You are on page 1of 9

Collisional ionization between alkali atoms and some methane derivatives: Electron affinities for CH3NO2, CF3I, and

CF3Br
R. N. Compton, P. W. Reinhardt, and C. D. Cooper Citation: J. Chem. Phys. 68, 4360 (1978); doi: 10.1063/1.435514 View online: http://dx.doi.org/10.1063/1.435514 View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v68/i10 Published by the American Institute of Physics.

Additional information on J. Chem. Phys.


Journal Homepage: http://jcp.aip.org/ Journal Information: http://jcp.aip.org/about/about_the_journal Top downloads: http://jcp.aip.org/features/most_downloaded Information for Authors: http://jcp.aip.org/authors

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Collisional ionization between alkali atoms and some methane derivatives: Electron affinities for CH 3N02 , CF31, and CF3Br8 )
R. N. Compton and P. W. Reinhardt
Chemical Physics Section, Health and Safety Research Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37830

C. D. Cooper
Department of Physics and Astronomy, University of Georgia, Athens, Georgia 30602 (Received 30 December 1977)

The negative ion products resulting from collisions between orthogonal beams of alkali atoms (Na, K, Cs) and the methane derivatives CH 3CN, CH 3N02 CF3Br, and CF3I have been studied in the energy range from reaction thresholds to ~40 eV (LAB). Stable negative ions with masses corresponding to the last three molecules were detected and the following electron affinities are derived from measurements of the energy threshold for the ion pair production reactions: E.A.(CH 3N02 ) = 0.44 :,:g:j eV; E.A.(CF3Br) =0.910.2 eV, and E.A.(CF3I)= 1.570.2 eV. From measurements of the difference between the energy threshold for the appearance of various fragment ions and the parent ion, the following bond dissociation energies are deduced: D(CH r N02) = 0.S60.2 eV; D(CFrBC) = 0.540.2 eV and D(CFrI-) = 0.320.2 eV. An argument is presented which adds further strength to the suggestions of Williams et al. and Jordan and Wendoloski that electron binding to CH 3CN is dominated by the dipole field.

I. INTRODUCTION

II. EXPERIMENTAL

Collisions between thermal alkali atoms and methyl halides resulting in methyl radicals and alkali halides have been thoroughly studied in molecular beam experiments. 1 These reactions are believed to proceed via an "electron jump" mechanism from the covalent to the ionic surface. For the case of alkali atoms colliding with methyl iodide, the reaction proceeds through a rebound mechanism in which the alkali iodide product rebounds into the backward direction. 2 At sufficiently high collision energies (i. e., energies greater than the threshold for chemiionization) the ionic state can separate into ions and neutrals. A lucid description of these reactions has been presented by Moutinho, Aten, and Los. 3 Rothe, Tang, and Reck4 have studied the collisional ionization reactions between energetic cesium atoms and CF 3X (X= Cl, Br, and I), which produce Xand F -. From consideration of the intensity ratio F-/ X- as a function of collision energy they suggested that the reaction proceeds via a stripping mechanism. In the cases where parent negative ions are formed, reasonably accurate ( O. 2 e V) electron affinities can be determined. In this manner, Tang, Mathur, Rothe, and Reck 5 derived an electron affinity for CF 31 of 1. 4 O. 2 eVfrom the reaction Cs+CF 3I-Cs++CF 3I", whereas McNamee, Lacmann, and Herschbach 6 obtained a value of 2.2 0 2 eV from the reaction K + CF 3I- K ++ CF 3I-.
0

The basic collisional ionization apparatus used in this work has been described earlier. 7 This apparatus was later modified 8 to employ crossed atomic and molecular beams to minimize Doppler broadening in the center of mass system, thereby improving the energy resolution. The energy thresholds of mass identified negative ion signals were determined by comparing previously determined thresholds such as 0;; from O2, NO;; from N02, and SFii and SF; from SF 6 (see ReL 8 for a complete treatment of this procedure). Except in cases where clean thresholds are not observed (i. eo, when considerable "tailing" is observed due to the nature of the cross section), the onsets are believed to be accurate to within 0.2 eV (LAB). Expected accuracy in the center of mass system is improved somewhat depending upon the masses of the collidants. The dependence of the reaction threshold upon the temperature of the target gas was determined in a few instances. The method employed will be described at the appropriate point in the text.
III. RESULTS AND DISCUSSION A. Methyl cyanide (acetonitrile, CH 3 CN)

