Top Banner
UNCORRECTED PROOF CCR 110408 1–10 Coordination Chemistry Reviews xxx (2005) xxx–xxx Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles 3 4 Mohammad A. Omary a , Ahmed A. Mohamed b , Manal A. Rawashdeh-Omary a , John P. Fackler Jr. a,5 6 a Department of Chemistry, University of North Texas, Denton, TX 76203, USA 7 b Department of Chemistry, Texas A&M University, P.O. Box 30012, College Station, TX 77843, USA 8 Received 2 August 2004; accepted 10 December 2004 Contents 9 10 Abstract .................................................................................................................. 00 11 Keywords ................................................................................................................ 00 12 1. Introduction .......................................................................................................... 00 13 2. Isolated trinuclear compounds .......................................................................................... 00 14 2.1. Nucleophilic nature ............................................................................................. 00 15 2.2. Photophysics ................................................................................................... 00 16 3. Adducts with organic electrophiles ...................................................................................... 00 17 3.1. Perfluoronaphthalene adduct ..................................................................................... 00 18 3.2. Hexafluorobenzene adduct ....................................................................................... 00 19 3.3. Tetracyanoquinodimethane adduct ................................................................................ 00 20 4. Conclusions .......................................................................................................... 00 21 Acknowledgments ......................................................................................................... 00 22 References ............................................................................................................... 00 23 24 25 Abstract 26 Two nucleophilic trinuclear Au(I) ring complexes, Au 3 ((p-tolyl)N = C(OEt)) 3 , 1, and Au 3 (-C 2 ,N 3 -bzim) 3 , 2, form sandwich adducts with organic Lewis acids (C 6 F 6 and C 10 F 8 ) and electron acceptors (tetracyanoquinodimethane, TCNQ), a neutral polyfunctional inorganic Lewis acid (Hg 3 (o-C 6 F 4 ) 3 ), and naked heavy metal cations (Tl + and Ag + ). Fascinating photophysical properties are associated with the adducts formed, including external heavy-atom effect leading to room-temperature phosphorescence of aromatic molecules, reversible quenching and regeneration of luminescence upon exposure to organic vapors, and sensitization of metal-centered emissions in supramolecular stacks with luminescence thermochromism. Intermolecular interactions, including M–M bonding and acid–base interactions, play the major role in influencing the luminescence properties. 27 28 29 30 31 32 33 © 2005 Published by Elsevier B.V. 34 Keywords: Trinuclear Au(I) complexes; Phosphorescence; Heavy-atom effect; Aurophilic bonding; X-ray structures 35 36 Corresponding author. Tel.: +1 979 845 0648; fax: +1 979 845 2373. E-mail address: [email protected] (J.P. Fackler Jr.). 1. Introduction 1 The Lewis donor–acceptor concept [1] and the re- 2 lated Ingold–Robinson nucleophile–electrophile concept [2] 3 are traditionally associated with donation and accep- 4 1 0010-8545/$ – see front matter © 2005 Published by Elsevier B.V. 2 doi:10.1016/j.ccr.2004.12.018
10

Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

May 10, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

F

Coordination Chemistry Reviews xxx (2005) xxx–xxx

Photophysics of supramolecular binary stacks consisting ofelectron-rich trinuclear Au(I) complexes and organic electrophiles

3

4

Mohammad A. Omarya, Ahmed A. Mohamedb,Manal A. Rawashdeh-Omarya, John P. Fackler Jr.a,∗

5

6

a Department of Chemistry, University of North Texas, Denton, TX 76203, USA7b Department of Chemistry, Texas A&M University, P.O. Box 30012, College Station, TX 77843, USA8

Received 2 August 2004; accepted 10 December 2004

Contents9

10

Abstract. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0011

Keywords. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0012

1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0013

14

15

16

17

18

19

20

21

22

23

24

25

26

hc Lewisductsnching and

r stacks within

27

28

29

30

31

32

33

34

35

36

1

2

TED

PR

OO2. Isolated trinuclear compounds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

2.1. Nucleophilic nature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 002.2. Photophysics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

3. Adducts with organic electrophiles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 003.1. Perfluoronaphthalene adduct. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 003.2. Hexafluorobenzene adduct. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 003.3. Tetracyanoquinodimethane adduct. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00Acknowledgments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

Abstract

Two nucleophilic trinuclear Au(I) ring complexes, Au3((p-tolyl)N = C(OEt))3, 1, and Au3(�-C2,N3-bzim)3, 2, form sandwich adducts witorganic Lewis acids (C6F6 and C10F8) and electron acceptors (tetracyanoquinodimethane, TCNQ), a neutral polyfunctional inorganiacid (Hg3(o-C6F4)3), and naked heavy metal cations (Tl+ and Ag+). Fascinating photophysical properties are associated with the adformed, including external heavy-atom effect leading to room-temperature phosphorescence of aromatic molecules, reversible queregeneration of luminescence upon exposure to organic vapors, and sensitization of metal-centered emissions in supramoleculaluminescence thermochromism. Intermolecular interactions, including M–M bonding and� acid–base interactions, play the major roleinfluencing the luminescence properties.

1

2

t 3

- 4

CO

RR

EC

CCR 110408 1–10

© 2005 Published by Elsevier B.V.

Keywords: Trinuclear Au(I) complexes; Phosphorescence; Heavy-atom effect; Aurophilic bonding; X-ray structures

∗ Corresponding author. Tel.: +1 979 845 0648; fax: +1 979 845 2373.E-mail address:[email protected] (J.P. Fackler Jr.).

