Nano Letters - ACS Publications - American Chemical Society

19 hours ago - Air Pollution and Industrial Hygiene · Apparatus and Plant Equipment ... The mechanisms of chemical reactions, including the transforma...
0 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Communication

Label-Free Dynamic Detection of SingleMolecule Nucleophilic Substitution Reaction Chunhui Gu, Chen Hu, Ying Wei, Dongqing Lin, Chuancheng Jia, Mingzhi Li, Dingkai Su, Jianxin Guan, Andong Xia, Linghai Xie, Abraham Nitzan, Hong Guo, and Xuefeng Guo Nano Lett., Just Accepted Manuscript • DOI: 10.1021/acs.nanolett.8b00949 • Publication Date (Web): 06 Jun 2018 Downloaded from http://pubs.acs.org on June 6, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Label-Free Dynamic Detection of Single-Molecule Nucleophilic Substitution Reaction Chunhui Gu,1† Chen Hu,2† Ying Wei,3 Dongqing Lin,3 Chuancheng Jia,1 Mingzhi Li,1 Dingkai Su,1 Jianxin Guan,1 Andong Xia,4 Linghai Xie,3 Abraham Nitzan,5 Hong Guo,2* Xuefeng Guo 1* 1

Beijing National Laboratory for Molecular Sciences, State Key Laboratory for Structural

Chemistry of Unstable and Stable Species, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871, P. R. China; 2

Center for the Physics of Materials and Department of Physics, McGill University,

Montreal, Quebec H3A 2T8, Canada; 3

Center for Molecular Systems and Organic Devices, Key Laboratory for Organic

Electronics & Information Displays and Institute of Advanced Materials, Nanjing University of Posts & Telecommunications, Nanjing 210023, P. R. China; 4

Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China;

5

Department of Chemistry, University of Pennsylvania, Philadelphia, PA 19104-6323,

USA;

ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT: The mechanisms of chemical reactions, including the transformation pathways of the electronic and geometric structures of molecules, are crucial for comprehending the essence and developing new chemistry. However, it is extremely hard to realize at the single-molecule level. Here we report a single-molecule approach capable of electrically probing stochastic fluctuations under equilibrium conditions and elucidating time trajectories of single species in nonequilibrated systems. Through molecular engineering, a single molecular wire containing a functional center of 9-phenyl-9-fluorenol was covalently wired into nanogapped graphene electrodes to form stable single-molecule junctions. Both experimental and theoretical studies consistently demonstrate and interpret direct measurement of the formation dynamics of individual carbocation intermediates with a strong solvent dependence in a nucleophilic substitution reaction. We also show the kinetic process of competitive transitions between acetate and bromide species, which is inevitably through a carbocation intermediate, confirming the classical mechanism. This unique method opens up plenty of opportunities to carry out single-molecule dynamics or biophysics investigations in broad fields beyond reaction chemistry through molecular design and engineering. 1 2

KEYWORDS: Single-molecule detection, reaction dynamics, molecular electronics,

3

nucleophilic substitution

4 5 6 7

ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

8

Revealing the dynamic processes of detailed molecular transformations of chemical

9

reactions is crucial for comprehending the essence and developing new chemistry.13

10

Previous reports proved that the pathway of chemical reactions seemed to be more complex

11

than we imagine. For instance, even for a simple bimolecular nucleophilic substitution (SN2)

12

reaction, the mechanism is complicated and not fully understood.2,3 In general, it is

13

extremely hard to access the detailed reaction pathways in macroscopic experiments

14

because the ensemble always shows thermodynamic quasi-equilibrium conditions. In

15

contrast, single-molecule analysis of chemical reactions can effectively avoid ensemble-

16

average effects, discover new phenomena/species and reveal the chronological reaction

17

processes.4 To date, discrete single-molecule detection technologies, including optical

18

methods57 and nanopores,810 have been developed to achieve the dynamic investigation

19

of biological macromolecules. However, these techniques seem to be unsuitable for the

20

detection of general organic molecules and their reaction trajectories, mainly limited by the

21

challenges including the lack of precise molecular immobilization, the requirement of

22

fluorescent labelling and low temporal resolution. Therefore, the development of robust

23

single-molecule detection platforms with label-free capability is invaluable to offer plenty

24

of room for probing molecular mechanisms of basic chemical reactions.

25

In this regard, the electrical approach may be a suitable choice.11,12 In particular,

26

electrical platforms based on single-molecule junctions (SMJs) are potentially attractive

27

because these platforms might overcome the major challenges mentioned above. In the past

28

two decades, different approaches to build SMJs,13 particularly mechanically controllable

29

break junction14 and conductive scanning probe microscope,1517 were developed, which

30

have made remarkable contributions to investigate the electronic properties of particular 3 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

31

molecules. In our previous studies, we developed a reliable methodology to build a new

32

type of SMJs based on graphene electrodes,18 where amine-terminated molecular wires can

33

be immobilized in the nanogap between carboxylic acid-terminated graphene electrodes

34

through covalent amide bonds. Such molecular nanocircuits benefit from the unique

35

coupling modes and robust contacts at the electrode-molecule interface, resulting in the

36

strong stability during long-term conductance measurements. As a single molecule

37

sandwiched between source and drain electrodes is the key contributor to the device

38

conductance, the conductance of SMJs is generally ultrasensitive to the electronic

39

structures of molecules. Through molecular engineering, specially-designed molecules

40

were integrated into SMJs to construct molecular electronic devices with specific functions

41

such as rectifiers,1921 field-effect transistors,22 switches23 and memories.24 These examples

42

demonstrate the capacity for in-situ modulating the electronic structure (and thus the

