
of DMA and (DMA)2 were optimized using B3LYP hybrid Compared with the monomer, much less structural informa-
density functional theory and QCISD ab initio theory with the tion is available on the dimer. Motivated by the potential
aug-cc-pVDZ and aug-cc-pVTZ basis sets, and using CCSD theory similarity between the DMA dimer and the ammonia dimer,
with the cc-pVDZ basis set. The optimization threshold criteria Tubergen and Kuczkowski obtained sufficient experimental
of the MOLPRO calculations were set to: gradient = 1  105 au., information to rule out many possible dimer conformers and
12098 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104
The Journal of Physical Chemistry A ARTICLE

concluded that the two most likely structures have nonlinear changes of DMA upon dimerization is very similar for confor-
hydrogen bonds.23 With the help of ab initio calculations and a mers A and B. The QCISD/aug-cc-pVDZ energy of conformer B
distributed multipole analysis to estimate the electrostatic en- is 1.6 kJ mol1 lower than that of conformer A, and for simplicity
ergy, they concluded that one particular structure is most only conformer B was optimized at higher levels and used in our
likely. Wales et al. studied the potential energy surface of the discussion. We also calculated the NH-stretching frequencies
DMA dimer by using a semiempirical potential function and and intensities for conformer A with a few methods and found
identified two minima on the potential surface.17 One of them the results to be close to those of conformer B. The comparison
corresponded to the structure proposed by Tubergen and between conformer A and B results is given in the Supporting
Kuczkowski;23 however, the other minimum that previously Information.
had not been reported agreed very well with experimental values. As shown in Figure 1, conformer B is a hydrogen-bonded
Mayer et al. performed an ab initio study on the DMA dimer at complex of Cs symmetry. To discuss more clearly, geometric
the MP2/6-31G* level and found only one minimum, which was parameters are defined: RNH is the NH bond length in DMA,
similar to the structure derived from microwave spectrum.44 and R(NHb) and R(NHf) are the bonded and free NH bond
They claimed that no other stable orientation of the two length in the dimer, respectively, d is the length of the hydrogen
monomers could be found on the MP2/6-31G* surface. To bond (the H 3 3 3 N distance), ψ is the angle between the CNC
eliminate these discrepancies, Cabaleiro-Lago and Ríos studied plane of the acceptor DMA molecule and the donor H, and θ is
the DMA dimer using DFT and MP2 methods in conjunction the NH 3 3 3 N angle or H-bond angle. The R(NHb), R(NHf), and
with different basis sets.25 They identified minima for three change in NH bond length upon complexation (Δr), together
different dimer conformers, which they called 2A, 2B, and 2C. with the d, ψ, and θ angels obtained at different computational
Among them, structure 2B is a minimum on the potential surface levels, are listed in Table 1. The z-matrix of DMA and (DMA)2
with all the methods and basis sets used and has the lowest energy at the QCISD/aug-cc-pVTZ level are given in Supporting
compared with the other two structures. The properties of Information.
structure 2C depend largely on the particular method used Generally, due to the lack of dispersion forces, DFT methods
and introducing dispersion with the MP2 method substantially predict complexes too loosely bound. As seen in Table 1, the
alters the structure of this minimum. Structure 2A is similar to 2B B3LYP method predicts larger d value compared to the other
except that the H-donor DMA molecule is rotated about the methods. At the QCISD/aug-cc-pVTZ level, the change in the
N 3 3 3 N axis.25 The region of the potential surface near structures NH bond length upon complexation is 0.0036 Å, while this
2A and 2B is very flat, and the molecules can rotate relatively change was reported to be 0.005 Å in the previous MP2 study.25
freely about the N 3 3 3 N axis. They concluded that the most The geometry change with dimerization is small, which suggests
stable minimum corresponds to structure 2B, which is similar to that the hydrogen bond in the dimer is weak. The NHf
that derived from the microwave spectrum for the dimer.25 bond length of the dimer is roughly the same as in the isolated
On the basis of previous studies, we include only the two molecule. The length of the hydrogen bond (the H 3 3 3 N
conformers A and B (Figure 1), which correspond to structure distance) is calculated to be 2.1810 Å (QCISD/aug-cc-pVTZ),
2A and 2B in ref 25. After preliminary examination, the geometry which is shorter than the hydrogen bond in the ammonia dimer
(2.38 Å) as expected, but it is close to the hydrogen bond length
in the CHCl3NH3 complex (2.25 Å).9 The angle between the
CNC plane of the acceptor and Hb is calculated to be 99 at the
QCISD/aug-cc-pVTZ level. This is smaller than the comparable
angle in DME-MeOH but larger than in DMS-MeOH.27 In an
ideal hydrogen bond, the NH 3 3 3 N angle θ is close to 180.45
In DMA dimer there is a 25 deviation from linearity of the
hydrogen bond, which again suggests that the hydrogen bond is
not strong in this complex. As in the ammonia dimer, this
deviation can be rationalized by electrostatic interactions.9
Besides the primary hydrogen bond, there is a secondary inter-
action between the positive region on the NH bond in the
hydrogen bond acceptor DMA with the negative region around
Figure 1. Optimized structures of DMA and two conformers of the partially negatively charged N of the donor which makes the
its dimer. deviation from linearity energetically favorable.9