1. Introduction

The Lewis donor–acceptor concept[1] and the re-lated Ingold–Robinson nucleophile–electrophile concep[2]are traditionally associated with� donation and accep

0010-8545/$ – see front matter © 2005 Published by Elsevier B.V.doi:10.1016/j.ccr.2004.12.018

N

U
Page 2: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

DO

F

0

2 M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx

tance of “electron pairs”[3]. But with the development of5

organometallic chemistry, it has become critical to expand6

these concepts to include� acids and bases. Convention-7

ally, the � base is an unsaturated organic molecule while8

metal centers are usually acids/acceptors. Our recent efforts,9

however, have demonstrated that the trinuclear Au(I) com-10

poundswith substituted imidazolate and carbeniate bridging11

ligands,Plate 1, act as� bases which form supramolecu-12

lar stacks with a variety of electrophiles[4–8]. These elec-13

trophiles include organic Lewis acids (C6F6 and C10F8) and14

electron acceptors (tetracyanoquinodimethane, TCNQ), a15

neutral polyfunctional inorganic Lewis acid (Hg3(o-C6F4)3),16

and naked heavy metal cations (Tl+ and Ag+). The resulting17

sandwich adducts exhibit interesting bonding and optoelec-18

tronic properties. The focus of this article will be on the pho-19

tophysical properties of the trinuclear Au(I) complexes and20

the sandwich� adducts thereof with organic electrophiles. A21

review of the structural properties of trinuclear Au(I) com-22

plexes in general has been reported elsewhere[9].23

Our efforts are complementary to those by many re-24

searchers in areas that span structural, chemical, and physi-25

cal properties of extended-chain materials and trinuclear d1026

ring complexes. Extended linear-chain compounds that con-27

tain metal atoms exhibit important aspects in chemical bond-28

ing as well as fascinating chemical and physical properties29

[10–12]. Many classes of coordination compounds that ex-30

h ad-31

v hitec-32

t nes-33

c ronic34

a this35

a iode36

f -37

c plex38

t sence39

o x,40

{ re-41

v42

a43

t44

c45

m ducts46

w and47

n etal-48

l49

An especially intriguing class of linear-chain species involves50

trinuclear d10 complexes, which have garnered considerable51

interest in recent years owing in large part to their fasci-52

nating luminescence properties. For example, Balch and co-53

workers reported that a trinuclear carbeniate Au(I) complex54

exhibits “solvoluminescence”, i.e., it produces spontaneous55

orange emission upon contact with solvent in samples that56

had been irradiated with long-wavelength UV light[21a]. The 57

same and related complexes have been found to form charge-58

transfer complexes with nitro-9-fluorenones[21b] and some 59

form hourglass figures on standing or when placed in an acid60

[21c]. The combined work reported by us[7], Balch and co- 61

workers[21], and Yang and Raptis[22] indicates that the 62

luminescence of the trinuclear Au(I) complexes, regardless63

of the bridging ligand, is related to intermolecular Au–Au in-64

teractions between adjacent trimers. Gabbaı and co-workers 65

have reported that a trinuclear Hg(II) complex forms 1:166

adducts with aromatic hydrocarbons, which become brightly67

phosphorescent at room temperature due to a mercury heavy-68

atom effect[23,24]. Finally, a recent report by Dias et al. 69

showed that a trinuclear Cu(I) pyrazolate complex exhibits70

bright emissions that can be tuned to multiple luminescence71

colors across the visible region by varying the temperature,72

solvent, or concentration[25]. 73

2 74

2 75

fo-76

c 77

b as 78

d 79

a nd 80

e ter-81

a l- 82

c s. 83

T c- 84

u tive85

M c ar-86

r the87

e - 88

g they89

e ence,90

t e 91

d y- 92

d er of93

a work94

o 95

2 m 96

s n 97

f fas-98

c ot 99

d r fo-100

c ducts101

w

UN

CO

RR

EC

TE

ibit such structures have been reported, giving rise toances in several areas that include supramolecular arcure, acid–base chemistry, metallophilic bonding, lumient materials, optical sensing, and various optoelectpplications. Among the many recent developments inrea are reports about a vapochromic light-emitting d

rom linear-chain Pt(II)/Pd(II) complexes[13], a luminesent switch consisting of an Au(I) dithiocarbamate comhat possesses a luminescent linear chain in the pref vapors of organic solvents[14], a vapochromic compleTl[Au(C6Cl5)2]}n, that shows a quantum dot effect andersible color changes upon binding to volatile organics[15],modified form of Magnus’ green salt [(NH3)4Pt][PtCl4],

hat exhibits semiconducting properties[16], a variety of d8

omplexes that interact with inorganic[17] and organic[18]olecules to form donor–acceptor extended-chain adith interesting conducting and/or magnetic properties,ew interesting complexes that exhibit strong heterobim

ic bonding between different closed-shell metal ions[19,20].

Plate 1. Sketch of the trinuclear AuI compounds studied.

PR

O

CCR 110408 1–1

. Isolated trinuclear compounds

.1. Nucleophilic nature

The two trinuclear Au(I) ring complexes that we haveused on, Au3((p-tolyl)N=C(OEt))3, 1, and Au3(�-C2,N3-zim)3, 2, are shown inPlate 1. The two compounds stackimers in the solid state with a chair arrangement in1 [26]nd a prismatic arrangement in2 [27]. Hence, each compouxhibits both intramolecular and intermolecular Au–Au inctions, i.e.,aurophilic bonding[28]. Density-functional caulations have indicated that1and2are electron-rich speciewo views are shown inFig. 1to illustrate the negative molelar electrostatic potential (MEP); the positive and negaEP regions in space are shown for the crystallographi

angement of1 while the MEP values are mapped onlectron density surface of2 [7]. The most nucleophilic reions are clearly at the center of the trinuclear ring andxtend in the space perpendicular to the ring plane. Hhe� base concept for1 and2 is borne out according to thata inFig. 1. This is similar to the action of aromatic hrocarbons as� bases that can attract cations at the centnd perpendicular to the aromatic ring plane; see thef Dougherty on cation-� interactions[29]. In fact, 1 andare involved in strong cation-� interactions, as they for

andwich adducts with Tl+ and Ag+ ions, which have beeully characterized by X-ray crystallography and exhibitinating luminescence properties[4,5]. Such adducts are niscussed in detail in this review, however, as the majous herein is on the trinuclear species alone and their adith organic electrophiles.

Page 3: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

CTE

D P

RO

OF

0

M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx 3

Fig. 1. Representation of the molecular electrostatic potential (MEP) from DFT calculations[7]. The positive (green) and negative (red) MEP regions in spaceare shown for the crystallographic arrangement of a dimeric unit of1 (left) while the MEP values are mapped on the electron density surface of a molecule of2 (right).

