35
BNL-112077-2016-JA Identification of Ion-Pair Structures in Solution by Vibrational Stark Effects J. Hack Submitted to Journal of Physical Chemistry B February 2016 Chemistry Department Brookhaven National Laboratory U.S. Department of Energy [DOE Office of Science] Notice: This manuscript has been authored by employees of Brookhaven Science Associates, LLC under Contract No. DE- SC0012704 with the U.S. Department of Energy. The publisher by accepting the manuscript for publication acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes.

Identification of Ion-Pair Structures in Solution by

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Identification of Ion-Pair Structures in Solution by

BNL-112077-2016-JA

Identification of Ion-Pair Structures in Solution by Vibrational Stark Effects

J. Hack

Submitted to Journal of Physical Chemistry B

February 2016

Chemistry Department

Brookhaven National Laboratory

U.S. Department of Energy [DOE Office of Science]

Notice: This manuscript has been authored by employees of Brookhaven Science Associates, LLC under Contract No. DE- SC0012704 with the U.S. Department of Energy. The publisher by accepting the manuscript for publication acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes.

Page 2: Identification of Ion-Pair Structures in Solution by

DISCLAIMER

This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, nor any of their contractors, subcontractors, or their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or any third party’s use or the results of such use of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof or its contractors or subcontractors. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

Page 3: Identification of Ion-Pair Structures in Solution by

1

Identification of Ion-Pair Structures in Solution by

Vibrational Stark Effects

John Hack, †‡ David C. Grills,† John R. Miller,† and Tomoyasu Mani†*

†Chemistry Department, Brookhaven National Laboratory, Upton NY 11973-5000, United States

‡Chemical Engineering Department, University of Virginia, 102 Engineers’ Way, PO Box

400741, Charlottesville, VA 22904-4741, United States

ABSTRACT:

Ion pairing is a fundamental consideration in many areas of chemistry and has implications in a

wide range of sciences and technologies that include batteries and organic photovoltaics. Ions in

solution are known to inhabit multiple possible states, including free ions (FI), contact ion pairs

(CIP), and solvent-separated ion pairs (SSIP). However, in solutions of organic radicals and non-

metal electrolytes, it is often difficult to distinguish between these states. In the first part of this

work, we report evidence for the formation of SSIPs in low-polarity solvents and distinct

measurements of CIP, SSIP, and FI, by using the ν(C≡N) infrared (IR) band of a nitrile-

substituted fluorene radical anion. Use of time-resolved IR detection following pulse radiolysis

allowed us to unambiguously assign the peak of the FI. In the presence of non-metal electrolytes,

two distinct red-shifted peaks were observed and assigned to the CIP and SSIP. The assignments

Page 4: Identification of Ion-Pair Structures in Solution by

2

are interpreted in the framework of the vibrational Stark effect (VSE) and are supported by 1) the

solvent-dependence of ion-pair populations, 2) the observation of a cryptand-separated sodium

ion pair that mimics the formation of SSIPs, and 3) electronic structure calculations. In the

second part of this work, we show that a blue-shift of the ν(C≡N) IR band due to the VSE can be

induced in a nitrile-substituted fluorene anion by covalently tethering it to a metal-chelating

ligand that forms an intramolecular ion pair upon reduction and complexation with sodium ion.

This adds support to the conclusion that the shift in IR absorptions by ion pairing originates from

the VSE. These results combined show that we can identify ion-pair structures by using the VSE,

including the existence of SSIPs in a low-polarity solvent.

1. INTRODUCTION

Ion pairing between oppositely charged ions is fundamental to many areas of chemistry

and applications including batteries and organic photovoltaics.1-3 It is widely accepted that ion

pairs exist in solvents having dielectric constants De <30.3 Ion pairing affects both the energetics

and dynamics of ions. Ion pairings of electrolytes and those between organic anion radicals and

metal cations are well studied, but studies on ion pairing between organic radicals and non-metal

electrolytes are scarce despite their significant implications in the control of the energetics of

these molecules, including redox potentials measured by means of electrochemistry. Ion pairing

is particularly important in solvents of low dielectric constant (De <11, here we use the definition

by Van Der Hoeven and Lyklema4 based on Fuoss’s theory5) where Coulomb interactions are

less screened by the solvent.

The exact form of ion pairs in solution is a topic of debate,3 and is often difficult to

identify experimentally. In general, we classify ion pairs into three populations: contact ion pairs

Page 5: Identification of Ion-Pair Structures in Solution by

3

(CIPs) or tight ion pairs, solvent-separated ion pairs (SSIPs) or loose ion pairs, and solvent-

shared ion pairs (SIPs) (Figure 1). We can consider SIP as one case of SSIP with one solvent

molecule between the two ions (therefore shared). For simplicity and due to the currently

achievable experimental resolution, we do not attempt to make any distinction between SIPs and

SSIPs, and call them SSIPs collectively, assuming that there is a single solvent molecule between

ions.

The formation of a SSIP can have significant implications for the kinetics of charge

separation and thermodynamic properties, including electrical conductivity, of a solution6-7 and

photo-induced charge transfer processes in condensed phases.8-9 The concept of a SSIP was first

introduced to explain unusual conductivity and kinetic data by Fuoss, Grunwald, and Winstein in

three independent studies.6-7, 10 The proposed SSIP species rests in a relative potential energy

minimum, separated from the CIP by a solvation energy barrier.

Figure 1. Pictorial representation of free ions (FI), a solvent-separated ion pair (SSIP) or solvent-

shared ion pair (SIP), and a contact ion pair (CIP). Void circle = solvent molecule.

Various studies have been performed to identify and characterize SSIPs, but many are limited to

pairing with metal cations.3 Identifying and distinguishing SSIPs from CIPs and free ions (FIs) in

solution, especially of organic radicals and non-metal electrolytes, is not a trivial task. Until

recently, spectroscopic techniques such as electron paramagnetic resonance, visible, and

vibrational spectroscopy have been able to observe distinct peaks assigned to CIPs and SSIPs,

- -

+

-

+

FI CIPSSIP (or SIP)

+

Page 6: Identification of Ion-Pair Structures in Solution by

4

but unambiguous measurements of FIs in neat solvents, especially in media of low polarity, were

difficult.2, 11-13 This technical limitation left some ambiguity as to the assignment of SSIPs as

distinct thermodynamic states with unique properties and behaviors from FIs. In previous work,

we reported a distinct IR absorption band of the free anion of F1CN, a nitrile-substituted fluorene

(Chart 1), in moderately non-polar neat tetrahydrofuran (THF, De = 7.52),14 using nanosecond

time-resolved infrared (TRIR) detection15 following pulse radiolysis at the Laser-Electron

Accelerator Facility (LEAF) at Brookhaven National Laboratory16. The ν(C≡N) IR peak for the

free ion of F1CN-• (νmax = 2110 cm-1) was red-shifted in comparison to the peak in the neutral

form (νmax = 2224 cm-1). Two further red-shifted peaks were observed upon ion pairing with

tetrabutylammonium ion (TBA+), which we assigned to the CIP (νmax = 2094 cm-1) and the SSIP

(νmax = 2108 cm-1) (Figure 2). We also determined the rate constants for the formation of ion

pairs by monitoring the growth of the ion pair’s IR band.14

Chart 1. Structures of the Nitrile-Substituted Fluorenes Studied.