The present paper describes measurements of thresholds for the production of negative ions from collisions between fast alkali atoms and seven methane derivatives. Electron affinities and bond dissociation energies of the negative ions are deduced for CH3N02 , CF 3Br, and CF 31.

a)Research sponsored by the Department of Energy under contract with the Union Carbide Corporation.
4360

Takeda and Williams 9 first observed an ESR spectrum of CH 3CN" upon gamma irradiation of the lower crystalline phase of acetonitrile at 77 OK. Williams and coworkers 10 were interested in the possibility that the large dipole moment and high degree of polarizability afforded by the nitriles could trap electrons in glasses and crystals. In fact, Bonin et al. l1 presented evidence suggesting that the trapped electrons were bound in the dipole fields of the highly polar CH3CN molecules for which IJ. (CH3CN) = 3. 92 D. This suggestion was proposed following theoretical calculations 12 which showed
1978 American Institute of Physics

J. Chern. Phys. 68(10), 15 May 1978

0021-9606/78/6810-4360$01.00

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Compton, Reinhardt, and Cooper: Electron affinities for CH 3N0 2 , CF31, and CF3Br that the fixed dipole can bind an electron if J1?- 1. 625 D. Later calculations indicated that electrons would be permanently bound to real polar molecules with dipole moments greater than ~ 2.5 D. 13 14 The CH3CN- ions observed in the experiments of Williams and co-workers have a calculated CCN angle of 130. 15 Gas phase CH 3CN- ions have been produced by charge exchange between atoms (or molecules) in highly excited Rydberg states and CH3CN.16-18 An upper limit to the thermal electron attachment rate constant for CH3CN was measured 17 to be <:; 1. 24 X 10- 14 cm 3 sec-I in an electron swarm experiment for pressures up to a few Torr. Mothes et al. 19 used the ECR technique to measure a value of 7.2 x 10- 12 cm 3sec-I. However, the value by Mothes et al. 19 is likely to be too large (see footnote to Ref. 19). The rapid reaction of CH 3CN with Rydberg states is not in accord with its small thermal electron attachment rate. Acetonitrile appears to be the only exception thus far to the rule suggested by Matsuzawa 20 that electron attachment rates and Rydberg reaction rates are equivalent. We will offer a possible explanation for this interesting contradiction in what follows. In the present experiment, we have been unable to detect CH 3CN- ions via collisions of fast Na, K, or Cs atoms with CH3CN. However, the dissociative ions CNand CH 2CN- were observed in all three cases. The onset for CN- occurs at 6.92 eV (CM) for Na collisions (see Fig. 1). If we subtract the ionization potential of sodium from this energy onset and assume that the products are created with zero internal energy and no kinetic energy in the center of mass, we find the energy difference D(CH3-CN) - E. A. (CN) to be 1. 65 0.2 eV (average of eight measurements). This onset occurs to within O. 2 e V of the onset observed in the dissociative electron attachment reaction reported earlier (see Fig. 4 of Ref. 17). The thermochemical value is L 5 eV as