Table 1. Selected Optimized Geometric Parameters (Ångstroms and Degrees) in the DMA Dimer
conformer method R(NHb) R(NHf) Δra d ψ θ

A B3LYP/aug-cc-pVTZ 1.0160 1.0120 0.0048 2.2912 133.1 171.5


CCSD/cc-pVDZ 1.0256 1.0240 0.0023 2.2059 143.6 152.9
QCISD/aug-cc-pVDZ 1.0235 1.0199 0.0033 2.1888 111.6 165.4
B B3LYP/aug-cc-pVTZ 1.0161 1.0117 0.0049 2.2698 114.1 167.5
CCSD/cc-pVDZ 1.0256 1.0231 0.0023 2.1932 100.3 147.8
QCISD/aug-cc-pVDZ 1.0234 1.0198 0.0031 2.1655 99.6 155.0
QCISD/aug-cc-pVTZ 1.0149 1.0115 0.0036 2.1810 99.3 154.6
a
Change in the NH bond length upon complexation, R(NHb)  RNH.

12099 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104


The Journal of Physical Chemistry A ARTICLE

Table 2. Calculated BEs of the DMA Dimer (Conformer B)


method BE/kJ mol1 BE (CP corrected)/ kJ mol1

B3LYP/aug-cc-pVTZ a 7.65 7.35


CCSD/cc-pVDZ a 23.60 7.49
QCISD/aug-cc-pVDZ a 21.41 13.37
QCISD/aug-cc-pVTZ a 18.34 14.93
a
BE includes zero-point vibrational energy (ZPVE) correction (ZPVEtot =
ZPVEdimer  2ZPVEmonomer), obtained from unscaled B3LYP/aug-cc-
pVTZ harmonic frequencies and amount to 2.79 kJ mol1.

Figure 3. The DMA dimer spectra in the fundamental NH-stretching


region as a result of spectral subtraction of DMA spectrum from the
spectra measured at different vapor pressures. The spectra were mea-
sured with a 10 cm cell at 296 K. The higher pressure spectra are offset to
avoid overlapping between the spectra. The absorbance is obtained by
log(I0/I).