N

O

OO

OO

N

F1CN

15-crown-5-Ph-F1CN

Page 7: Identification of Ion-Pair Structures in Solution by

5

Figure 2. IR spectra of the ν(C≡N) stretch of F1CN anion in THF. The FI (νmax = 2110 cm-1) was

obtained 50 ns after pulse radiolysis. A further red-shift occurs when F1CN is reduced

electrochemically in the presence of 0.1 M TBA+PF6- or TBA+BF4

-. The ion pair spectrum fits to

two Voigt functions, assigned to the CIP (νmax = 2094 cm-1) and SSIP (νmax = 2108 cm-1). The

data shown in Figure 2 were taken from ref.14

The observed red-shifts in ν(C≡N) IR frequency of the F1CN anion with ion pairing can

be interpreted in the framework of the vibrational Stark effect (VSE).14 This interpretation for

ion pairing between F1CN-• and positive charge was supported by electronic structure

calculations.14 The VSE relates shifts in the vibrational energy to the dot product of the

difference dipole of the vibration (between ground and excited states) and an exerted external

electric field. Vibrational shifts of nitrile and carbonyl probes in neutral molecules have been

calibrated to a known external field in frozen glass and in solution.17-18 These neutral probes can

be used to obtain a quantitative measure of the local electric field in proteins, other

macromolecular complexes, or environments, pioneered by Boxer and colleagues.18-19 Neutral

molecules (benzonitrile)20 and charged species (e.g. thiocyanate)21-22 can be used to investigate

the properties of ionic liquids. In the case of ion pairing between F1CN-• and TBA+, the inert

non-metal ion TBA+ induces an external electric field on the C≡N dipole (the difference dipole

1.0

0.8

0.6

0.4

0.2

0.0N

orm

aliz

ed A

bsor

ban

ce2160214021202100208020602040

Wavenumber (cm-1

)

FIIon Pairs

CIP

SSIP

Page 8: Identification of Ion-Pair Structures in Solution by

6

moment points N→C),23 causing a change in the energy of its vibrational transition (Supporting

Information Figure S1). The observed red-shift in the ion pairs of F1CN-• and TBA+ implies that

a cation is oriented on the nitrile side of F1CN-• in the ion pairs. We assumed that the ion pair

spectrum, shown in Figure 2, includes both CIP (F1CN-•, TBA+) and SSIP (F1CN-•||TBA+) based

on the goodness of fits, but more support is required for a clearer confirmation of the existence of

a SSIP as one of the distinct ion-pair structures and as a thermodynamically stable species.

Within the framework of the VSE, we also hypothesize that we can induce a blue-shift in ν(C≡N)

IR frequency of the nitrile-substituted fluorenes when a cation is positioned on the other side of

the nitrile of anions, away from the CN. Such a blue-shift was previously observed in the cases

of simple salts such as Na+SCN-,24 Li+SCN-,25 Li+SeCN-,26 and inorganic cyanide compounds,27

but the relationships between the VSE and ion-pair structures are often obscured because of the

lack of a priori structural information in such intermolecular complexes in solution.

In the first part of the current paper, we present evidence that supports the existence of

the SSIP in the ion pairs of F1CN anion and TBA+ by 1) the solvent-dependence of the SSIP/CIP

ratio in dimethylformamide (DMF)/THF mixtures, 2) the agreement of peak position in

cryptand-separated sodium ion pair that mimics the formation of a SSIP, and 3) electronic

structure calculations. In the second part of the paper, we report a blue shifted ν(C≡N) IR

vibration due to the VSE in the nitrile-functionalized fluorene radical anion by using a newly

synthesized fluorene derivative (Chart 1) whose intramolecular ion-pair structure is known a

priori. These works combined highlight the strength of the VSE in the studies of inter- and

intramolecular ion pairs. We will discuss ion-pair structures in media of low polarity and their

implications in the energetics of ions in such environments.

Page 9: Identification of Ion-Pair Structures in Solution by

7

2. EXPERIMENTAL METHODS

2.1. General. All reagents and solvents were used as received from standard sources. Silica gel

(pore size 60 Å, 70-230 mesh, Sigma-Aldrich) was used for column chromatography. FTIR

spectra were obtained using a Thermo Nicolet Nexus 670 FTIR spectrometer. UV-Vis-NIR

spectra were obtained using a Cary 5 spectrophotometer (Varian) or Cary 5000

spectrophotometer (Agilent). 1H and 13C NMR spectra were obtained with a Bruker Avance III

spectrometer operating at 400.16 and 100.62 MHz, respectively. Samples were prepared under

an atmosphere of argon for measurements other than NMR spectroscopy.

2.2. Synthesis. Syntheses and characterizations of F1CN,14 BrF1CN,28 and 4-Bromobenzo-15-

crown-529-30 are reported elsewhere.

2.2.1. 9,9-dihexyl-7-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)-9H-fluorene-2-carbonitrile

(BpinF1CN) A dry flask with an argon atmosphere was charged with BrF1CN (0.61 g, 1.39

mmol), Bis(pinacolato)diboron (0.45 g, 1.74 mmol), Pd(dppf)Cl2 (6.0 mol%), and KOAc (0.273

g, 2.78 mmol). After addition of 5 mL of DMF, the reaction mixture was stirred at 80 oC for 5 h,

after which it was cooled down to room temperature and poured into 50 mL of H2O. The mixture

was extracted with DCM (50 mL). The organic layer was collected and washed with brine three

times, and then dried over MgSO4. After filtration and removal of the solvents in vacuo, the

crude mixture was purified by column chromatography on silica gel (hexane:DCM = 7:3) to

afford the title compound as a pale yellow solid (0.40 g, 59%): 1H NMR (CD2Cl2, 400 MHz)

7.81 (m, 4H), 7.65 (m, 2H), 2.03 (m, 4H), 1.36 (s, 12H), 1.10-0.97 (m, 12H), 0.76 (t, J = 7.1 Hz,

6H), 0.51 (m, 4H); 13C NMR (CD2Cl2, 100 MHz) 52.4, 151.0, 145.9, 142.4, 134.2, 131.6,

127.2, 121.0, 120.4, 120.1, 110.8, 84.4, 56.0, 40.4, 31.9, 29.9, 25.1, 24.1, 22.9, 14.1.

Page 10: Identification of Ion-Pair Structures in Solution by

8

2.2.2. 9,9-dihexyl-7-(2,3,5,6,8,9,11,12-octahydrobenzo[b][1,4,7,10,13]pentaoxacyclopentadecin-

15-yl)-9H-fluorene-2-carbonitrile (15-crown-5-Ph-F1CN). The compound BpinF1CN (0.200 g,

412 μmol), 4-bromobenzo-15-crown-5 (0.158 g, 453 μmol), Pd(OAc)2 (11 mg, 49 μmol), SPhos

(41 mg, 99 μmol), and K3PO4 (0.179 g, 824 μmol) were combined in a 10 mL round-bottom

flask. Toluene (2 mL) and H2O (0.1 mL) were added and the resulting mixture was

deoxygenated by a couple of evacuation-refill (Ar) cycles. The mixture was stirred at 80 oC for 2

h and then cooled down and the solvents were removed in vacuo. The crude mixture was purified

by column chromatography on silica gel (5% THF in DCM) to afford the title compound as a

yellow solid (0.183 g, 71%): 1H NMR (CD2Cl2, 400 MHz) 7.81-7.79 (m, 2H), 7.66-7.64 (m,

2H), 7.60 (dd, J = 7.9, 1.5 Hz, 1H), 7.57 (s, 1 7.23 (dd, J = , 8.3, 2.1 Hz, 1H), 7.19 (d, J = 2.1

Hz, 1H), 6.98 (d, J = 8.2 Hz, 1H) 4.22-4.15 (m, 4H), 3.91-3.87 (m, 4H), 3.74-3.67 (m, 10 H),

2.06-2.01 (m, 4H), 1.83-1.80 (m, 2H), 1.13-1.01 (m, 12H), 0.76 (t, J = 7.1 Hz, 6H), 0.61 (m, 4H);

13C NMR (CD2Cl2, 100 MHz)

, 120.3, 114.3, 113.6, 110.1, 71.19. 71.16, 70.7, 60.6, 69.9, 69.8, 69.6, 69.2, 40.5, 31.9,

29.9, 24.1, 22.9, 14.1.