4361

determined from measured values of D(CH3 - CN) (<:; 5. 320.03 eV)21 and E.A. (CN) (3. 820. 002 eV).22 Measurements of CN- produced by Cs and K collisions are also in good agreement with this thermochemical value. Let us consider the known information about electron binding to acetonitrile. (1) CH 3CN possesses 16 valence electrons and from molecular orbital consideration electron binding to CH3CN should be weak. (2) The dipole moment of CH 3CN is 3.92 D and well exceeds the minimum moment required to bind an electron to a real molecule. 13 14 In this connection, Jordan and Wendoloski 23 calculate that the (Born-Oppenheimer) E.A. of CH 3CN is at least 2. 5 and perhaps as large as 12 times greater than the rotational constant. Crawford and Garrett have also shown 24 that this criterion is sufficient to insure electron binding in a more exact treatment of the problem. (3) Reactions of highly excited Rydberg atoms with CH 3CN produce copious amounts of CH3CNalthough the thermal electron attachment rate for CH3CN is small (<:; 1. 24 X 10- 14 cm 3 sec- I),17 (4) Charge exchange between fast alkali atoms (which contain a reasonably well bound outer s electron) and CH 3CN produces no detectable CH3CN-. The assumption that electrons can be bound to the dipole field of CH 3CN can lead to an explanation for the last three experimental observations. Electrons bound to a dipole field exist in large orbits and are described by very diffuse wavefunctions, much like those of an electron in a Rydberg orbital of large n. 25 Thus, charge exchange in this case would be efficient because of favorable overlap between the two electronic wavefunctions. Loosely stated, the electron gently changes "centers" of force in the collision. Free electron attachment into the dipole field of a molecule to form a bound state would require radiative or collisional stabilization, which may be unlikely during the short lifetime (~1O- 15 sec) of the electron-molecule sys-

Ii) :!::

2
Z

:0 ...

... ....

>. ... o

IW

0:: 0::

::::>

FIG. 1. Negative ion yields resulting from collisions of fast sodium atoms with acetonitrile as a function of the sodium atom energy. Similar yields were recorded for collision between cesium or potassium and acetonitrile.

10

15

20

25

30

SODIUM ATOM ENERGY (eV)


J. Chern. Phys., Vol. 68, No. 10,15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

4362

Compton, Reinhardt, and Cooper: Electron affinities for CH 3 N0 2 , CF31, and CF3Br

FIG. 2. Negative ion yields resulting from collisions of fast cesium atoms with nitromethane and nitromethane-d 3 as a function of the cesium atom energy.

10

15

20

25

30

CESIUM ATOM ENERGY (eV)

tem. Also, because of the fragile nature of the ion, collisional detachment may be more probable than stabilization. The fact that CHaCN- was not produced in measurable quantities in the present experiment may be related to the gross difference between the wavefunction representing an electron in the lowest s orbital of an alkali and that representing a very diffuse electron in the field of a molecular dipole,25 The distance at which the covalent (Cs + CHaCN) surface crosses with the ionic (Cs + CH3 CN-) surface is roughly 3 A assuming that the electron affinity of CH3CN is approximately zero and assuming only the point charge Coulomb interaction. Depending upon the exact value of the polarizability of CH 3CN -, which may be large, the crossing distance could be extended to ~ 7 A. Because of the extreme diffuseness of the wavefunction describing a weakly bound electron in a dipole field, the probability that an electron would be transferred at this distance would be small. Therefore, our inability to observe CHaCNfrom alkali collisions may be related to the nature of the orbital into which the electron is attached and may not be due to the smallness of the electron affinity of CH3CN. For example, the reaction of Cs with NO (E.A. = O. 024 eV)26 produces NO- with a reasonably large cross section. 7
It is also possible that collisional ionization does occur between a fast alkali atom and CH 3CN to produce the CH 3CN- ion in a state of rotational excitation sufficient for rapid detachment of the weakly bound electron. 13 Furthermore, the electric fields present in the time-offlight mass spectrometer may also be large enough to detach the weakly bound electron, Both of these possibilities are being examined in a separate experiment now in progress, In summary, the interpretation of ex-

perimental observations presented here and earlier 17 support earlier notions l1 2a that electron binding to CHaCN occurs primarily through the dipole field, Further studies of energy and angular distributions of photodetached electrons from CHaCN- or studies of reactions of well defined Rydberg states with CHaCN may be more conclusive. The latter study may afford a measurement of the electron affinity of CHaCN.