2.2. Photophysics102

The luminescence emission and excitation spectra of1 in103

the solid state are shown inFig. 2. The solid exhibits a feeble104

blue luminescence that becomes bright at 77 K. The lifetime105

is ∼10�s [30], hence the emission is phosphorescence from106

a formally triplet excited state. The emission shows a struc-107

tured profile with a uniform vibronic spacing (∼1400 cm−1).108

This spacing corresponds to the Raman spectrum of1 in so-109

lution, which shows a peak at∼1422 cm−1 and an overtone110

thereof (Fig. 3, right). This peak is assignable to theνC–N111

vibration of the bridging carbeniate ligand. The higher fre-112

quency region shows other vibrations characteristic of the113

aliphatic and aromatic substituents in1 on the blue and red114

side of 3000 cm−1, respectively. At first sight, it is tempting115

to assign the blue emission of1 to a ligand-centered phospho-116

rescence from a single molecular unit of1. This is because117

the emission is highly structured, typical of ligand-centered118

emissions in monomeric organic molecules/ligands and be-119

cause the vibronic spacing corresponds to theνC–N vibration 120

in a dilute solution. However, further data in Section3.2be- 121

low suggest that the emission likely occurs from a dimeric122

unit. Furthermore, the luminescence excitation spectra show123

peaks at too low energies to be assigned to a monomeric124

species. The emitting state, therefore, is likely a ligand to125

metal–metal charge transfer (LMMCT). 126

3. Adducts with organic electrophiles 127

3.1. Perfluoronaphthalene adduct 128

Reaction of1 with octafluoronaphthalene, C10F8, forms a 129

bright yellow phosphorescent solid[31]. The crystal structure 130

itation

NC

OR

RE

Fig. 2. Emission (right) and corrected exc

U

CCR 110408 1–1

(left) spectra of a crystalline sample of1 at 77 K.

Page 4: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

ED

PR

OO

F

CCR 110408 1–10

4 M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx

Fig. 3. Raman spectrum of a solution of1 in CH2Cl2 at ambient temperature.

shows a 1:1 sandwich adducts,1·C10F8, with a supramolecu-131

lar infinite-chain structure. The packing of the molecules in a132

{1·C10F8}∞ chain is shown inFig. 4. Similar 1:1 stacks form133

between1 and another perfluorinated aromatic molecule,134

C6F6 (vide infra). Given the nucleophilic character of1 and135

the known electrophilic character of the Lewis acids C8F10136

and C6F6, electrostatic interactions are expected to be the137

dominant forces that stabilize the supramolecular structure138

of both {1·C10F8}∞ and {1·C6F6}∞ � acid–base stacks.139

The distance between the centroid of C10F8 to the centroid140

of 1 is 3.509A compared to the slightly longer distance of141

3.565A in 1·C6F6. The Au· · ·Au intramolecular interactions142

in 1 (3.224, 3.288 and 3.299A) become longer upon adduct143

formation in 1·C10F8 (3.279, 3.280, 3.337A). There is no144

significant change in the distances in the fluorinated naphtha-145

lene rings upon adduct formation. This is consistent with a146

recent report by Schafer and co-workers[32] in which the au-147

thors have shown that the neutral and anionic forms of C10F8148

both possess a D2h symmetry and the electron affinity of the149

neutral species is 1.01 eV compared to 0.69 eV for hexafluo-150

F with1

robenzene. The larger electron affinity of C10F8 than that of 151

C6F6 explains the shorter interplanar separation in1·C10F8 152

compared to1·C6F6. 153

The1·C10F8 solid adduct exhibits a yellow emission band154

that is bright even at ambient temperature.Fig. 5shows that 155

the emission spectra of crystals of1·C10F8 in comparison 156

with the solid uncomplexed C10F8 aromatic molecule. A 157

structured profile is observed for1·C10F8 with a better resolu- 158

tion obtained upon cooling to 77 K. The profile is essentially159

the same as that of crystals of uncomplexed octafluoronaph-160

thalene but there is a slight red shift. At ambient tempera-161

ture, the yellow organic-centered luminescence is very bright162

for 1·C10F8 but undetectable for solid octafluoronaphthalene.163

F 10 8 tand solid C10F8. No luminescence was detected at room temperature fromsolid C10F8.

NC

OR

RE

CT

ig. 4. Thermal ellipsoid drawing of the stacked octafluoronaphthalene.