2.3. Spectroelectrochemistry. FTIR spectroelectrochemical measurements were performed in a

spectroelectrochemical transmission cell GS20900 (Specac Limited), coupled with a CHI 600E

Potentiostat (CH Instruments, Inc.). All samples contained 0.1 M TBA+PF6-.

2.4. FTIR of Reduced Species. Chemical titrations were performed with sodium biphenyl

(Na+BP-) or sodium chunks, either in the absence or presence of cryptand-222 (C222), in a

glovebox. The concentration of C222 was at least one equivalent of neutral molecules before

Page 11: Identification of Ion-Pair Structures in Solution by

9

reduction unless otherwise noted. The solutions were transferred into a demountable liquid flow

cell DLC-S25 (Harrick Scientific) equipped with 2 mm thick CaF2 windows and a 950 μm thick

spacer.

2.5. Pulse Radiolysis with Time-resolved Infrared (TRIR) Detection. Pulse radiolysis

experiments were conducted in LEAF at Brookhaven National Laboratory. A detailed

description of the experimental setup is given elsewhere.15 An airtight IR solution flow cell was

used (1.1 mm pathlength), equipped with 0.35 mm thick CaF2 windows. Continuous wave

external-cavity quantum cascade lasers (models 21047-MHF and 21043-MHF, Daylight

Solutions, Inc.) were used as the IR probe source. Samples (5-20 mM) were dissolved in THF or

DMF and purged with argon.

2.6. Quantum Calculations. Computations were performed with either Gaussian0931 or Q-

Chem 4.232 using density functional theory (DFT). All calculations on anions were spin-

unrestricted. All hexyl groups at the 9-position of fluorenes were replaced by ethyl groups. The

geometry optimizations were performed without symmetry constraints. Calculations used the 6-

31G(d) basis set and the LC-ωPBE (ω = 0.1 bohr-1)28, 30, 33 or B3LYP34-35 functional. A PCM

solvation model36-37 (Gaussian) or C-PCM solvation model38-40 (Q-Chem) for THF was used.

3. RESULTS AND DISCUSSION

3.1. SSIP vs. CIP. 3.1.1. Ion Pair Populations in DMF/THF Mixtures In the first part of this

work, we examine the existence of a SSIP in the ion pairs of F1CN-• and TBA+ in solvents of low

polarity. In THF solution containing 0.1 M TBA+PF6-, which is a typical total concentration of

electrolytes in electrochemical measurements, the results indicated that >95% of the ions exist as

ion pairs based on fits such as those shown in Figure S2. We determined that the spectrum of the

Page 12: Identification of Ion-Pair Structures in Solution by

10

SSIP is distinct from that of the FI. To further support the assignment of the SSIP peak (Figure

2), IR spectra in the region of the ν(C≡N) stretch were measured spectroelectrochemically in

mixtures of DMF (De = 38.25) and THF containing 0.1 M TBA+PF6- (Figure 3). The solvent

properties of the mixtures are reported in Supporting Information Table S1. The spectra obtained

in mixtures up to 20% DMF by volume fit to two Voigt functions centered at 2094 cm-1,

assigned to the CIP, and 2108 cm-1, assigned to the SSIP, with a SSIP/CIP ratio that increases

with DMF composition. The characteristics of their spectra are listed in Table 1. The two Voigt

functions were constrained to the same position, full width at half maximum (FWHM), and

shape as the functions used to fit the spectrum of ion pairs in pure THF solvent. Broader

linewidths observed for the CIP and the SSIP bands in the presence of TBA+ compared to that

for the FI indicate that multiple conformations exist in both the CIP and the SSIP. However, as

mentioned above, in this work we do not attempt to distinguish these conformers. The fittings of

the spectra at each concentration are reported in Supporting Information Figure S2.

Figure 3. Spectroelectrochemical measurements of the ν(C≡N) stretch of F1CN-• in mixtures of

DMF and THF containing 0.1 M TBA+PF6-. The spectra obtained in mixtures up to 20% DMF

CIP SSIP

Abs

orb

an

ce (

a.u

.)

21402120210020802060

Wavenumbers (cm-1

)

100%

70%

50%30%20%

15%10%

5%

0%

% DMF

Page 13: Identification of Ion-Pair Structures in Solution by

11

by volume fit well to two Voigt functions centered at 2094 cm-1, assigned to the CIP, and 2108

cm-1, assigned to the SSIP, with a SSIP/CIP ratio that increases with DMF composition.

Beyond 20% DMF, the peaks shift noticeably and no longer fit cleanly to the same two

Voigt functions. As DMF becomes a larger fraction of the bulk solvent, several confounding

effects may occur. As the dielectric constant of the bulk solvent increases the ν(C≡N) peak

maximum can shift to the red probably due to solvent-induced VSE,17-18 as seen for F1CN-• free

ions, νmax = 2101 cm-1 in DMF, vs. νmax = 2110 cm-1 in THF (the 9 cm-1 red-shift can be seen in

Supporting Information Figure S3). As the dielectric constant of the solvent increases, the

Coulomb attractions between oppositely charged species are more screened. The free ion

becomes more stabilized, and may begin to contribute to the spectra. The observed change in FI

from THF to DMF may be explained by a simple Onsager-like model of solvation as in the case

of neutral aromatic nitriles, but more local solute-solvent interactions41 are likely to play an

important role in anions. Interestingly, the spectrum reported in 100% DMF solvent containing

0.1 M electrolyte is centered, except for a slight electrolyte-induced red-shift, at almost the same

position, 2101 cm-1, as the free ion peak in neat DMF, observed by TRIR detection following

pulse radiolysis (Supporting Information Figure S4a). We can attribute this red-shift to the

presence of ion pairs. We can fit the spectra to two functions by fixing one function to the

spectrum of the free ion in DMF while adjusting the other, assumed to arise from ion pairs

(Supporting Information Figure S4b). Here, we did not make any distinctions among ion pairs,

and we find adequate, though not excellent fits. Assuming that the ratio of the areas under the

curves reflects the population ratio, we can estimate the association constant for ion pairing

between F1CN-• and TBA+ in DMF to be KA ~ 3 M-1, which is about four orders of magnitude

smaller than that in THF (KA = 9.2 × 104 M-1).14 The value KA ~ 3 M-1 is one order of magnitude

Page 14: Identification of Ion-Pair Structures in Solution by

12

smaller than that for TBA+PF6- in acetonitrile reported by Geiger and coworkers (KA = 35.7 M-

1),42-43, which seems reasonable given that KA(F1CN-•,TBA+) is about four times smaller than

KA(TBA+,PF6

-) in THF (see Supporting Information Table S2). An alternative is to interpret a

slight broadening in the redder side in the spectroelectrochemical peak in terms of changes in the

dielectric constants due to the presence of 0.1 M electrolyte. The first interpretation which

identifies a reasonable KA(F1CN-•,TBA+) seems preferable. In both cases, the observation

supports the picture that a radical anion is mainly “free” in a polar solvent in the presence of

electrolytes, while some may be stabilized by ion pairings. This is different from the picture we

previously put forward that ions are mainly stabilized by a single ion pair in non-polar

environments.14

Table 1. Observed ν(C≡N) (cm-1) of Anion Species in THFa

Molecules FIb

Ion Pairs

TBA+ Na+

CIPc SSIPc Na+d Na(C222)+d

F1CN 2110

(12.1)

2094

(21.5)

2108

(18.8)

2094

(22.8)

2108e

(15.9)

2106f

(20.0)

15-crown-5-Ph-F1CN 2124

(18.4)

2106

(23.6)

2122

(26.3)

2136

(23.1)

2122

(24.0)

aThe numbers in parentheses are the fwhm values determined by fitting with a Voigt function. Uncertanities are ± 1 cm-1. bObtained by pulse radiolysis. cObtained by spectroelectrochemistry. dObtained by chemical reductions. eFitted to two Voigt functions. The reported value in the Table is assigned to the species that “mimics” the SSIP. The other peak was observed at 2097 cm-1. fFitted to one Voigt function.