increases with pressure. To confirm that the difference between


the two spectra arises from DMA dimer, a series of spectra at
different pressures were measured.
We measured DMA spectra at pressures of 2, 70, 204, 315,
445, and 700 Torr. The spectrum at 2 Torr was measured with a
Figure 2. DMA spectra in the fundamental NH-stretching region path length of 4.8 m and the other spectra with a path length of
measured at 700 and 2 Torr vapor pressures with a path length of
10 cm and 4.8 m, respectively, at 296 K. The left-hand side ordinate
10 cm. The spectrum at 2 Torr is subtracted with appropriate
corresponds to the spectrum at 700 Torr. The absorbance is obtained weights taking into account the pressure and path length
by log(I0/I). difference of the FTIR spectra recorded at the five higher
pressures. All spectra were measured at 1.0 cm1 resolution.
We also measured a few spectra at 0.2 cm1 resolution. These
The binding energy (BE) defined as the energy difference
higher resolution spectra were almost identical to the spectra
between the dimer (ED) and the two monomers (EM), i.e., BE =
recorded with 1.0 cm1 resolution, and no fine structure was
ED  2EM, was calculated at various levels and are given in
observed at higher resolution. We assign the residual to be the
Table 2. Zero-point vibrational energy corrections (ZPVE) were
fundamental NH-stretching band of DMA dimer located at
included in the BE and obtained from the unscaled B3LYP/aug-
3339 cm1. This position is found by fitting a Lorentzian
cc-pVTZ harmonic frequencies. The B3LYP/aug-cc-pVTZ har-
function to the observed band. In the region from 2400 to
monic frequencies are given in Table S1 of the Supporting
7000 cm1, we did not observe any other transitions that we
Information. The binding energy of the QCISD/aug-cc-pVTZ
could assign to the DMA dimer. The spectra obtained after
optimized dimer is calculated to be 18.3 kJ mol1, which the
subtraction are shown in Figure 3.
CP correction on the energy changes to 14.9 kJ mol1. The
According to the chemical equilibrium for DMA dimerization,
magnitude of the CP correction decreases with increase in basis
set size, as expected. Very recently, optimization of the DMA 2DMA / ðDMAÞ2 ð3Þ
dimer with the explicitly correlated coupled cluster method
CCSD(T)-F12/VDZ-F12 found the binding energy of confor- Kp ¼ pD =pM 2 ð4Þ
mer B to be 18.6 kJ mol1.46 The CCSD(T)-F12/VDZ-F12
method have been found to give binding energies for small where Kp is the equilibrium constant and pD and pM are the vapor
complexes that are in good agreement with CCSD(T) results at pressures of the dimer and monomer, respectively. At a given
the complete basis set limit.10 temperature, Kp is a constant, and therefore, pD is proportional to
FTIR Spectra of the DMA Dimer. Fang et al. measured the pM2. In the IR absorption spectrum, the integrated intensity of
temperature and pressure dependence of the NH-stretching the DMA dimer band is proportional to pD, providing the same
overtone intensities of DMA and no relative intensity changes path length is used. We integrated the band areas shown in
of the various NH-stretching overtone bands were observed Figure 3, and plot the band area vs pM2 in Figure 4. The results fit
with a DMA vapor pressure range from 3 to 100 Torr and a well with a linear fit (Figure 4), which suggests that the complex
temperature from 0 to 80 C.1 However, when we focused on the formed at higher pressures is (DMA)2 and not large aggregates
fundamental NH-stretching region and changed the vapor and also support our assignment of the measured band to the
pressure from 2 to 700 Torr, a new feature appears in the spectra DMA dimer.
to the low energy side of the DMA NH-stretching band We have calculated IR frequencies and intensities of DMA and
(Figure 2). The relative intensity of the low energy shoulder DMA dimer both with a harmonic oscillator linear dipole (HO)
12100 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104
The Journal of Physical Chemistry A ARTICLE