U

ig. 5. Emission spectra of single crystals of the1·C F stacked adduc

Page 5: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

ED

OF

0

M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx 5

Lifetime measurements show that the yellow luminescence164

of 1·C10F8 is phosphorescence. The decay curve at 550 nm165

shows a single exponential fit withτ = 3.57± 0.07 ms. The166

magnitude of the lifetime did not change significantly with167

variations in either the laser excitation wavelength or the168

monitored emission wavelength.169

The structured profile, energy, and lifetime of the yel-170

low emission of the1·C10F8 adduct indicate that the emis-171

sion should be assigned to monomer phosphorescence of the172

octafluoronaphthalene, for which the energy of the T1 → S0173

radiative transition is red shifted only slightly. The 3.6 ms174

lifetime of the yellow emission of1·C10F8 is two orders of175

magnitude shorter than the lifetime of the octafluoronaph-176

thalene phosphorescence, which was reported to be in the177

range of 0.25–0.38 s in frozen glasses[33–35]. The bright-178

ness of the yellow phosphorescence at room temperature for179

the1·C10F8 solid adduct along with the great reduction in the180

lifetime compared to free octafluoronaphthalene are indica-181

tive of a strong gold heavy-atom effect. The slight red shift in182

the emission energy of1·C10F8 is also a known consequence183

of the heavy-atom effect[33,35]. The interactions between184

1 and octafluoronaphthalene are secondary� interactions, as185

evidenced by the relatively long crystallographic distances186

between the planes of the two components (∼3.5A between187

the centroids). The energies of such interactions have been es-188

timated to be in the range of only 1–2 kcal/mol in similar sys-189

t ffect190

s to191

t nds192

w ment193

( rac-194

t avy-195

a lent196

b -197

a dest198

d ase is199

m tom200

e201

B .11 s202

d tively203

[ n for204

1 ct of205

g n in206

τ ara-207

b ts.208

T f209

A210

B211

T ds212

t u-213

n r214

o ich215

a with216

t217

f218

t s of219

5d10 heavy metal ions lead to an unusually strong external220

heavy-atom effect that is comparable to the internal heavy-221

atom effect in organic compounds. The long-range ordering222

of the acid–base stacks in which each organic triplet emitter223

is surrounded by six heavy metal atoms is likely the major224

contributing factor to the unusually strong external heavy-225

atom effect seen here. This effect, however, is smaller than226

the internal effect in which metals are involved in coordinate227

covalent bonds with the organic moiety, whereinτP values 228

of 101–102 �s are common for ligand-centered emissions in229

5d10 complexes[45] and even in lighter 3d10 and 4d10 com- 230

plexes[46]. The phosphorescence lifetimes are even shorter231

for metal-centered emissions, whereinτP values of a few mi- 232

croseconds and even 102 ns are quite common, e.g., for Au(I)233

complexes that show gold-centered emissions[47]. 234

In order to gain insight into the photophysical processes235

that lead to the enhanced phosphorescence in1·C10F8, ab- 236

sorption, diffuse–reflectance, and luminescence excitation237

spectra were obtained.Fig. 6 shows the diffuse reflectance238

spectrum of the solid adduct1·C10F8. Because the adduct239

stacks in a 1:1 manner, the diffuse reflectance data are com-240

pared with the absorption spectra for dilute solutions of1 241

and of C10F8, which represent monomers of these species.242

Fig. 6 shows the1·C10F8 adduct exhibits not only absorp-243

tions characteristic of its two monomer components, but also244

n tend245

t sible246

r these247

n that248

l 249

t ar 335250

a ed251

d tures252

b ence253

e ance254

s rre-255

s 256

Fa dilutes

UN

CO

RR

EC

T

ems[36,37]. Hence, one can consider the heavy-atom eeen here for the1·C10F8 adduct to be more comparablehe external heavy-atom effect known for organic compouhen a heavy atom is present in the luminophore environ

e.g., in the solvent, host matrix, or “through space” inteions within the molecule), as opposed to the internal hetom effect where a heavy atom is involved in a direct covaonding to the luminophore[38–42]. The external heavytom effect in organic compounds usually leads to a moecrease in phosphorescence lifetimes while this decreuch more significant in the case of the internal heavy-affect. For example, substituting naphthalene (τP = 2.6 s) withr and norborane-Br reduces the lifetime to 0.02 and 0ue to internal and external heavy-atom effects, respec

38,39]. The reduction in phosphorescence lifetime see·C10F8 to ms levels due to an external heavy-atom effeold is much more significant than the common reductioP due to external heavy-atom effects and is more comple to the reduction inτP due to internal heavy-atom effeche spin–orbit coupling “�” parameter for the 5d orbital ou(I) is 5100 cm−1 [43], comparable to theζ values for ther and I atoms of 2460 and 5700 cm−1, respectively[39,44].he internal heavy-atom effect in�-iodonaphthalene lea

o τP = 2 ms[39]. Meanwhile, lifetime data recently commicated by Omary and Gabbaı showτP values on the ordef 10−1 to 100 ms for several 1:1 adducts of electron-rromatic hydrocarbons (naphthalene; biphenyl; pyrene)

he macrocyclic Lewis acid�-(o-C6F4)3Hg3 [24]. There-ore, it is reasonable to conclude that secondary� interac-ions of aromatic luminophores with trinuclear complexe

PR

O

CCR 110408 1–1

ew red-shifted features (designated by arrows) that exhe absorption edge of the adduct to approach the viegion. Luminescence excitation spectra suggest thatew absorptions represent the major excitation route

eads to the yellow luminescence of1·C10F8. Fig. 7 showshat distinct luminescence excitation peaks appeared nend 360 nm for1·C10F8, which correspond to the red-shiftiffuse–reflectance features. The fact that these new feaecome much more clearly discernible in the luminescxcitation spectrum than they were in the diffuse reflectpectrum indicates the central role played by their coponding transitions in the excitation route of1·C10F8. For

ig. 6. Solid-state diffuse reflectance spectrum of the1·C10F8 stackeddduct in comparison to its free molecular components, represented byolutions of1 and octafluoronaphthalene in acetonitrile.

Page 6: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

CTE

DO

F

0

6 M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx

Fig. 7. Overlay of various electronic spectra of1·C10F8 and free C10F8 forpurposes of comparison.