Page 15: Identification of Ion-Pair Structures in Solution by

13

In order to calculate association constants for the CIP and SSIPs, the SSIP/CIP ratios

obtained in mixtures up to 20% DMF were fit to the model represented in Scheme 1. This model

allows for the formation of a CIP, a SSIP separated by one THF molecule, and a SSIP separated

by one DMF molecule. We abbreviate these as SSIPTHF and SSIPDMF. The SSIP peak in Figure 3

is taken to be a combination of contributions from both SSIP species. The model makes the

following assumptions. First, we assume that there is only one solvent molecule separating ions

in the SSIP species, though it is possible that there are multiple.3 Second, we assume that in

mixtures up to 20% DMF we can neglect formation of free ions and assume that DMF does not

preferentially solvate either type of ion pair. Third, we assume that the ratio of the extinction

coefficients of ν(C≡N) absorption in CIP and SSIP is unity. The second assumption is supported

by the absence of a clear peak corresponding to that of the FI in DMF. The last assumption is

supported by DFT calculations, which show that the extinction coefficients of ν(C≡N) absorption

of anions differ by only 1-1.5 times over a distance of 4-5 Å between F1CN-• and a positive

charge.14 The maximum difference in the distance between SSIP and CIP was reported to likely

be ~4-5 Å.8-9 This difference is partially confirmed by computations (Supporting Information

Figure S5). Importantly, using a ratio of 1.5 does not yield significantly different results from

those reported (Supporting Information Table S3). A similar approach estimating extinction

coefficients of IR absorbance using the outputs from electronic structure calculations was

recently adopted by Ludwig and coworkers for far-IR absorption bands.12-13

Scheme 1. Ion Pairing Equilibria in DMF/THF Mixtures

TBA+ + PF6- ⇌ (TBA+, PF6

-) (1)

F1CN-• + TBA+ ⇌ CIP (F1CN-•, TBA+) (2)

Page 16: Identification of Ion-Pair Structures in Solution by

14

F1CN-• + THF + TBA+ ⇌ SSIPTHF (F1CN-•, THF, TBA+) (3)

F1CN-• + DMF + TBA+ ⇌ SSIPDMF (F1CN-•, DMF, TBA+) (4)

Calculated association constants for the formation of CIP and SSIPs are reported in Table

2. The association constants KA for TBA+PF6- were taken to be mole-fraction weighted linear

combinations of association constants determined by Geiger and coworkers42 in THF (3.73 x 105

M-1), and DMF (35.7 M-1),43 and not adjusted. The association constant for the formation of

SSIPDMF, K4, is about an order of magnitude larger than for SSIPTHF, K3. This result matches the

expectation that DMF, a more polar solvent than THF, has a higher association constant for SSIP

formation, and explains the observed growth in the combined SSIP peak with increasing DMF

composition. Standard free energy changes (ΔG0) are reported in Table 2 as well. K3 was

multiplied by the concentration of THF so that the energy of the SSIPTHF could be compared to

the energy of the CIP. It was found that the CIP (ΔG0 = -300 ± 5 meV) is stabilized by 30 meV

compared to the SSIPTHF (ΔG0 = -270 ± 10 meV). This value is not much larger than the average

thermal energy, kBT at room temperature.

Table 2. Ion Pairing Equilibria for F1CN-• with TBA+

Equationa KAb ΔG0 (meV)b

2 K2 = (9.8 ± 1.8) × 104 M-1 -300 ± 5

3 K3 = (2.9 ± 1.1) × 103 M-2 -200 ± 10

3 K3 [THF] = (3.4 ± 1.3) × 104 M-1 -270 ± 10c

4 K4 = (2.6 ± 1.1) × 104 M-2 -260 ± 10

a Equations as defined in Scheme 1. b The values are the averages from five different DMF/THF ratios. c100% THF ([THF] = 12.3 M).

Page 17: Identification of Ion-Pair Structures in Solution by

15

3.1.2. Macrocycle Polyether-Separated Intermolecular Ion Pair.

The FTIR spectrum of F1CN-• in THF obtained by chemical reduction of F1CN by

sodium biphenyl is shown in Figure 4. The ν(C≡N) absorption band of the CIP (F1CN-•, Na+)

appears at 2094 cm-1, within one wavenumber of that of the CIP with TBA+ in THF (Figure 2

and Supporting Information Figure S6). Here we assign the peak to CIP based on the previous

studies by Szwarc and coworkers showing that Na+ preferentially forms a CIP with anions in

THF except for bulky anions (e.g. tetraphenylboride).1, 44 This agreement, despite the large

difference in ionic radii of Na+ and TBA+ (crystal ionic radii for Na+ and TBA+ are 0.95 and 4.13

Å, respectively),1, 45-46 may support the theoretical model presented by Fry47-48 that the alkyl

chains in TBA+ contribute negligible steric hindrance to ion pairing with anions. This model is

similar in nature to the concept of penetrated ion pairs.49-51

Figure 4. FTIR spectra obtained for F1CN anions in THF. The red spectrum, peaking at 2094

cm-1, labeled F1CN-•, Na+, was produced by a chemical reduction with sodium biphenyl. The

green spectrum, labeled F1CN-•||Na(C222)+,was produced by reduction with excess sodium

biphenyl in the presence of cryptand-222 (C222). A main peak is observed at 2108 cm-1 with an

Page 18: Identification of Ion-Pair Structures in Solution by

16

additional shoulder at 2097 cm-1. C222 encapsulates the Na+ ion, creating a “separated” ion pair,

which mimics the SSIP.52-54 The FI spectrum, labeled F1CN-•, (νmax = 2110 cm-1) was recorded

with TRIR detection obtained 50 ns after pulse radiolysis in neat THF. The balls in the top of the

figure represent Na+.

In the presence of cryptand-222 (C222), a main peak is observed at 2108 cm-1 together

with a shoulder at 2097 cm-1, the former of which is within one wavenumber of the SSIP

observed with TBA+. C222, a chelating agent, encapsulates the Na+ ion, creating a “separated”

ion pair with F1CN-• as F1CN-•||Na(C222)+, which can mimic the SSIP (F1CN-•||TBA+). This

strategy prevents close contact between alkali metals and anions, as shown elegantly by Kochi

and coworkers.52-54 The notable agreement between the observed peak and the peak at 2108 cm-1

in the presence of TBA+ further supports the assignment of the SSIPTHF (F1CN-•, THF, TBA+)

observed in Figures 2 and 4. The shoulder observed at 2097 cm-1 (FWHM = 15.8 cm-1 from

multi-peak curve fitting) is relatively close to that of the CIP observed with TBA+ and Na+ (νmax

= 2094 cm-1). This may indicate either an incomplete encapsulation of Na+ by C222 or the

possibility that some Na(C222)+ can be positioned close enough to resemble the CIP. Changing

the concentration of C222 (from one equivalent of neutral F1CN to three equivalents) appears

not to change the ratio of the two peaks, which suggests the latter is the case. It is important to

note that the observed spectrum of “separated” ion pair with Na(C222)+ is clearly different from

that of the FI, showing that while the cryptands can reduce the effects of ion pairing they will not

create FIs in media of low polarity.