Reactions of fast Na, K, or Cs atoms with nitromethane produce CHaNO;;, NO;;, and 0-. Figure 2 illustrates the signals observed for the case of Cs colliding with CH 3N0 2 and CD 3N0 2 No major differences between CH aN0 2 and CD aN0 2 are observed except that the NO'2 and 0- ions occur approximately 0.5 eV (LAB) lower in the deuterated case. This is approximately the energy difference expected due to the different reduced masses in the two collisions. The CD 3N0'2 ion onset is approximately 0.2 eV (LAB) lower than the CH aNO'2 onset. The energy scales were calibrated using NOi from N0 2 as shown in Fig. 2. The electron affinities calculated for CH 3N0 2 and CD 3N0 2 are 0,43 and 0.39 eV, respectively. The difference between these two values is insignificant, although such an isotope effect would be expected. In cases where the vibrational frequencies of the negative ion are of lower energy than that of the neutral molecule, the difference in zero point energies requires that the heavier molecule have a smaller electron affinity. The electron affinity derived from the data shown in Fig. 2 happens to be in accord with this expectation; however, the precision of the present experiments is insufficient to reproduce this difference (see Table I where the average of many data runs gives the reverse ordering in magnitude for the electron affinities),

J. Chem. Phys., Vol. 68, No. 10, 15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Compton, Reinhardt, and Cooper: Electron affinities for CH 3 N0 2 , CF31, and CF3Br
TABLE I. Electron affinities for CF 3Br, CF3I, CH 3N0 2 and bond dissociation energies of their negative ions. D(CF 3-X-) or D(CH3-X-) 0.540.2 eV 0.350.2 eV 0.29 0.2 eV 0.60.2 eVd 0.S0.2 eyd 0.60.2 eVd 1.20.2 eVa 0.380.2 eV b

4363

Alkali
K

Molecule CF 3Br CF3I CF3I CF3I CH3N0 2


CH3NO z

Electron affinity 0.910.2eV 1.540.2eV 1.60.2 eV 0.35O.2 eV 0.460.2 eV O. 5 O. 2 eV

Others

Others

Na
K

Cs
K

2.20.2eVa 1.40.2 eVb D.45O.OS eVo

Cs Cs

CD3 N0 2

ap. E. McNamee, K. Lacmann, and D. R. Herschbach, Faraday Discuss. Chern. Soc. 55, 318 (1973). bs. Y. Tang, B. P. Mathur, E. W. Rothe, and G. P. Reck, J. Chern. Phys. 64, 1270 (1976). oR. Freeman, Ph. D. Dissertation, Univ. of Houston (1971) [private communication, Dr. W. E. Wentworth, July (1977)]. <Qnset of NO:; is corrected for the temperature dependence on threshold. No temperature dependence was found for the threshold for CHsNOi.

Figure 3 shows the negative ion thresholds from collisions of crossed beams of potassium and CH 3N02. The threshold for NO; from nitrogen dioxide is used as an energy scale calibration. Our electron affinity for CH 3N0 2 determined from five such measurements is, 0.35 0.2 eV (see Table I). We have also investigated the influence of target gas temperature upon reaction thresholds. Internal vibration-rotation energy in the target molecule can effectively lower the ion-pair reaction threshold. Nitromethane vapor was allowed to reach temperature equilibrium in a .resistance heated 1. 6 mm diam platinum tube. The hot vapor effuses from a 75 J.1. wide slit and crosses a fast potassium beam before being terminated by a liquid nitrogen cold trap. The temperature of the platinum tube was determined by a thermocouple and is checked in situ by observing the tube with an optical

pyrometer through a glass port. Observations were made over the temperature range from room temperature to ~ 1000 oK. No temperature dependence of the energy threshold for CH 3NO; could be detected. The onset for NOi decreased at the rate of ~ O. 0015 eV (LAB)/oK. The thermochemical values shown in Table I are corrected for this temperature dependence by linear extrapolation of the room temperature value to T = 0 oK. The threshold for 0- formation in the case of potassium and cesium is different. A threshold of 17.6 eV (LAB) is predicted for the case of Cs + CH 3N0 2 from knowledge of the ionization potential of cesium (3.893 eV), the bond-dissociation energy D(CHaNO-O) 3.1149 eV27 and the electron affinity of atomic oxygen 1. 462 eV. 28 The onset for 0- from CH 3N0 2 or CD 3N0 2 is ~ 21 eV (LAB) which indicates that the products of the collision are produced with internal and/or kinetic energy.