comparison,Fig. 7 also shows the luminescence excitation257

spectrum of C10F8 alone, which exhibits luminescence ex-258

citation peaks at shorter wavelengths and correspond to the259

monomer absorption peaks in the absorption spectrum of di-260

lute solutions of C10F8.261

We assign the diffuse reflectance/excitation peaks at∼335262

and 360 nm for1·C10F8 to charge-transfer transitions in263

the ground state adduct. The trinuclear Au(I) complex1 is264

strongly nucleophilic (Section2.1), so it acts as a� base and265

donates electron density to the� acid C10F8. The resulting266

charge-transfer adduct exhibits characteristic absorptions that267

are red-shifted from the absorptions of its individual compo-268

nents. Kisch et al. suggested that donor-to-acceptor charge-269

transfer bands should appear in the diffuse–reflectance spec-270

tra of 1:1 donor:acceptor inorganic:organic ionic solid stacks271

[48]. Also, previous work by us (vide infra; Section 2.3)272

[8] and by Balch and co-workers[21b] showed that charge-273

transfer absorption bands in the visible and near IR re-274

gion arise inneutral adducts that form upon interaction of275

trinuclear Au(I) complexes with organic acceptors such as276

TCNQ and nitro-substituted fluorenones. No luminescence277

data were reported in these previous reports, but they sup-278

port the charge-transfer assignment of the lowest-energy279

diffuse–reflectance/excitation peaks of1·C10F8. Although280

the mechanism of the external heavy-atom effect of organic281

luminophores has been subject to numerous interpretations282

[ rigin,283

i a role284

i t has285

b a for286

r f re-287

l toms288

(289

o arge-290

t hown291

i or-292

g to an293

e294

The conventional heavy-atom effect in organic molecules295

usually invokes a phosphorescence route that entails simply296

the enhancement of the S1–T1 intersystem crossing of the 297

organic compound following direct absorption from S0 to ei- 298

ther S1 or a higher singlet (e.g., S2) that then relaxes to S1 299

via internal conversion[38,39]. In contrast, the spectral data300

for the 1·C10F8 adduct here suggest a different excitation301

route for the phosphorescence, as depicted inFig. 8. Absorp- 302

tion occurs directly to the resulting charge-transfer states in303

the1·C10F8 adduct. The significantly lower intensity of the304

lowest-energy feature near 360 nm in the diffuse reflectance305

and excitation spectra of the adduct suggests that it represents306

a triplet charge transfer state (3CT) while the higher energy 307

feature near 335 nm is assigned to a singlet charge transfer308

state (1CT). Fortunately, these charge transfer states lie higher309

in energy than the energy of the T1 state of the organic com- 310

ponent such that the latter state will not be depopulated as a311

result of the charge transfer process. Thus, the lowest-energy312

emitting state in the adduct remains as the T1 state with little 313

perturbation of its original energy in the organic component314

alone. The spectral data suggest that the S1 state of C10F8 is 315

not involved in the charge transfer process because vibronic316

features corresponding to this state remain essentially unper-317

turbed in the spectrum of the binary adduct (see the dashed318

lines inFig. 6). Hence, we illustrate inFig. 8that the3CT and 319

1CT states arise from the molecular orbital interaction of the320

T 321

w g 322

i rum323

t s 324

t r low-325

e irect326

a s 327

f 328

3 329

t 330

e 331

F en thee the1

UN

CO

RR

E49] and there seems to be lack of consensus on its ot has been suggested that charge transfer might playn the external heavy-atom effect. Evidence to this effeceen gathered, typically by analysis of the emission datigid glasses at cryogenic temperatures for a number oated luminophores that contain one or more heavy-ausually halogens), or by theoretical studies[49]. But, tour knowledge, distinct peaks due to the suggested ch

ransfer process in the ground-state adduct, like those sn Figs. 6 and 7, have not been reported for conventionalanic luminophores that exhibit phosphorescence duexternal heavy-atom effect.

PR

O

CCR 110408 1–1

2 and S2 states of C10F8 with suitable frontier orbitals in1hile both the S1 and T1 states of C10F8 remain non-bondin

n the1·C10F8 adduct. The luminescence excitation specthat monitors the yellow phosphorescence (Fig. 7) suggesthat the charge transfer process represents the majonergy excitation route while excitation peaks due to dbsorption to the triplet are absent (e.g., S0 → T1 absorption

or naphthalenes lie between 400 and 500 nm)[39].

.2. Hexafluorobenzene adduct

Unlike the1·C10F8 adduct, the1·C6F6 adduct does noxhibit phosphorescence from the T1 state of C6F6 on com-

ig. 8. Proposed energy level diagram showing the interaction betwexcited states of1 and C10F8 to form the charge transfer (CT) states in·C10F8 adduct.

Page 7: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

ED

OF

CCR 110408 1–10

M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx 7

plexation with1. A reasonable explanation may lie in the fact332

that previous experimental data in the organic literature sug-333

gest that perfluorination leads to an internal heavy-atom effect334

for the octafluoronaphthalene ring but not for the hexafluo-335

robenzene ring (for further details, see refs.[34,50]). The336

1·C6F6 solid adduct is actually not luminescent at all even337

at cryogenic temperatures. Because1 exhibits intermolec-338

ular Au–Au interactions between adjacent trimers (Section339

2.1) and exhibits blue phosphorescence (Section2.2) while340

both these interactions are absent in the non-luminescent 1:1341

stacks of the1·C6F6 solid adduct, we hypothesized that the342

presence of luminescence in1 and its absence in1·C6F6 are343

related to the presence or absence of intermolecular Au–Au344

interactions between adjacent trimers. This hypothesis was345

substantiated by our observation that, when1·C6F6 is im-346

mersed in a solvent that does not dissolve1, e.g., Et2O, the347

compound looses its crystallinity and the resulting powder348

exhibits the same blue photoluminescence characteristic of349

1, indicating that C6F6 is liberated from1·C6F6. In contrast,350

when solid1 is suspended in C6F6, its blue luminescence351

starts to quench with time. These observations prompted a352

study of the interaction of these complexes with vapors of353

organic compounds.Fig. 9 shows a representative example.354

A solid sample of1 was placed at a fixed position in a closed355

chamber and its luminescence spectra were acquired as a356

function of time in the presence of CF vapors. The vapors357

w bient358

t359

s ham-360

b d361

a l the362

n of363

C n the364

l365

that,366

1 lear367

A ab-368

s b-369

s370

F tt 51, 65,9

for even nearly saturated solutions of1·C6F6 in CDCl3 show 371

one peak at the same chemical shift (−168 ppm) as that ob- 372

served for C6F6 alone in CDCl3. 373

3.3. Tetracyanoquinodimethane adduct 374

Reaction of a saturated solution of 7,7,8,8-tetracyano-375

quinodimethane with a sample of2added dropwise results in 376

an instant dramatic color change from light yellow to intense377

green. Crystallization produced dark crystals whose struc-378

ture showed infinite 2:1 stacks,{[2]2·TCNQ}∞, as shown 379

in Fig. 10 [8]. The TCNQ molecule is sandwiched between380

two units of2 from each side, in a face-to-face manner so381

that a molecule of the2·TCNQ adduct is best represented by382

the formula (�-2)(µ-�-TCNQ)(�-2). The cyanide groups are383

clearly not coordinated to the gold atoms; hence,� acid–base 384

interactions are favored to� coordination of TCNQ to the 385

trinuclear2 despite the strong� donation ability of TCNQ 386

and the low coordination number, 2, for the Au atoms in2. 387

Although the distance between the centroid of TCNQ to the388

centroid of the Au3 unit is very long, 3.964A, consistent with 389

the aforementioned nature of such secondary� interactions, 390

Fig. 10. Crystal structure of the2·TCNQ adduct showing a view of thecolumnar structure formed by the repeat unit. The benzyl groups are omittedfor clarity [8].