Combining the solvent dependence and chemical reduction data clearly shows that we

can distinguish the SSIP from the FI and in turn supports our initial assumption that anions in

THF containing 0.1 M TBA+PF6- primarily exist as ion pairs. While it is expected from the

Page 19: Identification of Ion-Pair Structures in Solution by

17

thermodynamic equilibria (KA = 9.2 × 104 M-1 for ion pairs as a whole14) that 2% of F1CN-• will

be present as FIs in such a solution, our IR measurements cannot readily detect this small

population. Analysis indicates that it would likely be possible to detect if at least 5% FIs are

present.

3.1.3. Energetics of Ion Pairs. We now turn our attention to the effect of ion parings on

the energetics of the ions, obtaining estimates of the standard reduction potential (E0) in the

absence of electrolytes and in the presence of electrolytes (typically 0.1 M) determined by cyclic

voltammetry. The standard energetics of the SSIPTHF and CIP in THF (vs. Fc+/0) are shown in

Figure 5, including the previously determined energies of the FI and Ion Pairs.14 We further

compare the experimental values with estimates by electronic structure calculations. The typical

calculated structures of the SSIPTHF and the CIP are displayed in Supporting Information Figure

S5.

Figure 5. Standard energy diagram in THF (E0 vs. Fc+/0 in eV) of the Ion Pairs, CIP, SSIP, and

FI of F1CN-• in THF. CIP = (F1CN-•, TBA+). SSIP = (F1CN-•, THF, TBA+). FI = F1CN-•. The

energy of the Ion Pairs is taken from ref28 and the energy of the FI was determined using

FI-3.01

CIP-2.71

∆G0IP = -290 meV

SSIP-2.74

∆ ∆ G0

= -30 meV

Ion Pairs-2.72

Page 20: Identification of Ion-Pair Structures in Solution by

18

equation 5 with the value of KA from ref14. Note in ref14 the reduction potential was measured

with 0.1 M TBA+BF4-, which is -2.74 V vs. Fc+/0. Although the difference is only 20 meV, for

clarity, we used the reduction potential determined under the same conditions as those used in

the current study (0.1 M TBA+PF6-). The energy of the FI shifts by 20 meV accordingly while KA

does not change. The energy spacing in the graph is arbitrary.

The SSIPTHF was calculated as a F1CN-•, TBA+ pair surrounded by four explicit THF molecules.

We found that the TBA+ ion is stabilized by approximately four THF molecules (Supporting

Information Figure S7) in a similar manner to DMF and DMSO.55 The CIP was calculated by

removing one THF molecule and allowing F1CN-• to move closer to the cation. The distances

between the nitrogen atoms in F1CN-• and TBA+ (dNN) are ~3.8 Å for the CIP and ~7.8 Å for the

SSIP. We did not fully optimize the structures, but selected the structures, obtained in the course

of the optimization processes, which are consistent with our IR observations and interpretation

by VSE; namely the configuration in which TBA+ is positioned on the same side as the nitrile of

F1CN-•. (See Supporting Information pp. 2-3 for the details). Our current result of intramolecular

ion pairs further supports that the IR shift is induced by the VSE (vide infra). The CIP was

calculated to be more stable than the SSIPTHF by a free energy difference of approximately 40

meV. This value is in a good agreement with the experimentally determined value (~30 meV).

However, we would like to note that the structures used here are among the many stable

conformations and more samplings of conformations of ion pairs would be required to further

validate these preliminary computational results.

In a solution of single solvent where the concentration of solvent is constant, the

SSIP/CIP ratio should not change (eqs 2 and 3). We previously demonstrated that the reduction

potentials in the absence of electrolytes can be determined by measuring the association

Page 21: Identification of Ion-Pair Structures in Solution by

19

constants for ion pairing.14 In the previous analysis,14 we did not distinguish SSIP/CIP

contributions, instead treating ion pairing as a mixture of the two. Here, we could determine the

respective standard energies of the CIP and SSIP within a couple of assumptions. The estimated

energy difference ~kBT makes it difficult to measure the contribution of individual species by

cyclic voltammetry measurements whose experimental uncertainties are typically about 10-20

mV.56 While it would be desirable to measure the difference by slowly scanning the potential, we

do not expect to alter the ratio of CIP and SSIP peaks by adjusting potentials if the two forms are

in equilibrium with each other for each ion pair. While the SSIP/CIP ratios are likely to depend

on the molecules of interest and electrolytes of choice, based on our current study, we can safely

say that both species are present in ion pairs between other molecules and electrolytes in a

similar environment. Therefore, the standard reduction potentials measured in the presence of

counter-ions M+ from electrolytes by electrochemical methods (E0([M+])) should be considered

to reflect the contributions from both species. In general, E0([M+]) is related to the standard

reduction potential in the absence of electrolytes (E([M+] = 0)) and to the association constants

(KA) of ion pairing between redox-active species and counter-ions from electrolytes through the

following equation:57-58

M M 0 ln 1 M (5)

where R is the gas constant, T is the temperature, F is the Faraday constant, and [M+] is the

concentration of free counter-ions M+ from the electrolyte. The energy diagram in Figure 5 and

our previous report14 used a standard condition for all solutes including the electrolytes, giving

E0([M+] = [M+]0 = 1M). We now consider the case where CIP and SSIP are present. Assuming

that only free ions contribute to the electron transfer reactions at the electrode, we obtain the

following equation:

Page 22: Identification of Ion-Pair Structures in Solution by

20

M M 0 ln 1 M Solv M (6)

where KCIP and KSSIP are defined in Supporting Information pp. 4-5. This shows that the standard

energy of ion pairs E0([M+]), obtained by cyclic voltammetry, is affected by the additional

contribution of stabilization by SSIPs. In the current case, the standard energies of the CIP and

ion pairs obtained by cyclic voltammetry are within the experimental errors (± 10 meV) and we

are unfortunately not able to clearly observe the contributions from the SSIP in the standard

redox potentials measured electrochemically.

3.1.4. Differences in Standard and Experimental Conditions. As mentioned above, in our

previous report,14 we defined the reduction potential in the presence of electrolytes at the

standard condition, meaning that we have E0([M+]0), where usually [M+]0 = 1M and use E1/2

from the measurement of cyclic voltammetry as E0([M+]0) referenced to the ferrocene couple (vs.

Fc+/0).59 This convention allows us to use equation 5 in the simplest form, making it easy to

estimate and compare the redox potentials in the absence of electrolytes within the assumption

that the major stabilization of redox-active ions by electrolytes comes from single ion pairs.

However, the experimental concentration of free counter-ions from 0.1 M of typical electrolytes

is much lower in media of low polarity (e.g. [M+] = [TBA+] ~ 500 μM for TBA+PF6- in THF). In

this regard, it may be more appropriate to define the E1/2 measured by cyclic voltammetry as the

reduction potential in the presence of electrolytes at the specific experimental condition. Here the

concentration of free counter-ions in equation 5 is no longer [M+]0 and becomes the value at the

experimental condition in specific media with specific electrolytes. Based on this definition, we

obtained the contribution of ion pairs as ~100 meV for the condition of 0.1 M TBA+PF6- in THF.

The reduction potential of F1CN in the absence of electrolytes is then calculated to be -2.82 V vs.

Fc+/0 where the oxidation potential of ferrocene is defined under the same conditions. This value

Page 23: Identification of Ion-Pair Structures in Solution by

21

is ~ 190 mV more positive than the energy estimated under the standard conditions (Figure 5).

We would like to emphasize that this difference is solely from the differences in the definition,

either at standard or specific experimental conditions for electrolytes, and it still holds that the

measurements of association constants between redox-active species and counter-ions from

electrolytes allow us to estimate the redox potentials in the absence of electrolytes, assuming that

only ion pairing shifts the potentials.