~ C
~

>. ... o

::::I

... ,
:0
IZ

a:: a:: => u

z o

FIG. 3. Negative ion yields resulting from collisions of fast potassium atoms and nitromethane as a function of the potassium atom energy. Also shown is the NOi signal from collisions between potassium and nitrogen dioxide which is used for energy scale calibration.

10

15

20

POTASSIUM ATOM ENERGY (eV)


J. Chern. Phys., Vol. 68, No. 10, 15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

4364

Compton, Reinhardt, and Cooper: Electron affinities for CH]N0 2 , CF31, and CF3Br

(J)

IZ

0:: 0:: :J U

FIG. 4. Negative ion yields resulting from collisions of fast sodium atoms with CFsBr as a function of the sodium atom energy.

10

12

14

16

18

SODIUM ATOM ENERGY (eV)

Dissociative electron attachment producing 0- ions occurs considerably above the thermodynamic threshold of 1. 65 eV.17 For the case of collisions between potassium and CH aN0 2, the calculated onset for 0- production is 9.79 eV (LAB). The measured onset is 9.8 O. 2 eV. Nitromethane attaches slow electrons and also reacts with highly excited Rydberg states to form CHaNO;.17 The dipole moment of nitromethane is 3.46 D and is sufficient to bind an electron into a stable orbit. However, the electron affinity observed in this work is too large to be attributed solely to dipole binding. The "dipole bound" state may be considered as an excited state of CHaNO;. INDO calculations for CHaNO; support conclusions based upon 14N and laC hyperfine splittings that the nitro-anion-radicals are pyramidal about the nitrogen atom. 29 Experiment and theory predict that the plane containing the two oxygen atoms is tilted with respect to the C-N direction by ~ 6. Thus a large geometry change of the CH aN0 2 structure is not expected upon addition of an extra electron. The onsets observed for the reactions (Cs, K) + CH aN0 2 - (Cs" K + CHg + NO; can be employed to deduce an electron affinity for N02 using the known bond dissociation energy of CH 3-N0 2 and assuming that the products are left in their ground states with no excess energy in the center of mass. If we assume that D(CH a-N0 2 ) = 2. 52 eV, gO a value of E. A. (N0 2) = 2.3 0.2 eV is obtained. This value is very close to the photodetachment value of 2.36 O. 1 eV. 31 Also the bond dissociation energy of the negative ion, D(CHa-NO;), has been determined by measuring the difference in threshold for NO; and CHaNO; (see Table I).

Fig. 4. Table I summarizes the various energy thresholds for the ions observed. Tang et al. 5 have previously reported on ions resulting from cesium collisions with CFaBr. They observed CFaBr- but were unable to deduce an electron affinity due to insufficient signal to noise at threshold. The electron affinity of CF aBr is determined in this work to be 0.91 0.2 eV, which is an average of 22 measurements. From the onset of Bra value of - O. 4 0.2 eV is determined for the difference in the bond dissociation energy, D(CFa-Br) and the electron affinity of bromine. We could detect no change in the threshold for either CFgBr- or Br- as the CFaBr was heated. The bond dissociation energy of CFa-Br is 2.95 O. 15 eV,a2 and the electron affinity of Br is 3.364 e V, 28 yielding a difference of - O. 41 e V which is identical to our value.

D. Perfluoromethyl iodide (CF 3 1)


Collisions of fast sodium or cesium atoms with CF 31 produce the ions CF aI", 1-, F -, and CF"3 as shown in Fig. 5. Pertinent data are summarized in Table I. Again, no temperature dependence was observed for any of the thresholds. Tang et al. 5 also did not detect a shift in the 1- threshold at a target gas temperature of 577C. From measurements of the onsets for C F al" and 1", an ionic bond dissociation energy of O. 32 O. 2 e V is obtained. This value compares well with the value of O. 38 O. 16 eV obtained by Tang et al. 5 but is considerably different from the value of 1. 2 O. 2 e V reported by McNamee, Lacmann, and Herschbach. 6 In principle, the CFg-I bond dissociation energy could be calculated from the 'measured onset for the formation of I" since the electron affinity of atomic iodine is well known 3.061 eV. 28 This exercise results in a value for D(CF 3-1) of 2. OO. 2 eV compared to the experimental

C. Perfluoromethyl bromide (CF 3 Br)


Collisions between fast sodium atoms and CF gBr produce the ions CFaBr-, Br-, F -, and CF; as shown in

J. Chern. Phys., Vol. 68, No. 10, 15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Compton, Reinhardt, and Cooper: Electron affinities for CH 3N0 2 , CF31, and CF3Br

4365

FIG. 5. Negative ion yields resulting from collisions of fast sodium atoms with CFaI as a function of the sodium atom energy.