We believe that a VTZ quality basis set is necessary to describe


the dimer well.
In the spectrum of DMA, two transitions at 3374 and
3359 cm1 are observed in the NH-stretching region, due to
splitting by tunneling in the inversion mode. The higher fre-
quency transition arise from the ground state and is observed to
be slightly more intense.3 In our measurements, the DMA dimer
transition observed at 3339 cm1 leads to an observed gas-phase
frequency redshift of 35 cm1. Our QCISD/aug-cc-pVTZ AO
calculated red shift of the NHb-stretching vibration is 61 cm1.
However, it has been observed that the coupling to lower
frequency modes ignored in the AO model can change the
stretching frequencies.48 If we compare the HO and AO calcu-
lated redshift with the B3LYP/aug-cc-pVTZ method we find that
the shift is about 20 cm1 larger with the AO model. A 20 cm1
decrease of the QCISD/aug-cc-pVTZ calculated shift would
bring it in very good agreement with the observed shift.
In the Ar matrix experiment, the NH-stretching fundamental
Figure 4. The linear fitting between integrated area of the DMA dimer transition in DMA is assigned to a peak at 3377 cm1, and only
absorption band and pM2. These integrated areas are obtained from one transition is assigned. However, this transition is higher than
room temperature spectra at 296 K. both transitions observed in the gas phase, and with the usual
redshift caused by the Ar matrix we believe that this assignment is
perhaps incorrect. It should also be noted that the NH-stretching
Table 3. Observed and Calculated NH-Stretching Funda- transition in DMA is about 100 times weaker than normal
mental Frequency (cm1) and Relative Intensities for the NH-stretching transitions, and perhaps difficult to assign in the
DMA Dimer matrix experiment.3 Thus, the reported redshift of 73 cm1
from the solid Argon matrix FTIR experiment is likely too
method υNH Δυa fD/fMb
high.22
harmonic B3LYP/aug-cc-pVTZ 3445 82 363 One should also take into consideration that we measure at
QCISD/aug-cc-pVDZ 3477 45 952 room temperature but calculate at 0 K. Previously, a difference
anharmonic B3LYP/aug-cc-pVTZ 3277 103 835 between room temperature redshift and low temperature
QCISD/aug-cc-pVDZ 3313 57 144
redshift has been found for the CHCl3NH3 complex.9 The
redshift of the CH-stretching vibration in the jet-cooled spec-
QCISD/aug-cc-pVTZ 3348 61 571
trum (38 cm1) is somewhat larger than the observed redshift at
Bohn and Andrewsc
observed 3304,3313 73
room temperature (17.5 cm1). The reason for this discrepancy
this work 3339 35 is a temperature effect. At room temperature, the low-frequency
a
Δv = vmonomer  vdimer b fD/fM is the oscillator strength of dimer vs intermolecular vibrational modes are expected to be populated,
monomer. c Taken from ref 22. which lead to a broadening of the band and an overall redshift of
the band maximum.9
approximation and with an anharmonic oscillator local mode The fundamental NH-stretching transition in DMA is very
(AO) model. The calculated DMA dimer NH-stretching funda- weak is due to a very small dipole moment derivative.3 Thus,
mental frequency and intensity relative to that of the monomer deviations from a harmonic oscillator and a linear dipole become
are listed in Table 3. Additional results of the calculated important, and intensity ratio between dimer and monomer
frequencies and absolute intensities are given in Tables S1S3 NH-stretching transitions (fD/fM) is likely more reliable with the
of Supporting Information. The calculations at all levels predict AO model. We believe that fD/fM ≈ 600 is a reasonable value for
an H-bonded dimer with a characteristic redshift and a very the intensity enhancement based on our QCISD/aug-cc-pVTZ
pronounced intensity increase of the NHb-stretching vibration. calculations. In comparison, the recent CCSD(T)-F12/VDZ-
The measured peak of the dimer band at 3339 cm1 is in F12 value for fD/fM is 611.46 This is significantly higher than the
reasonable agreement with our QCISD/aug-cc-pVTZ anharmo- predicted values for trimethylamine-methanol (TMA-MeOH),
nic oscillator (AO) calculated NHb-stretching vibration fre- dimethylsulfide-methanol (DMS-MeOH), and dimethylether-
quency of 3348 cm1. It is also in good agreement with peaks methanol (DME-MeOH) dimers,26,27 which all lie in the range
at 3312.8 and 3304.3 cm1 observed in the Ar-matrix and of fD/fM ≈ 1368.
assigned to (DMA)2.22 In an argon matrix, one typically observes An IR intensity increase is often observed for the hydrogen
a redshift of transitions by around 25 cm1.47 stretching vibration of a hydrogen bond donor and is often
It is clear from Tables 2 and 3 that there is a large variation in considered as a defining feature of a typical hydrogen bond.45
the calculated binding energies and the calculated frequency shift In the ammonia dimer, the NH-stretching vibration is increased
and intensity enhancement with the range of theoretical methods by up to a factor of 9 compared to the monomer.49 However,
used. The large CP correction in the calculated binding suggests the NH-stretching fundamental transition in (DMA)2 is calcu-
that the cc-pVDZ and aug-cc-pVDZ basis sets are too small to lated to be ∼600 times stronger than that of the monomer.
describe well the weak interactions in the dimer. The calculated The IR intensity is determined mainly by the derivative of
IR intensities also vary significantly between the HO and AO the electronic dipole moment of the molecule, which changes
vibrational models when the aug-cc-pVDZ basis set is used. significantly between the monomer and the dimer. Furthermore,
12101 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104
The Journal of Physical Chemistry A ARTICLE

the NH-stretching fundamental vibration of DMA is very weak


due to a very small first derivative of dipole moment function.3,50
Comparison of the experimental integrated band intensity and
the QCISD/aug-cc-pVTZ calculated oscillator strength of the
NHb-stretching band indicates that there is only about 0.1%
dimer present at room temperature with a total DMA pressure of
700 Torr. The large intensity enhancement of the NH-stretching
fundamental in the DMA dimer makes it possible to observe
the dimer in the gas phase, even at room temperature where
only a very small fraction of DMA undergoes dimerization.
The intensity of the NHf-stretching transition is calculated to
be 2.6 times stronger than the NH-stretching transition in DMA
monomer; however this small enhancement is not sufficient for
us to observe the NHf-stretching transition.
For comparison, we also calculated the intensity of the first
NH-stretching overtone transition of (DMA)2 at the QCISD/
aug-cc-pVTZ level with our AO local mode model, to see if we Figure 5. The temperature dependence of the NH-stretching band in
could record NH-stretching overtone transitions of the DMA the DMA dimer. A path length of 2.4 m was used. At a given temperature,
dimer. For the DMA monomer, it was found that the first the spectrum was measured with a total pressure of 200 Torr.
NH-stretching overtone was stronger than the fundamental
transition.3 However, we predict the first NHb-stretching over-
tone of (DMA)2 to have only 0.01% of the intensity of the
fundamental vibration of the dimer and only 1% of the intensity
of the first overtone transition of DMA. This intensity decrease
of the first NHb-stretching overtone transition is also typical for
NH or OH bonds involved in hydrogen bonding.47,50 This is part
of the reason why overtones of complexes are difficult to measure
at room temperature.1,51,52
Enthalpy of Hydrogen Bond Formation. If we assume
that the enthalpy of a chemical equilibrium is independent of
temperature, we can use the van’t Hoff equation to write53
ln Kp ¼  ΔH=RT þ C ð5Þ