3.2. Blue-shift VSE. 3.2.1. Molecular Design. In the second part of the study, we examine

the possibility of inducing a blue-shift by VSE. Recall that a positively charged TBA+ sits on the

same side of the molecule as the nitrile group in F1CN-•, inducing a red-shift in ν(C≡N). In the

framework of the VSE a positive charge positioned on the opposite side, away from the nitrile

group, might produce a blue-shift in ν(C≡N). Such a blue-shift was previously observed in

simple inorganic slats.25-27 The blue-shift was interpreted based in the framework of the VSE:

The difference dipole moment of ν(C≡N) is aligned antiparallel to the external electric field

induced by the metal cation and cyanide-containing anions. However, the relationships between

the VSE and ion-pair structures are often obscured because of the lack of a priori structural

information in these complexes in solution.

Here, to obtain a clearer relationship between the VSE and ion-pair structures and to

further clarify the origin of IR shifts, we synthesized and studied a new molecular system whose

ion-pair structure is better defined through covalent bonding. The new molecule is named 15-

crown-5-Ph-F1CN where a benzo crown ether (15-crown-5-Ph) is tethered to the phenyl on the

opposite side, away from the nitrile group (Chart 1). The molecule can take either syn or anti

conformations (Supporting Information Figure S8). DFT calculations predict that the syn

conformer is only slightly (< 1 meV) more stable in the neutral molecule and that there is a ~ 150

Page 24: Identification of Ion-Pair Structures in Solution by

22

meV torsional energy barrier between the syn and anti conformers. For the anion, the syn

conformers are again only slightly (~3-4 meV) more stable and the syn-anti torsional barrier is

higher, ~ 200-250 meV (Supporting Information Figure S9). Higher energy barriers are

consistent with a similar torsional motion previously reported in structurally related

oligofluorenes.60-61 We did not determine experimentally which conformer is dominant, but

based on the computational results, we depict the molecular structures as syn conformers here.

Figure 6. IR spectra of F1CN-• anions in THF. The blue line represents the FTIR spectrum of the

species obtained by chemical reduction with excess sodium biphenyl, yielding the species

Na(15-crown-5)+-Ph-F1CN (νmax = 2136 cm-1). The FI (black, νmax = 2124 cm-1) and ion pair

(red) spectra were taken with TRIR detection obtained 60 and 100 ns after pulse radiolysis in

neat THF and in the presence of 0.1 M TBA+PF6-, respectively. The ball in the top of the figure

represents Na+.

3.2.2. Intramolecular Ion Pair. Crown ethers are known to encapsulate alkali metal ions,62-

63 and are widely used as sensors for metal ions in materials and biological sciences.64 They were

also employed in the studies of intermolecular ion pairs.52-54 When 15-crown-5-Ph-F1CN is

reduced with sodium or sodium biphenyl in the absence of any other chelating agents, Na+ is

Page 25: Identification of Ion-Pair Structures in Solution by

23

expected to be captured by the crown ether within the molecule while the excess electron is

delocalized in the –conjugated part. This ensures that a positive charge is on the opposite side

of the molecule, away from the nitrile group. The complexation forms the neutral spin ½ species

Na(15-crown-5)+-Ph-F1CN-• that we can classify as an intramolecular ion pair (Figure 6 Top).

The nitrile mid-IR peak from the FI of the molecule, (15-crown-5)-Ph-F1CN-•, is shifted

to higher frequency by ~14 cm-1 (νmax = 2124 cm-1) compared to that of F1CN-• due to the further

delocalization of an electron in the –conjugated part of the molecule.28 Upon reduction and

complexation with Na+, the IR absorption from Na(15-crown-5)+-Ph-F1CN-• shifts further blue

by another ~12 cm-1 (νmax = 2136 cm-1). DFT calculations support our experimental observations

(Supporting Information Table S4). Computations predict that the anti conformer exhibits a

larger shift than the syn conformer. When electrochemically reduced in the presence of 0.1 M

TBA+PF6-, it forms ion pairs with TBA+, which cannot fit into the crown, and its IR absorption

band shifts to a lower frequency compared to that of the FI (νmax = 2110 cm-1) as expected. The

spectrum can be decomposed into CIP (15-crown-5)-Ph-F1CN-•, TBA+) and SSIP ((15-crown-

5)-Ph-F1CN-•, THF, TBA+) contributions in a similar manner to the case of F1CN-•. Their

spectroscopic characteristics are summarized in Table 1. In contrast to the IR spectra, the optical

spectra of anions of this molecule can distinguish between the FI and the intramolecular ion pair,

but not clearly between the FI and the intermolecular ion pairs (Supporting Information Figure

S10). We note that a slightly larger shift in the IR peak from the FI was observed for the CIP

with TBA+ (Δ ~-18 cm-1) than for the intramolecular ion pair with Na+ (Δ ~12 cm-1) in ((15-

crown-5)+-Ph-F1CN-•. This result further supports the VSE interpretation of the observed IR

peak shifts upon ion pairings and points to the possibility of calibrating these shifts to obtain

Page 26: Identification of Ion-Pair Structures in Solution by

24

what has been called a Stark “tuning rate”,53 meaning the sensitivity of the vibrational frequency

to an external electric field, of ν(C≡N) in anions by careful molecular design.

4. CONCLUSIONS

Ion pairing can control the energetics and dynamics of ions in various environments. In

the first part of the study, we presented evidence for the formation of a SSIP between organic

radicals and non-metal electrolytes in moderately non-polar solvent THF, and distinct spectral

measurements of the CIP, SSIP, and FI using the ν(C≡N) IR absorption in a nitrile-substituted

fluorene. We support our assignment of the SSIP peak with 1) shifting ion pair populations in

DMF/THF mixtures, 2) agreement with the peak position in C222-separated sodium ion pair, and

3) electronic structure calculations. We report the association constants for SSIPs formed by

F1CN-• and TBA+ separated by either a THF or DMF molecule in DMF/THF mixtures of low

DMF composition. Our findings suggest that SSIPs can be formed and exist as

thermodynamically stable species in non-polar environments. In our system, CIPs are slightly

more stable than SSIPs. The existence of the SSIP implies that the redox potentials measured by

electrochemistry in such environments should be considered to reflect the contributions from

both species. In the second part of the study, we show that we can induce a blue-shift in the

ν(C≡N) IR band of a nitrile-substituted fluorene anion by covalently tethering it to a metal-

chelating ligand. This adds further support that the induced shift in IR absorptions by ion pairing

originates from the vibrational Stark effect. These results combined show the capability of

identifying ion-pair structures based on vibrational information.

Page 27: Identification of Ion-Pair Structures in Solution by

25

ASSOCIATED CONTENT

Supporting Information. Computational details, detailed description of contribution of ion

pairings in redox potentials, Figures S1-S10, and Tables S1-4. This material is available free of

charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION

Corresponding Author

*[email protected]

ACKNOWLEDGMENT

This material, and use of the LEAF facility of the BNL Accelerator Center for Energy

Research and the computer Cluster at the Center for Functional Nanomaterials, is supported by

the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Division of

Chemical Sciences, Geosciences & Biosciences under contract no. DE-SC0012704. This project

was also supported in part by the U.S. Department of Energy, Office of Science, Office of

Workforce Development for Teachers and Scientists (WDTS) under the Science Undergraduate

Laboratory Internships Program (SULI). T.M. is grateful for the support by the Goldhaber

Distinguished Fellowship from Brookhaven Science Associates. The authors thank Bobby H.

Layne for technical assistance in LEAF experiments and Dr. Qin Wu for his assistance in

quantum calculations.