10

12

14

18

SODIUM ATOM ENERGY (eV)

value of 2. 35 O. 05. S2 Tang et al. 5 also obtained a value of 2.05 0.2 eV. The discrepancy does not appear to be attributable to the presence of internal energy in the CFsI target molecules owing to room temperature population of rotational and vibrational states. It is unlikely that the reported bond dissociation energy for CF sl (Table I) is incorrect by as much as 0.3 eV. Parent negative ions of the Freons CF sCI, CF sBr, and CF sl have been detected by ESR spectroscopySS upon irradiation at 77 OK of solid solutions containing up to 5 mole percent of the CFsX parent compound in various solvents. The results suggest that the unpaired electron resides in an at (a*) antibonding orbital which is composed largely of the p orbitals from carbon and the unique halogens which lie along the C sv symmetry axis of the radical anion. It is suggestedSs that the CF sXspecies are probably true radical anions rather than radical-anion adducts such as the weakly bound CHs-Brand CHs-I" previously observed 34 in the acetonitrile matrix. The rather large bond energies for CFs-Br- (0.54 eV) and CFs-I- (0.32 eV) observed in this work supports this contention.
E. Other methane derivatives: A rapid survey of collisions of potassium atoms with other methane derivatives revealed the following information.

for D(CHs-Br) and E.A. (Br) are 2.9 eVand 3.364 eV (Refs. 36 and 28), which yield a difference of - O. 46 eV. The ions C2H- (strong) and CsH; (weaker) were observed from collisions of potassium with methyl acetylene.
CONCLUSIONS

In this work we have pre,sented electron affinities and bond dissociation energies for negative ions from a number of methane derivatives. In Fig. 6 an approximate sketch is shown of one slice of the potential energy surfaces for the negative ions studied in this work. The shapes are arbitrarily drawn to fit the asymptotic and equilibrium values. The surface for CF sBr-, CF sl", and CHsNOi are drawn through the maximum in the wavefunction representing the ground vibrational state of the neutral since these molecules are known to capture slow electrons. The minima in the CFsBr- and CF sl - surfaces are shifted to larger distances relative to the corresponding neutral to be consistent with the CNDO calculations, which predict that occupation of the a* orbital by the unpaired electron lengthens the C-X bond (e. g., C-CI distance in CFaCI increases from 1. 75 to 1. 95 A).37 The dashed surfaces which are drawn congruent with the CHsCN and CHsNOz surfaces represent the "dipole bound" negative ion state. The fact that stable CHsCN- ions are not produced by alkali collisions may be explained by assuming either that electron transfer from the alkali to the assumed "dipole bound" CHsCN- state is improbable due to poor overlap of electronic wavefunctions or that CHaCN- is produced in a rotational state with excitation sufficient to detach the electron tS (i. e., total energy is positive). The possibility also remains that feeble CHsCN- ions are produced but are destroyed by collisions with background CHsCN molecules or by field detachment due to ion draw out fields in the mass spectrometer. Electrons

Collisions of potassium atoms with methyl alcohol produce OH- and CHaO- ions of almost equal abundance. The onset for OH- yields D(CHa-OH)-E.A. (OH) = 1. 94 0.2 eV compared with a calculated value of 3.951. 825 eV (Ref. 35 and 28) or 2. 12 eV. Potassium atom collisions with CHaBr produced no detectable parent negative ions. The Br- ion is observed at a threshold collision energy of 4.03 eV (CM); subtracting the ionization potential of potassium (4.32 eV) gives D(CHs-Br)-E.A. (Br)=-0.3 eV. Accepted values

J. Chern. Phys., Vol. 68, No. 10, 15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

4366
6

Compton, Reinhardt, and Cooper: Electron affinities for CH 3N0 2 , CF31, and CF3Br

Ir-~I

manently bind an electron may be a facile way to produce "dipole bound" negative ions.