UN

CO

RR

EC

T

6 6ere produced as a result of the vapor pressure at am

emperature and pressure of a liquid sample of neat C6F6 in amall beaker that is placed at the bottom of the closed cer. The quenching of the luminescence of1was first detectefter only several minutes and continued gradually untiext day (Fig. 9). A control experiment in the absence6F6 vapors was carried out and showed no quenching i

uminescence of1 with time.The evidence we have gathered thus far suggests

·C6F6 dissociates in solution into its component trinucu(I) compound and organic Lewis acid. The electronicorption spectra for solutions of1·C6F6 show the same aorption peaks as those for1and C6F6. The19F NMR spectra

ig. 9. Emission spectra of1 vs. exposure time to C6F6 vapor at ambienemperature and pressure. The exposure time was (top-to-bottom) 0,8, 148, 208, 1322, 1439, and 1462 min, respectively.

PR

O

Page 8: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

CTE

D P

RO

OF

0

8 M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx

Fig. 11. Visible absorption spectrum of2 TCNQ in CH2Cl2 at room temperature. The insert shows the SEM image of the crystals taken without a conductivefilm coating[8].

dramatic changes in the physical and electronic properties391

take place as a result of these interactions. Multiple evidence392

has been collected to suggest the presence of partial charge393

delocalization in the{[2]2·TCNQ}∞ stacks. First, the in-394

termolecular Au–Au distances were shortened significantly;395

they were even shorter than the intramolecular Au–Au dis-396

tances in which ligand assistance plays a role. This is ascrib-397

able to charge-transfer from the electron-rich Au center to398

the electron acceptor TCNQ; a partial oxidation of the Au(I)399

atoms leads to a shortening of Au–Au distances. In the lim-400

iting case, upon oxidation to Au(II), a gold–gold single bond401

forms[51]. Second, the intense dark green color of the adduct,402

in contrast to the colorless crystals of1 and the pale brown403

crystals of TCNQ, is an indication of charge transfer; similar404

color changes were used as evidence of charge transfer in405

a previous studies by Balch for adducts of trinuclear Au(I)406

carbeniate complexes with nitro-9-fluorenones[21b]. Like-407

wise, Perutz and co-workers reported a red charge-transfer408

complex that forms as a result of a reaction between bis(�6-409

benzene)chromium and hexafluorobenzene[52]. There is410

more relevance for the latter study with the2·TCNQ adduct411

discussed here than with the1·C6F6 adduct discussed in the412

previous section. The charge transfer adduct forming between413

2 and TCNQ is believed to remain intact in solution because414

the solutions of the adduct are deeply colored.Fig. 11shows415

the absorption spectrum of the adduct in a CH2Cl2 solution.416

T ab-417

s er be-418

c tral419

T er420

a421

p -422

s all423

a -424

s cted425

t orp-426

tions due to the neutral form and the fully reduced anionic427

TCNQ− form. Third, a scanning-electron microscope (SEM)428

image for crystals of2·TCNQ is shown in the inset ofFig. 11. 429

The observation of a clear SEM image for crystals of2·TCNQ 430

that were not coated with a conducting material suggests that431

these crystals are non-insulators, because the charge from the432

electron beam accumulates on the surface of insulating mate-433

rials, which precludes the observation of a clear SEM image434

for an insulator[53]. Conducting or semiconducting materi-435

als, on the other hand, can diffuse the charge and, thus, allow436

for the observation of a clear SEM image. We have attempted437

to carry out preliminary resistivity measurements for a large438

crystal of2·TCNQ that we were able to grow from a saturated439

CH2Cl2 solution containing1 and TCNQ at 4◦C. However, 440

the resistivity of the crystals at ambient temperature was too441

high to be measured with the setup we had access to, which442

can only measure resistivities for conducting materials. The443

apparent optical band gap of2·TCNQ (the absorption edge) 444

is about 1.4 eV, which is within the band gap region for semi-445

conductors, not conductors. 446

4. Conclusions 447

Although metal centers are usually electrophilic, we have448

shown via solid evidence based on X-ray crystallography,449

e , and450

q u(I)451

c - 452

o with453

a oto-454

p ding455

o itiza-456

t s via457

a ered458

e ac-459

UN

CO

RR

E

he spectrum shows visible and NIR absorptions. Theorption peak near 500 nm must be due to charge transfause2 and TCNQ alone absorb below 400 nm. Both neuCNQ and TCNQ− absorb in the NIR region with strongbsorptions for the anionic form (εmax= 43,300 M−1 cm−1);artially reduced TCNQ will likely exhibit similar NIR aborptions. The spectrum inFig. 11appears to show a smmount of anionic TCNQ− but it is hard to distinguish aborptions due to the partially reduced TCNQ that are expeo be the dominant TCNQ species in solution from abs

CCR 110408 1–1

lectronic absorption and luminescence spectroscopyuantum mechanical calculations that the trinuclear Aomplexes studied herein are strong� bases. Their nuclephilic nature allows them to form acid–base adductsvariety of organic and inorganic electrophiles. The ph

hysical properties vary in the different adducts depenn the interacting electrophile. These included the sens

ion of the phosphorescence of the aromatic electrophilegold heavy-atom effect, the quenching of the Au-centmission by disrupting the Au–Au intermolecular inter