REFERENCES

1. Szwarc, M., Ions and Ion Pairs. Accounts Chem. Res. 1969, 2, 87-96.

Page 28: Identification of Ion-Pair Structures in Solution by

26

2. Smid, J., Structure of Ion-Pair Solvation Complexes. Angew. Chem. Int. Edit. 1972, 11,

112-127.

3. Marcus, Y.; Hefter, G., Ion Pairing. Chem. Rev. 2006, 106, 4585-4621.

4. Van Der Hoeven, P. H. C.; Lyklema, J., Electrostatic Stabilization in Nonaqueous Media.

Adv. Colloid. Interfac. 1992, 42, 205-277.

5. Fuoss, R. M., Ionic Association. III. The Equilibrium between Ion Pairs and Free Ions. J.

Am. Chem. Soc. 1958, 80, 5059-5061.

6. Winstein, S.; Clippinger, E.; Fainberg, A. H.; Heck, R.; Robinson, G. C., Salt Effects and

Ion Pairs in Solvolysis and Related Reactions. III. Common Ion Rate Depression and Exchange

of Anions during Acetolysis. J. Am. Chem. Soc. 1956, 78, 328-335.

7. Sadek, H.; Fuoss, R. M., Electrolyte Solvent Interaction. IV. Tetrabutylammonium

Bromide in Methanol Carbon Tetrachloride and Methanol Heptane Mixtures. J. Am. Chem. Soc.

1954, 76, 5897-5901.

8. Mentel, K. K.; Nunes, R. M. D.; Serpa, C.; Arnaut, L. G., Dynamics of Radical Ion Pairs

Following Photoinduced Electron Transfer in Solvents with Low and Intermediate Polarities. J.

Phys. Chem. B 2015, 119, 7571-7578.

9. Vauthey, E., Investigations of Bimolecular Photoinduced Electron Transfer Reactions in

Polar Solvents Using Ultrafast Spectroscopy. J. Photoch. Photobio. A 2006, 179, 1-12.

10. Grunwald, E., Interpretation of Data Obtained in Nonaqueous Media. Anal. Chem. 1954,

26, 1696-1701.

11. Hogen-Esch, T. E.; Smid, J., Studies of Contact and Solvent-Separated Ion Pairs of

Carbanions. I. Effect of Temperture Counterion and Solvent. J. Am. Chem. Soc. 1966, 88, 307-

318.

Page 29: Identification of Ion-Pair Structures in Solution by

27

12. Fumino, K.; Stange, P.; Fossog, V.; Hempelmann, R.; Ludwig, R., Equilibrium of

Contact and Solvent-Separated Ion Pairs in Mixtures of Protic Ionic Liquids and Molecular

Solvents Controlled by Polarity. Angew. Chem. Int. Edit. 2013, 52, 12439-12442.

13. Fumino, K.; Fossog, V.; Stange, P.; Wittler, K.; Polet, W.; Hempelmann, R.; Ludwig, R.,

Ion Pairing in Protic Ionic Liquids Probed by Far-Infrared Spectroscopy: Effects of Solvent

Polarity and Temperature. Chemphyschem 2014, 15, 2604-2609.

14. Mani, T.; Grills, D. C.; Miller, J. R., Vibrational Stark Effects to Identify Ion Pairing and

Determine Reduction Potentials in Electrolyte-Free Environments. J. Am. Chem. Soc. 2015, 137,

1136-1140.

15. Grills, D. C.; Farrington, J. A.; Layne, B. H.; Preses, J. M.; Bernstein, H. J.; Wishart, J.

F., Development of Nanosecond Time-Resolved Infrared Detection at the LEAF Pulse

Radiolysis Facility. Rev. Sci. Instrum. 2015, 86, 044102 1-10.

16. Wishart, J. F.; Cook, A. R.; Miller, J. R., The LEAF Picosecond Pulse Radiolysis Facility

at Brookhaven National Laboratory. Rev. Sci. Instrum. 2004, 75, 4359-4366.

17. Boxer, S. G., Stark Realities. J. Phys. Chem. B 2009, 113, 2972-2983.

18. Fried, S. D.; Boxer, S. G., Measuring Electric Fields and Noncovalent Interactions Using

the Vibrational Stark Effect. Accounts Chem. Res. 2015, 48, 998-1006.

19. Ritchie, A. W.; Webb, L. J., Understanding and Manipulating Electrostatic Fields at the

Protein–Protein Interface Using Vibrational Spectroscopy and Continuum Electrostatics

Calculations. J. Phys. Chem. B 2015, 119, 13945-13957.

20. Zhang, S. G.; Zhang, Y.; Ma, X. Y.; Lu, L. J.; He, Y. D.; Deng, Y. Q., Benzonitrile as a

Probe of Local Environment in Ionic Liquids. J. Phys. Chem. B 2013, 117, 2764-2772.

Page 30: Identification of Ion-Pair Structures in Solution by

28

21. Ren, Z.; Brinzer, T.; Dutta, S.; Garrett-Roe, S., Thiocyanate as a Local Probe of Ultrafast

Structure and Dynamics in Imidazolium-Based Ionic Liquids: Water-Induced Heterogeneity and

Cation-Induced Ion Pairing. J. Phys. Chem. B 2015, 119, 4699-4712.

22. Ren, Z.; Ivanova, A. S.; Couchot-Vore, D.; Garrett-Roe, S., Ultrafast Structure and

Dynamics in Ionic Liquids: 2D-IR Spectroscopy Probes the Molecular Origin of Viscosity. J.

Phys. Chem. Lett. 2014, 5, 1541-1546.

23. Andrews, S. S.; Boxer, S. G., Vibrational Stark Effects of Nitriles I. Methods and

Experimental Results. J. Phys. Chem. A 2000, 104, 11853-11863.

24. Baiz, C. R.; Kubarych, K. J., Ultrafast Vibrational Stark-Effect Spectroscopy: Exploring

Charge-Transfer Reactions by Directly Monitoring the Solvation Shell Response. J. Am. Chem.

Soc. 2010, 132, 12784-12785.

25. Lee, K. K.; Park, K. H.; Kwon, D.; Choi, J. H.; Son, H.; Park, S.; Cho, M., Ion-Pairing

Dynamics of Li+ and SCN- in Dimethylformamide Solution: Chemical Exchange Two-

Dimensional Infrared Spectroscopy. J. Chem. Phys. 2011, 134, 064506 1-10.

26. Park, K. H.; Choi, S. R.; Choi, J. H.; Park, S.; Cho, M., Real-Time Probing of Ion Pairing

Dynamics with 2DIR Spectroscopy. Chemphyschem 2010, 11, 3632-3637.

27. Gill, J. B., Solute-Solute Interactions in Liquid-Ammonia Solutions - a Vibrational

Spectroscopic View. Pure Appl. Chem. 1981, 53, 1365-1381.

28. Mani, T.; Grills, D. C.; Newton, M. D.; Miller, J. R., Electron Localization of Anions

Probed by Nitrile Vibrations. J. Am. Chem. Soc. 2015, 137, 10979-10991.

29. Okano, T.; Iwahara, M.; Konishi, H.; Kiji, J., Synthesis and Properties of Crown Ether-

Modified Phosphines and Their Use as Ligands in Transition-Metal Catalysts. J. Organomet.

Chem. 1988, 346, 267-275.

Page 31: Identification of Ion-Pair Structures in Solution by

29

30. Kimura, M.; Shi, K.; Hashimoto, K.; Hu, Z. Z., Establishment of an Efficient Synthetic

Route to 3,4 : 3',4'-bis(3,6,9-trioxaundecane-1,11-dioxy)benzil. Heterocycles 2006, 68, 2375-

2380.