4
~

r---~

>

>-

a::

<9

w z 2 w

0
-i

--~

2
~

>

z 0 w
-i

a::

><9

-2
RRFIG. 6. Approximate sketch of negative ion potential energy "curves" derived from the present data. The dashed lines represent the hypothetical "dipole bound" states. In the case of CH 3N0 2 we assume that the "dipole bound" state and the valence state are separate and distinct. We further assume that the ionic surfaces cross the neutral surface in the region of the ground state vibrational wavefunction in accordance with the fact that these molecules attach thermal electron. We caution that this assumption, which is commonly invoked, needs further scrutiny [see, e.g., O. H. Crawford and B. J. D. Koch, J. Chem. Phys. 60, 4512 (1974)].

bound to permanent dipoles are expected to be weakly bound and diffuse, a condition which results in extreme fragility. No stable parent ions were observed in two other cases when potassium was made to collide with molecules possessing large dipole moments. Formamide (H 2NCOH) has a dipole moment of 3. 25 D. Only the ions NCO- and HNCOH- were produced. Both appeared at a collision energy of ~ 6.3 eV (C. M. ); 0- was also recorded at a higher energy. The dipole moment of dimethyl sulfoxide, (CH 3 hSO, is also quite large (3.96 D), yet only the ions CH 3SO-, SO-, OH -, and S- were produced upon collision with fast potassium atoms. Further studies of Rydberg reactions in well defined energy levels and alkali energy loss measurements are planned for molecules with large dipole moments (e. g., CH 3CN, H2NCOH, NH 2CN). Acetonitrile, HCN, formamide, or cyamamide (NH 2CN, j.L = 4.27 D) may be the best systems, aside from the alkali halides, in which to study stable states of the electron-dipole system. Furthermore, reactions of highly excited Rydberg states with molecules possessing dipole moments sufficient to per-