Page 9: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

ED

OF

0

M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx 9

tions by the organic electrophile, the formation of intense460

charge transfer absorptions in the visible and NIR regions,461

and metal-centered thermochromic luminescence bands de-462

localized along the stacking M–M chain axis when the elec-463

trophiles contains a heavy metal. These fascinating photo-464

physical properties promise a great potential for the utiliza-465

tion of such supramolecular acid–base stacks in a variety466

of optoelectronic applications that include molecular LEDs,467

selective optical sensing of hazardous small molecules and468

heavy metals, optical telecommunication devices, and solar469

cell dyes.470

Acknowledgments471

The financial support of this project has been provided by472

the Robert A. Welch Foundation of Houston, TX, to J.P.F.473

and M.A.O., a National Science Foundation career award to474

M.A.O. (CHE-0349313), and an Advanced Technology Pro-475

gram grant from the Texas Higher Education Coordinating476

Board to M.A.O. We also thank Professor Alfredo Burini of477

the University of Camerino for initiating collaborative studies478

with the trinuclear Au(I) complexes.479

R480

481

les,482

483

484

485

Cor-486

487

ory488

don,489

490

ork,491

492

.A.493

58.494

ples,495

496

ry,497

Am.498

499

.A.500

Soc.501

502

assi,503

504

em.505

506

[ 3.507

[508

[ 1–3,509

510

[ .511

512

[ R.513

514

[15] (a) E.J. Fernandez, J.M. Lopez-de-Luzuriaga, M. Monge, M.E. Ol-515

mos, J. Perez, A. Laguna, A.A. Mohamed, J.P. Fackler Jr., J. Am.516

Chem. Soc. 125 (2003) 2022; 517

(b) E.J. Fernandez, J.M. Lopez-de-Luzuriaga, M. Monge, M. Mon- 518

tiel, M.E. Olmos, J. Perez, A. Laguna, F. Mendizabal, A.A. Mo- 519

hamed, J.P. Fackler Jr., Inorg. Chem. 43 (2004) 3573–3581. 520

[16] W.R. Caseri, H.D. Chanzy, K. Feldman, M. Fontana, P. Smith, T.A.521

Tervoort, J.G.P. Goossens, E.W. Meijer, A.P.H.J. Schenning, I.P. Dol-522

byna, M.G. Debije, M.P. de Haas, J.M. Warman, A.M. van de Craats,523

R.H. Friend, H. Sirringhaus, N. Stutzmann, Adv. Mater. 15 (2003)524

125. 525

[17] (a) H. Tanaka, Y. Okano, H. Kobayashi, W. Suzuki, A. Kobayashi,526

Science 291 (2001) 285; 527

(b) F. Bigoli, P. Deplano, M.L. Mercuri, M.A. Pellinghelli, L. Pilia, 528

G. Pintus, A. Serpe, E.F. Trogu, Inorg. Chem. 41 (2002) 5241; 529

(c) N. Robertson, L. Cronin, Coord. Chem. Rev. 227 (2002) 93; 530

(d) H. Kisch, B. Eisen, R. Dinnebier, K. Shankland, W.I.F. David,531

F. Knoch, Chem. Eur. J. 7 (2001) 738. 532

[18] B.W. Smucker, J.M. Hudson, M.A. Omary, K.R. Dunbar, Inorg.533

Chem. 42 (2003) 4714. 534

[19] M.A. Rawashdeh-Omary, M.A. Omary, J.P. Fackler Jr., Inorg. Chim.535

Acta 334 (2002) 376. 536

[20] M. Stender, R.L. White-Morris, M.M. Olmstead, A.L. Balch, Inorg.537

Chem. 42 (2003) 4504. 538

[21] (a) J.C. Vickery, M.M. Olmstead, E.Y. Fung, A.L. Balch, Angew.539

Chem. Int. Ed. Engl. 36 (1997) 1179; 540

(b) M.M. Olmstead, F. Jiang, S. Attar, A.L. Balch, J. Am. Chem.541

Soc. 123 (2001) 3260; 542

(c) A. Hayashi, M.M. Olmstead, S. Attar, A.L. Balch, J. Am. Chem.543

Soc. 124 (2002) 5791. 544

[22] G. Yang, R.G. Raptis, Inorg. Chem. 42 (2003) 261. 545

[ 4 546

547

[ .P.548

549

[ .A.550

. 551

[ m.552

553

[ Jr.,554

555

[ istry556

557

[ Sci.558

559

[ J.P.560

561

[ kler562

563

[ 2. 564

[ 977)565

566

[ . 567

[ 568

[ 968)569

570

[ 31.571

[ ngs,572

573

[ of574

969.575

[ 954)576

577

[ 578

[ 579

[ sity580

581

UN

CO

RR

EC

T

eferences

[1] G.N. Lewis, J. Franklin Inst. 226 (1938) 293;G.N. Lewis, Valence and the Structure of Atoms and MolecuChemical Catalog Co., New York, 1978, p. 141.

[2] (a) C.K. Ingold, J. Chem. Soc. (1933) 1120;C.K. Ingold, Chem. Rev. 15 (1934) 225;C.K. Ingold, Structure and Mechanism in Organic Chemistry,nell, Ithaca, New York, 1953, Chapter V ;(b) R. Robinson, Outline of an Electrochemical (Electronic) Theof the Course of Organic Reactions, Institute of Chemistry, Lon1932, p. 12.

[3] W.B. Jensen, The Lewis Acid–base Concepts, Wiley, New Y1980.

[4] A. Burini, R. Bravi, J.P. Fackler Jr., R. Galassi, T.A. Grant, MOmary, B.R. Peitroni, R.J. Staples, Inorg. Chem. 39 (2000) 31

[5] A. Burini, J.P. Fackler Jr., R. Galassi, B.R. Pietroni, R.J. StaChem. Commun. (1998) 95.

[6] A. Burini, J.P. Fackler Jr., R. Galassi, A. Macchioni, M.A. OmaM.A.-O. Rawashdeh, B.R. Pietroni, S. Sabatini, C. Zuccaccia, J.Chem. Soc. 124 (2002) 4570.

[7] A. Burini, J.P. Fackler Jr., R. Galassi, T.A. Grant, M.A. Omary, MRawashdeh-Omary, B.R. Pietroni, R.J. Staples, J. Am. Chem.122 (2000) 11264.

[8] M.A. Rawashdeh-Omary, M.A. Omary, J.P. Fackler Jr., R. GalB.R. Pietroni, A. Burini, J. Am. Chem. Soc. 123 (2001) 9689.

[9] A. Burini, A.A. Mohamed, J.P. Fackler Jr., Comments Inorg. Ch24 (2003) 253.

10] J.K. Bera, K.R. Dunbar, Angew. Chem. Int. Ed. 41 (2002) 44511] R. Hoffmann, Angew. Chem., Int. Ed. Engl. 26 (1987) 846.12] J.S. Miller (Ed.), Extended Linear-Chain Compounds, vols.

Plenum Press, New York, 1982.13] Y. Kunugi, K.R. Mann, L.L. Miller, C.L. Exstrom, J. Am. Chem

Soc. 120 (1998) 589.14] M.A. Mansour, W.B. Connick, R.J. Lachicotte, H.J. Gysling,

Eisenberg, J. Am. Chem. Soc. 120 (1998) 1329.