31. Frisch, M. J., et al. Gaussian 09, Gaussian, Inc.: Wallingford, CT, USA, 2009.

32. Shao, Y., et al., Advances in Molecular Quantum Chemistry Contained in the Q-Chem 4

Program Package. Mol. Phys. 2014, 1-32.

33. Vydrov, O. A.; Scuseria, G. E.; Perdew, J. P., Tests of functionals for systems with

fractional electron number. J. Chem. Phys. 2007, 126, 154109 1-9.

34. Becke, A. D., Density-Functional Thermochemistry. III. The Role of Exact Exchange. J.

Chem. Phys. 1993, 98, 5648-5652.

35. Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J., Ab-Initio Calculation of

Vibrational Absorption and Circular-Dichroism Spectra Using Density-Functional Force-Fields.

J. Phys. Chem. 1994, 98, 11623-11627.

36. Tomasi, J.; Mennucci, B.; Cammi, R., Quantum Mechanical Continuum Solvation

Models. Chem. Rev. 2005, 105, 2999-3093.

37. Miertus, S.; Scrocco, E.; Tomasi, J., Electrostatic Interaction of a Solute with a

Continuum - a Direct Utilization of Abinitio Molecular Potentials for the Prevision of Solvent

Effects. Chem. Phys. 1981, 55, 117-129.

38. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V., Energies, Structures, and Electronic

Properties of Molecules in Solution with the C-PCM Solvation Model. J. Comput. Chem. 2003,

24, 669-681.

39. Barone, V.; Cossi, M., Quantum Calculation of Molecular Energies and Energy Gradients

in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995-2001.

Page 32: Identification of Ion-Pair Structures in Solution by

30

40. Truong, T. N.; Stefanovich, E. V., A New Method for Incorporating Solvent Effect into

the Classical, Ab-Initio Molecular-Orbital and Density-Functional Theory Frameworks for

Arbitrary Shape Cavity. Chem. Phys. Lett. 1995, 240, 253-260.

41. Levinson, N. M.; Fried, S. D.; Boxer, S. G., Solvent-Induced Infrared Frequency Shifts in

Aromatic Nitriles are Quantitatively Described by the Vibrational Stark Effect. J. Phys. Chem. B

2012, 116, 10470-10476.

42. LeSuer, R. J.; Buttolph, C.; Geiger, W. E., Comparison of the Conductivity Properties of

the Tetrabutylammonium Salt of Tetrakis(pentafluorophenyl)borate Anion with Those of

Traditional Supporting Electrolyte Anions in Nonaqueous Solvents. Anal. Chem. 2004, 76, 6395-

6401.

43. Geiger and coworkers reported the value of 35. 7 M-1 in acetonitrile (De = 37.5). We

assume that the value in DMF (De = 38.25) is similar.

44. Roberts, R. C.; Szwarc, M., Chemistry of Radical Anion of Tetraphenylethylene. J. Am.

Chem. Soc. 1965, 87, 5542-5548.

45. Masterton, W. L.; Bolocofsky, D.; Lee, T. P., Ionic Radii from Scaled Particle Theory of

Salt Effect. J. Phys. Chem. 1971, 75, 2809-2815.

46. Marcus, Y., Tetraalkylammonium Ions in Aqueous and Non-Aqueous Solutions. J.

Solution. Chem. 2008, 37, 1071-1098.

47. Fry, A. J., Computational Studies of Ion Pairing. 9. The "Steric" Effect of

Tetraalkylammonium Ions with Electrochemically Generated Anions is Not Steric. Electrochem.

Commun. 2013, 35, 88-90.

Page 33: Identification of Ion-Pair Structures in Solution by

31

48. Fry, A. J., Computational Studies of Ion Pairing. 10. Ion Pairing between

Tetrabutylammonium Ion and Inorganic Ions: A General Motif Confirmed. J. Org. Chem. 2015,

80, 3758-3765.

49. Boche, G., Penetrated Ion Pairs: a New Form of Ion-Pair. Angew. Chem. Int. Edit. 1992,

31, 731-732.

50. Yang, X. Q.; Zaitsev, A.; Sauerwein, B.; Murphy, S.; Schuster, G. B., Penetrated Ion-

Pairs - Photochemistry of Cyanine Dyes within Organic Borates. J. Am. Chem. Soc. 1992, 114,

793-794.

51. Abbott, A., Electrochemistry in Media of Low Dielectric-Constant. Chem. Soc. Rev.

1993, 22, 435-440.

52. Davlieva, M. G.; Lu, J. M.; Lindeman, S. V.; Kochi, J. K., Crystallographic Distinction

between "Contact" and "Separated" Ion Pairs: Structural Effects on Electronic/ESR spectra of

Alkali-Metal Nitrobenzenides. J. Am. Chem. Soc. 2004, 126, 4557-4565.

53. Lü, J. M.; Rosokha, S. V.; Neretin, I. S.; Kochi, J. K., Quinones as Electron Acceptors.

X-ray Structures, Spectral (EPR, UV-vis) Characteristics and Electron-Transfer Reactivities of

Their Reduced Anion Radicals as Separated vs Contact Ion Pairs. J. Am. Chem. Soc. 2006, 128,

16708-16719.

54. Lü, J. M.; Rosokha, S. V.; Lindeman, S. V.; Neretin, I. S.; Kochi, J. K., "Separated"

versus "Contact" Ion-Pair Structures in Solution from Their Crystalline States: Dynamic Effects

on Dinitrobenzenide as a Mixed-Valence Anion. J. Am. Chem. Soc. 2005, 127, 1797-1809.

55. Fry, A. J., Tetraalkylammonium Ions are Surrounded by an Inner Solvation Shell in

Strong Electron Pair Donor Solvents. Electrochem. Commun. 2009, 11, 309-312.

Page 34: Identification of Ion-Pair Structures in Solution by

32

56. Wardman, P., Reduction Potentials of One-Electron Couples Involving Free-Radicals in

Aqueous-Solution. J. Phys. Chem. Ref. Data 1989, 18, 1637-1755.

57. Evans, D. H., One-Electron and Two-Electron Transfers in Electrochemistry and

Homogeneous Solution Reactions. Chem. Rev. 2008, 108, 2113-2144.

58. Saveant, J. M., Effect of Ion Pairing on the Mechanism and Rate of Electron Transfer.

Electrochemical Aspects. J. Phys. Chem. B 2001, 105, 8995-9001.

59. Bard, A. J. F., Larry R., Electrochemical Methods: Fundamentals and Applications. 2nd

ed.; John Wiley & Sons, Inc.: 2001.

60. Mani, T.; Miller, J. R., Role of Bad Dihedral Angles: Methylfluorenes Act as Energy

Barriers for Excitons and Polarons of Oligofluorenes. J. Phys. Chem. A 2014, 118, 9451-9459.

61. Fratiloiu, S.; Grozerna, F. C.; Koizumi, Y.; Seki, S.; Saeki, A.; Tagawa, S.; Dudek, S. P.;

Siebbeles, L. D. A., Electronic Structure and Optical Properties of Charged Oligofluorenes

Studied by VIS/NIR Spectroscopy and Time-Dependent Density Functional Theory. J. Phys.

Chem. B 2006, 110, 5984-5993.

62. Pedersen, C. J., Cyclic Polyethers and Their Complexes with Metal Salts. J. Am. Chem.

Soc. 1967, 89, 2495-2496.

63. Pedersen, C. J., Cyclic Polyethers and Their Complexes with Metal Salts. J. Am. Chem.

Soc. 1967, 89, 7017-7036.

64. Gokel, G. W.; Leevy, W. M.; Weber, M. E., Crown Ethers: Sensors for Ions and

Molecular Scaffolds for Materials and Biological Models. Chem. Rev. 2004, 104, 2723-2750.

Page 35: Identification of Ion-Pair Structures in Solution by

33

Table of Contents Graphic