l D R. Herschbach, Adv. Chern. Phys. 10, 319 (1966). 2G. H. Kwai, J. A. Norris, and D. R. Herschbach, J. Chem. Phys. 52,1317 (1970); L. M. RaffandM. Karplus, J. Chem. Phys. 44, 1212 (1966); R. A. Budde, P. J. Kuntz, R. B. Bernstein, and R. D. Levine, Chern. Phys. Lett. 19, 7 (1973); R. Grice and D. R. Hardin, Mol. Phys. 21, 805 (1971). 3A. M. C. Moutinho, J. A. Aten, and J. Los, Chern. Phys. 5, 84 (1974). 4E. W. Rothe, S. Y. Tang, and G. P. Reck, Chern. Phys. Lett. 26, 434 (1974). 5S. Y. Tang, B. P. Mathur, E. W. Rothe, and G. P. Reck, J. Chem. Phys. 64, 1270 (1976). sp. E. McNamee, K. Lacmann, and D. R. Herschbach, Faraday Discuss, Chem. Soc. 55, 318 (1973). lS. J. Nalley, R. N. Compton, H. C. Schweinler, and V. E. Anderson, J. Chern. Phys. 59, 4125 (1973). SR. N. Compton, P. W. Reinhardt, and C. D. Cooper, J. Chern. Phys. 63, 3821 (1975). 9K Takeda and F. Williams, Mol. Phys. 17, 677 (1969). 10M. A. Bonin, K. Tsuji, and F. Williams, Nature 218, 946 (1968). 11M. A. Bonin, K. Takeda, K. Tsuji, and F. Williams, Chern. Phys. Lett. 2, 363 (1968). 12See O. H. Crawford, Proc. Phys. Soc. 91, 276 (1967). 13W. R. Garrett, Chern. Phys. Lett. 5, 393 (1970); Phys. Rev. 3A, 961 (1971). 140. H. Crawford, Mol. Phys. 20, 585 (1971). 15M. A. Bonin, Y. J. Chung, E. D. Sparague, K, Takeda, J. T. Wang, and F. Williams, in Nobel Symposium 22, edited by Per-Olof Kinell, Bengt Ranby, and Vera Runnstrom-Reio (Wiley, New York, 1973) p. 103. 1ST Sugiura and A. Arakawa, in Proceedings of the International Conference on Mass Spectrometry, (Univ, of Tokyo, Tokyo, 1970), p. 848. l1J. A. Stockdale, F. J. Davis, R. N. Compton, and C. E. Klits, J. Chem. Phys. 60, 4279 (1974). lSC. E. Klots, J. Chern. Phys. 62, 741 (1975). 19K. G. Mothes, E. Schultes, and R. N. Schindler, J. Phys. Chern. 76, 3758 (1972). It has been found lA. A. Christodoulides, R. Schumacher, and R. N. Schindler, J. Phys. Chern. 79, 1904 (1975)] that the ECR method often overestimated the attachment rates in cases where the rate is small. The attaching moleculse are beleived to react with the walls of the ECR cell resulting in a concentration dependence on tl).e attachment rate. In the case of HCI, for example, the measured rate decreases more than four orders of magnitude as the concentration of HCI increases, A similar dependence is expected for CH3CN (A. A. Christodoulides, private communication) 20M. Matsuzawa, J. Phys. Soc. Jpn. 32, 1080 (1972); 33, 1108 (1972). 21H. Okabe and V. H. DibeIar, J. Chern. Phys. 59, 2430 (1973). 22 J Berkowitz, W. A. Chupka, and T. A. Walter, J. Chern. Phys. 50, 1497 (1969). 23K. D. Jordan and J. J. Wendoloski, Chern. Phys. 21, 145 (1977). 240. H. Crawford and W. R. Garrett, J. Chern. Phys. 66, 4968 (1977). 25W. R. Garrett, J. E, Turner, and V. E. Anderson, Phys. Rev. 188, 513 (1969). 26M. W. Seigel, R. J. Celotta, J. L. Hall. J. Levine, and R. A. Bennett, Phys. Rev. 6, 607 (1972). 21B. A. Thrush and J. J. Zwolenik, Trans. Faraday Soc. 59, 582 (1968).

J. Chern. Phys., Vol. 68, No. 10, 15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Compton, Reinhardt, and Cooper: Electron affinities for CH 3N0 2 , CF31, and CF3Br
28H. Hotop and W. C. Lineberger, J. Phys. Chem. Ref. Data Ser. 4, 539 (1975). 29B. C. Gilbert and Micheal Trenwith, J. Chem. Soc. Perkins Trans. (14), 2010 (1973). 30A. H. Sehon and M. Szwarc, Ann. Rev. Phys. Chem. 8, 439 (1957). 31E. Herbst, T. A. Patterson, and W. C. Lineberger, J, Chem, Phys. 61, 1300 (1974). 32B. deB. Darwent, "Bond Dissociation Energies in Simple Molecules," Natl. Stand. Ref. Data Ser. Natl. Bur. Stand. 31. (1970). 33Akinori Hasegawa and Fframcon Williams, Chem. Phys.

4367

Lett. 46, 66 (1977). D. Sprague and F. Williams, J. Chem. Phys. 54, 5425 (1971); S. P. Mishra and M. C. R. Symons, J. Chem. Soc. Perkin Trans. II, 391 (1973); Y. Fujita, T. Katsu, M. Sato and K. Takahashi, J. Chem. Phys. 61, 4307 (1974). 3ST L. Cottrell, The Strengths of Chemical Bonds (Academic, New York, 1956). 3SG. Herzberg, "Electronic Spectra of Polyatomic Molecules (Van Nostrand, New York, 1966). 37Akinori Hasegawa, Masaru Shiotani and Ffrancon Williams, submitted to Faraday Discuss. on "Radiation Effects in Liquids and Solids."
34 E

J. Chern. Phys., Vol. 68, No. 10, 15 May 1978

Downloaded 13 Feb 2013 to 148.247.22.254. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

You might also like