PR

O

CCR 110408 1–1

23] M.R. Haneline, M. Tsunoda, F.P. Gabbaı, J. Am. Chem. Soc. 12(2002) 3737.

24] M.A. Omary, R.M. Kassab, M.R. Haneline, O. Elbjeirami, FGabbaı, Inorg. Chem. 42 (2003) 2176.

25] H.V.R. Dias, H.V.K. Diyabalanage, M.A. Rawashdeh-Omary, MFranzman, M.A. Omary, J. Am. Chem. Soc. 125 (2003) 12072

26] A. Tiripicchio, M.T. Camellini, G. Minghetti, J. Organomet. Che171 (1979) 399.

27] M.A. Rawashdeh-Omary, M.A. Omary, L. Daniels, J.P. Facklerin preparation.

28] H. Schmidbaur (Ed.), Gold: Progress in the Chemistry, Biochemand Technology, John Wiley & Sons, Chester, England, 1999.

29] S. Mecozzi, A.P. West, D.A. Dougherty, Proc. Natl. Acad.U.S.A. 93 (1996) 10566, and references therein.

30] M.A. Omary, J. Park, M.A. Rawashdeh-Omary, S.W. North,Fackler Jr., in preparation.

31] A.A. Mohamed, M.A. Rawashdeh-Omary, M.A. Omary, J.P. FacJr., in preparation.

32] Y. Xie, H.F. Schaefer, F.A. Cotton, Chem. Commun. (2003) 1033] J.C. Miller, J.S. Meek, S.J. Strickler, J. Am. Chem. Soc. 99 (1

8175, and references therein.34] H.B. Rosenberg, S.D. Carson, J. Phys. Chem. 72 (1968) 353135] V.L. Ermolaev, K.J. Svitashev, Opt. Spektrosk. 7 (1959) 664.36] E.F. Kiefer, W.L. Waters, D.A. Carlson, J. Am. Chem. Soc. 90 (1

5127.37] W. Kitching, G.M. Drew, V. Alberts, Organometallics 1 (1982) 338] N.J. Turro, Modern Molecular Photochemistry, Benjamin/Cummi

Menlo Park, CA, 1978.39] S.P. McGlynn, T. Azumi, M. Kinoshita, Molecular Spectroscopy

the Triplet State, Prentice-Hall: Englewood Cliffs, New Jersey, 140] D.S. McClure, N.W. Blake, P.L. Hanst, J. Chem. Phys. 22 (1

255.41] V.L. Ermolaev, K.J. Svitashev, Opt. Spektrosk. 7 (1959) 399.42] M. Kasha, J. Chem. Phys. 20 (1952) 71.43] J.S. Griffith, Theory of Transition Metal Ions, Cambridge Univer

Press, Cambridge, 1964.

Page 10: Photophysics of supramolecular binary stacks consisting of electron-rich trinuclear Au(I) complexes and organic electrophiles

DO

F

10 M.A. Omary et al. / Coordination Chemistry Reviews xxx (2005) xxx–xxx

[44] J. Friedrich, F. Metz, F. Dorr, Mol. Phys. 30 (1975) 289.582

[45] H-Y. Chao, W. Lu, Y. Li, M.C.W. Chan, C.-M. Che, K.-K. Cheung,583

N. Zhu, J. Am. Chem. Soc. 124 (2002) 14696.584

[46] M.A. Omary, M.A. Rawashdeh-Omary, H.V.K. Diyabalanage, H.V.R.585

Dias, Inorg. Chem. 42 (2003) 8612.586

[47] For reviews, see;587

(a) J.M. Forward, J.P. Fackler Jr., Z. Assefa, Photophysical and588

photochemical properties of gold(I) complexes, in: D.M. Roundhill,589

J.P. Fackler Jr. (Eds.), Optoelectronic Properties of Inorganic Com-590

pounds, Plenum, New York, 1999, Chapter 6 ;591

(b) V.W.-W. Yam, K.K. Lo, Chem. Soc. Rev. 28 (1999) 323.592

[48] (a) H. Kisch, B. Eisen, R. Dinnebier, K. Shankland, W.I.F. David,593

F. Knoch, Chem. Eur. J. 7 (2001) 738;594

(b) H. Kisch, Comments Inorg. Chem. 16 (1994) 113.595

[49] A comprehensive citation about the mechanism of the external596

heavy-atom effect is inappropriate but representative examples597

include [42] and the following with references therein;598

(a) T. Azumi, Chem. Phys. Lett. 19 (1973) 580;599

(b) A.K. Chandra, N.J. Turro, A.L. Lyons Jr., P. Stone, J. Am.600

Chem. Soc. 100 (1978) 4964;601

602

(c) B.F. Minaev, S. Knuts, H.Agren, Chem. Phys. Lett. 181 (1994) 603

15; 604

(d) N. Chattopadhyay, C. Serpa, M.M. Pereira, J. Seixas de Melo, L.G.605

Arnaut, S.J. Formosinho, J. Phys. Chem. A 105 (2001) 10025; 606

(e) I.R. Gould, J.A. Boiani, E.B. Gaillard, J.L. Goodman, S. Farid, J.607

Phys. Chem. A 107 (2003) 3515. 608

[50] D. Phillips, J. Chem. Phys. 46 (1967) 4679. 609

[51] For example, see; 610

(a) T.F. Carlson, J.P. Fackler Jr., J. Organomet. Chem. 596 (2000)611

237; 612

(b) V.W.W. Yam, S.W.K. Choi, K.K. Cheung, Chem. Commun.613

(1996) 1173; 614

J.P. Fackler Jr., Inorg. Chem. (2002) 6959. 615

[52] C.J. Aspley, C. Boxwell, M.L. Buil, C.L. Higgitt, C. Longb, R.N. 616

Perutz, Chem. Commun. (1999) 1027. 617

[53] J.I. Goldstein, D.E. Newbury, P. Echlin, D.C. Joy, A.D. Romig Jr.,618

C.E. Lyman, C. Fiori, E. Lifshin, Scanning Electron Microscopy and619

X-ray Microanalysis: A Text for Biologists, Materials Scientists, and620

Geologists, Plenum Press, New York, 1992. 621

TE

UN

CO

RR

EC

PR

O

CCR 110408 1–10