where Kp is the equilibrium constant, ΔH the enthalpy of a chemical


equilibrium, R the gas constant, T the absolute temperature, and C
an integration constant. The enthalpy ΔH can be determined if we
can obtain Kp at different temperatures. In our case, Kp is the
equilibrium constant of DMA dimerization and ΔH is the enthalpy
of hydrogen bond formation for the DMA dimer. As seen in eq 4, Figure 6. Linear least-squares fitting of the van’t Hoff equation plot for
Kp is proportional to pD on the condition of keeping pM a constant. the DMA dimer.
The dimer pressure pD is proportional to the measured integrated
absorbance of the dimer band. Therefore, we have 200 Torr for the high pressure spectra. Similarly, since we will use
0 appropriate weights on the low pressure spectrum used in the
lnðAÞ ¼  ΔH=RT þ C ð6Þ
spectral subtraction, the pressure cannot be too low, otherwise
where A is the integrated absorbance of the dimer band and C0 is a the uncertainties would be amplified. Finally, we decided to
new constant. use 20 Torr in the low pressure spectrum, which will need a
We recorded the IR spectra of (DMA)2 in the temperature weighting of around 10 in the subtraction. We used the ideal gas
range from 296 to 368 K to determine the enthalpy of hydrogen law to determine the filling pressure necessary to obtain 200 Torr
bond formation for the DMA dimer. In general, to get the dimer pressure at each of the elevated temperatures.
spectrum at a certain temperature, two absorption spectra are Representative IR spectra in the NH-stretching region of
needed: a high-pressure spectrum of DMA and its dimer and DMA dimer after spectral subtraction are shown in Figure 5
a low pressure spectrum where dimer contribution is minimal. and the van’t Hoff equation plot for DMA dimer is shown in
The latter is a pure DMA monomer spectrum necessary for Figure 6. Because of the weighting of the low pressure spectra,
subtraction. The vapor pressure of DMA should be kept constant the quality of the spectra in Figure 5 is not as good as the spectra
for all the high pressure measurements. The pressure of the in Figure 3. We used intergrated band areas as the measure
dimer in the mixture is less than 0.2% of the total pressure and of the absorbance in each of the different temperature spectra.
can therefore be neglected in the total pressure. We select After linear least-squares fitting, the dimerization enthalpy
the constant pressure such that the DMA dimer spectrum after ΔH in the temperature range 296368 K is determined to be
subtraction should be measurable and the absorbance of the 23.8 ( 2.2 kJ mol1 from the slope in Figure 6. We calculated
NH-stretching band in the spectrum should not saturate the the enthalpy of dimerization to be 23.1 kJ mol1 with
detector. After several test measurements, we decided to use the QCISD/aug-cc-pVDZ method and for the higher level
12102 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104
The Journal of Physical Chemistry A ARTICLE

QCISD/aug-cc-pVTZ method with a B3LYP/aug-cc-pVTZ most stable conformer of the dimer is characterized by a non-
thermodynamic correction we obtain 20.4 kJ mol1, both in linear hydrogen bond NH 3 3 3 N between the two DMA
good agreement with our experimental determination. molecules. The IR spectrum of the dimer shows a characteristic
The enthalpy ΔH of H-bond formation is directly related to small redshift of the NH-stretching vibrational band and a
the stabilization energy of a complex.54 In a stronger H-bonded pronounced intensity increase. The large increase in IR intensity
complex like methanol-trimethylamine (MeOH-TMA), the makes it possible to obtain the IR spectrum of the dimer in the
observed enthalpy is approximately 29 kJ mol1.55,56 The gas phase at room temperature. The detected fundamental NH-
enthalpies of H-bond formation in weaker complexes have been stretching band of the dimer is obtained by subtracting monomer
determined to be 14.8 and 19.1 kJ mol1 for DMS-MeOH spectra and shows the expected pressure square dependence.
and DME-MeOH, respectively.27 However, in ref 27, no correc- This provided evidence of the DMA dimer in the gas phase.
tion for change in total pressures was considered in the measure- On the basis of the measurements of the temperature depen-
ments at different temperatures; therefore, at temperatures dence of the dimer spectra, the enthalpy of hydrogen bond
higher than room temperature, the pressure of the monomers formation in DMA dimer in the temperature range of 296368 K
are slightly higher than the pressure read when the cell was filled was determined to be 23.8 ( 2.2 kJ mol1 in agreement with
with sample gas and the determined ΔH values are likely a few theoretical predictions. We combine the experimental integrated
kJ mol1 too small. For comparison, we also measured ΔH without intensity of the NH-stretching dimer band with anharmonic local
corrections for pressure changes and obtain a value of 16.9 ( mode intensity calculations to obtain an estimate of the thermo-
3.0 kJ mol1. The linear least-squares fitting of van’t Hoff equation dynamic dimerization equilibrium constant of 1.4  103 atm1
plot of DMA dimer for the experiment without pressure correction is at room temperature.
given in Supporting Information (Figure S1).
Thermodynamic Dimerization Equilibrium Constant. As
’ ASSOCIATED CONTENT
seen from eq 4, at a certain temperature, the thermodynamic
dimerization equilibrium constant Kp can be determined pro-
vided that the partial pressure of DMA (pM) and DMA dimer
bS Supporting Information. The linear least-squares fitting
of van’t Hoff equation plot of DMA dimer without pressure
(pD) could be measured. This has been used to determine Kp for correction; a plot of pD against pM2 to determine the thermo-
dimers with large Kp values.57 However, due to the small changes dynamic dimerization equilibrium constant; calculated frequen-
in pD for different total pressure, this is impossible for DMA. cies and intensities for DMA and DMA dimer with normal mode
Instead, we estimate the partial pressure of the dimer pD from the method; calculated NH-stretching fundamental and overtone
measured and anharmonically calculated NH-stretching intensi- frequencies and oscillator strengths for DMA dimer with an AO
ties. At the QCISD/aug-cc-pVTZ level, the NH-stretching local mode model; calculated NH-stretching fundamental fre-
fundamental transition in (DMA)2 is calculated to be 571 times quency and relative intensities for the two conformers A and B of
stronger than that of the monomer. The ratio of the experimental DMA dimer; z-matrices of DMA and DMA dimer (conformer B)
integrated band intensity and the theoretical oscillator strength at QCISD/aug-cc-pVTZ level. This material is available free of
gives the partial pressure of the dimer.36 A plot of pD against charge via the Internet at http://pubs.acs.org.
p2M for each of the experiments is shown in Figure S2 of Supporting
Information. The slope of the least-squares fitting of these data is the ’ AUTHOR INFORMATION
thermodynamic dimerization equilibrium constant, Kp, which
we determined to be 1.4  103 atm1 at a temperature of Corresponding Author
296 K. The Kp determination method depends on the calculated *E-mail: hgk@chem.ku.dk. Fax: 45-35320322. Phone: 45-35320334.
intensity enhancement, which varies with theoretical method used.
We have used our highest level value obtained with the QCISD/
aug-cc-pVTZ AO method, with a fD/fM ≈ 571. Uncertainty in the ’ ACKNOWLEDGMENT
calculated intensity will be reflected in uncertainty in the deter- We thank Lauri Halonen, Benjamin J. Miller, and Joseph R.
mined Kp value. However, at the aug-cc-pVTZ basis set level, the Lane for helpful discussions and Rene W. Larsen for help with the
intensity ratio of our different approaches is in the range (363 to FTIR measurements. We are grateful for support from The
835), and thus our Kp value is accurate to within a factor of 2. The Danish Council for Independent Research—Natural Sciences
decrease in band absorbance with increasing temperature, as and the Danish Center for Scientific Computing.
shown in Figure 5, clearly illustrate the decrease in the equilibrium
constant with increasing temperature. We have estimated Kp
theoretically with a statistical thermodynamics procedure and find ’ REFERENCES
it to be 3.8  105 atm1 at the B3LYP/aug-cc-pVTZ level and (1) Fang, H. L.; Swofford, R. L.; Compton, D. A. C. Chem. Phys. Lett.
2.6  103 atm1 at the QCISD/aug-cc-pVTZ level with B3LYP/ 1984, 108, 539.
aug-cc-pVTZ thermodynamic values.58 The loose binding of (2) Marinov, D.; Rey, J. M.; Muller, M. G.; Sigrist, M. W. Appl. Opt.
complexes with the B3LYP method is in agreement with the 2007, 46, 3981.
smaller Kp value obtained with this method. The better calculation (3) Miller, B. J.; Du, L.; Steel, T. J.; Paul, A. J.; Sodergren, A. H.; Lane,
agrees reasonably well with our combined experimental and J. R.; Henry, B. R.; Kjaergaard, H. G. J. Phys. Chem. A 2011submitted.
(4) Bako, I.; Palinkas, G. THEOCHEM 2002, 594, 179.
theoretical determined value.
(5) Boese, A. D.; Chandra, A.; Martin, J. M. L.; Marx, D. J. Chem.
Phys. 2003, 119, 5965.
’ CONCLUSIONS (6) Bricknell, B. C.; Ford, T. A. J. Mol. Struct. 2010, 972, 99.
(7) Buck, U.; Gu, X. J.; Krohne, R.; Lauenstein, C.; Linnartz, H.;
The dimethylamine dimer in the gas phase has been investi- Rudolph, A. J. Chem. Phys. 1991, 94, 23.
gated by FTIR and ab initio calculations. The geometry of the (8) Hippler, M. J. Chem. Phys. 2007, 127, 084306.

12103 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104


The Journal of Physical Chemistry A ARTICLE

(9) Hippler, M.; Hesse, S.; Suhm, M. A. Phys. Chem. Chem. Phys. (37) Simon, S.; Duran, M.; Dannenberg, J. J. J. Phys. Chem. A 1999,
2010, 12, 13555. 103, 1640.
(10) Lane, J. R.; Kjaergaard, H. G. J. Chem. Phys. 2009, 131, 034307. (38) Howard, D. L.; Robinson, T. W.; Fraser, A. E.; Kjaergaard, H. G.
(11) Lane, J. R.; Vaida, V.; Kjaergaard, H. G. J. Chem. Phys. 2008, Phys. Chem. Chem. Phys. 2004, 6, 719.
128, 034302. (39) Niefer, B. I.; Kjaergaard, H. G.; Henry, B. R. J. Chem. Phys. 1993,
(12) Mmereki, B. T.; Donaldson, D. J. J. Phys. Chem. A 2002, 106, 99, 5682.
3185. (40) Robinson, T. W.; Kjaergaard, H. G.; Ishiuchi, S. I.; Shinozaki,
(13) Nelson, D. D.; Fraser, G. T.; Klemperer, W. J. Chem. Phys. 1985, M.; Fujii, M. J. Phys. Chem. A 2004, 108, 4420.
83, 6201. (41) Rong, Z. M.; Kjaergaard, H. G. J. Phys. Chem. A 2002, 106, 6242.
(14) Nelson, D. D.; Fraser, G. T.; Klemperer, W. Science 1987, (42) Kjaergaard, H. G.; Yu, H. T.; Schattka, B. J.; Henry, B. R.; Tarr,
238, 1670. A. W. J. Chem. Phys. 1990, 93, 6239.
(15) Rodham, D. A.; Suzuki, S.; Suenram, R. D.; Lovas, F. J.; (43) Wollrab, J. E.; Laurie, V. W. J. Chem. Phys. 1968, 48, 5058.
Dasgupta, S.; Goddard, W. A.; Blake, G. A. Nature 1993, 362, 735. (44) Mayer, P. M.; Keister, J. W.; Baer, T.; Evans, M.; Ng, C. Y.; Hsu,
(16) Tubergen, M. J.; Kuczkowski, R. L. J. Mol. Struct. 1995, 352, C. W. J. Phys. Chem. A 1997, 101, 1270.
335. (45) Arunan, E.; Desiraju, G. R.; Klein, R. A.; Sadlej, J.; Scheiner, S.;
(17) Wales, D. J.; Stone, A. J.; Popelier, P. L. A. Chem. Phys. Lett. Alkorta, I.; Clary, D. C.; Crabtree, R. H.; Dannenberg, J. J.; Hobza, P.;
1995, 240, 89. Kjaergaard, H. G.; Legon, A. C.; Mennucci, B.; Nesbitt, D. J. Pure Appl.
(18) Lambert, J. D.; Strong, E. D. T. Proc. R. Soc. London, Ser. A 1950, Chem. 2011, 83, 1619.
200, 566. (46) Du, L.; Lane, J. R.; Kjaergaard, H. G. Unpublished, 2011.
(19) Wolff, H.; Gamer, G. J. Phys. Chem. 1972, 76, 871. (47) Schofield, D. P.; Kjaergaard, H. G. Phys. Chem. Chem. Phys.
(20) Bernardh, M. C.; Sandorfy, C. J. Chem. Phys. 1972, 56, 3412. 2003, 5, 3100.
(21) Odutola, J. A.; Viswanathan, R.; Dyke, T. R. J. Am. Chem. Soc. (48) Kjaergaard, H. G.; Garden, A. L.; Chaban, G. M.; Gerber, R. B.;
1979, 101, 4787. Matthews, D. A.; Stanton, J. F. J. Phys. Chem. A 2008, 112, 4324.
(22) Bohn, R. B.; Andrews, L. J. Phys. Chem. 1991, 95, 9707. (49) Slipchenko, M. N.; Sartakov, B. G.; Vilesov, A. F.; Xantheas,
(23) Tubergen, M. J.; Kuczkowski, R. L. J. Chem. Phys. 1994, 100, S. S. J. Phys. Chem. A 2007, 111, 7460.
3377. (50) Kjaergaard, H. G.; Low, G. R.; Robinson, T. W.; Howard, D. L.
(24) Pradeep, T.; Hegde, M. S.; Rao, C. N. R. J. Mol. Spectrosc. 1991, J. Phys. Chem. A 2002, 106, 8955.
150, 289. (51) Garden, A. L.; Halonen, L.; Kjaergaard, H. G. J. Phys. Chem. A
(25) Cabaleiro-Lago, E. M.; Rios, M. A. J. Chem. Phys. 2000, 113, 2008, 112, 7439.
9523. (52) Shillings, A. J. L.; Ball, S. M.; Barber, M. J.; Tennyson, J.; Jones,
(26) Howard, D. L.; Kjaergaard, H. G. J. Phys. Chem. A 2006, 110, R. L. Atmos. Chem. Phys. 2011, 11, 4273.
9597. (53) Paiva, J. C. M.; Gil, V. M. S.; Correia, A. F. J. Chem. Educ. 2002,
(27) Howard, D. L.; Kjaergaard, H. G. Phys. Chem. Chem. Phys. 2008, 79, 583.
10, 4113. (54) Curtiss, L. A.; Blander, M. Chem. Rev. 1988, 88, 827.
(28) Henry, B. R. Acc. Chem. Res. 1977, 10, 207. (55) Fild, M.; Swiniars., Mf; Holmes, R. R. Inorg. Chem. 1970, 9, 839.
(29) Henry, B. R. Acc. Chem. Res. 1987, 20, 429. (56) Millen, D. J.; Mines, G. W. J. Chem. Soc., Faraday Trans. 1974,
(30) Henry, B. R.; Kjaergaard, H. G. Can. J. Chem. 2002, 80, 1635. 70, 693.
(31) Howard, D. L.; Jorgensen, P.; Kjaergaard, H. G. J. Am. Chem. (57) Crawford, M. A.; Wallington, T. J.; Szente, J. J.; Maricq, M. M.;
Soc. 2005, 127, 17096. Francisco, J. S. J. Phys. Chem. A 1999, 103, 365.
(32) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; (58) Vaida, V.; Headrick, J. E. J. Phys. Chem. A 2000, 104, 5401.
Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.;
Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.;
Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson,
G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.;
Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.;
Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;
Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.;
Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.;
Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski,
V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick,
D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui,
Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;
Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith,
T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill,
P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.
Gaussian 03, revision E.01; Gaussian, Inc.: Wallingford, CT, 2004.
(33) Werner, H.-J.; Knowles, P. J.; Manby, F. R.; Sch€utz, M.; Celani,
P.; Knizia, G.; Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.;
Adler, T. B.; Amos, R. D.; Bernhardsson, A.; Berning, A.; Cooper, D. L.;
Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Goll, E.; Hampel, C.;
Hesselmann, A.; Hetzer, G.; Hrenar, T.; Jansen, G.; K€oppl, C.; Liu, Y.;
Lloyd, A. W.; Mata, R. A.; May, A. J.; McNicholas, S. J.; Meyer, W.; Mura,
M. E.; Nicklass, A.; Palmieri, P.; Pfl€uger, K.; Pitzer, R.; Reiher, M.;
Shiozaki, T.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T.; Wang,
M.; Wolf, A. MOLPRO, version 2009.1.
(34) Boys, S. F.; Bernardi, F. Mol. Phys. 1970, 19, 553.
(35) Hippler, M. J. Chem. Phys. 2005, 123, 204311.
(36) Chung, S.; Hippler, M. J. Chem. Phys. 2006, 124, 214316.

12104 dx.doi.org/10.1021/jp206762j |J. Phys. Chem. A 2011, 115, 12097–12104

You might also like