43

conductance) of SMJs by external stimuli, where the rearrangement of the molecular

44

electronic structures caused by chemical reactions can be deemed as possible external

45

stimuli.25 On the basis of this strategy, SMJs have been specially designed through

46

structure-function relationships and have successfully detected chemical reactions such as

47

complex formation26 and nucleophilic addition.27

48

In the present study, for the first time we test the potential of the utilization of SMJs

49

as an electrical platform for direct dynamic measurement of a reversible unimolecular

50

nucleophilic substitution (SN1) reaction at the single-molecule level. As a well-known

51

organic reaction, the classical pathway of a SN1 reaction involves two steps: (i) Heterolysis:

52

A leaving group separates from the reactant to form a carbocation intermediate. (ii)

53

Recombination: The carbocation intermediate combines with a nucleophile to form a 4 ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

54

product. The reaction potential energy surface is shown in Figure 1 (up-right). Thereinto,

55

carbocation, as a short-lived but common intermediate, has attarcted great interest in the

56

field of organic chemistry, in which, especially, Olah, G. A. made great contributions.28,29

57

We expect that the carbocation intermediate can be distinguished from the reactant/product

58

electrically due to the conductance increase caused by the transition from sp3 to sp2

59

hybridization, which has been reported previously.30 By real-time monitoring SMJs with a

60

9-phenyl-9-fluorenol center, we found that the conductance of SMJs was faithfully

61

synchronous to the chemical reaction, through which we observed the reversible SN1

62

reaction and its closely-following competitive reactions, respectively. Statistical analyses

63

in time domain further revealed significant information about the reaction kinetics, which

64

also provides the useful guidance to other chemical reactions.

65

Specially, 9-phenyl-9-fluorenol was applied as a reactant to investigate the acid-

66

catalyzed SN1 reaction in the mixed solution of acetic acid (HAc) and trifluoroacetic acid

67

(TFA). Such a mixed acid solution can provide the high proton environment to stabilize

68

the carbocation intermediate, which is favorable to further detection. This particular

69

reaction was first investigated by a series of macroscopic experiments. GC-MS confirmed

70

the complete consumption of the reagent and observed a reversible transition between a 9-

71

phenyl-9-fluorenyl cation and a 9-phenyl-9-fluorenyl acetate in the solution. The kinetic

72

and thermodynamic properties of the reaction were determined by UV-Vis experiments

73

and flash photolysis (see Supporting Information, Section 1). Through molecular

74

engineering, we covalently integrated a molecular wire with a functional center of 9-

75

phenyl-9-fluorenol into nanogapped graphene electrodes to form stable graphene-

76

molecule-graphene single-molecule junctions (GMG-SMJs) (Figure 1). The details of 5 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

77

molecular synthesis are provided in the Supporting Information (Section 2). To confirm

78

the formation of GMG-SMJs, the current-voltage (IV) curves of the devices at different

79

stages were measured as shown in Figure S9a. We found that the current decreased to zero

80

after precise oxygen plasma etching and recovered to some extent after molecular

81

immobilization, indicating the success of the device fabrication. Under optimized

82

conditions, the connection yield was found to be ~15%, that is, 25 of 169 devices on the

83

same silicon chip showed the increased conductance. These working devices showed

84

similar electrical properties in the following experiments, demonstrating the

85

reproducibility. On the basis of these data, statistical analysis demonstrated that charge

86

transport through the junction mainly resulted from single-molecule connection (see

87

Supporting Information, Section 3).

88 89 90

Figure 1. Schematic of a graphene-molecule-graphene single-molecule junction that shows the formation dynamics of a 9-phenyl-9-fluorenyl cation in a SN1 reaction. 6 ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

91

High temporal resolution electrical characterization was carried out to monitor the

92

conductance of GMG-SMJs in real time (Figure 2. Constant bias voltage: 0.3 V; Sampling

93

rate: 57.6 kSa/s). GMG-SMJs were initially measured in the air and then anhydrous

94

HAc/TFA solutions with different proportions were gradually added to GMG-SMJs with

95

the aid of a polydimethylsiloxane (PDMS) solvent reservoir (Figure S9b). Figures 2a and

96

2b show the representative current-time (It) trajectories. In comparison with the It

97

trajectory in the air, which was dominated by the flicker (1/f) noise, random telegraph

98

signals (RTSs) appeared when acid solutions were added. Correspondingly, in the current

99

distribution histograms (Figure 2c), we observed large-amplitude two-level conductance

100

states: a large numbered, narrowly broadened low conductance state (shown in blue) and a

101

less numbered, widely broadened high conductance state (shown in red). This observation

102

demonstrated that GMG-SMJs had two stable states in acid solutions. A previous report

103

demonstrated that the current itself might lead to the reconnection of graphene electrodes

104

at high bias voltages (~1.2 V for ~3 nm nanogaps, similar to the condition in this work).31

105

To rule out this possibility, we fixed an optimized bias voltage at 0.3 V (much lower than

106

1.2 V used in the literature) to maintain a stable conductance as well as a high signal-noise

107

ratio. Under such a condition, we proved that the RTS did not appear in the air and the RTS

108

behavior was highly related to the acid solution, demonstrating that RTS signals are

109

solvent-dependent, rather than the behavior of graphene electrodes. To further rule out

110

other potential artifacts, three different charge-transporting systems were established as

111

control devices: (i) Macroscopic uncut graphene ribbons,(ii) Partially-cleaved graphene

112

nanoconstrictions and (iii) SMJs bridged by a molecular wire containing an acid-inert

113

sexiphenyl core, which has the similar molecular length and conductance with the 97 ACS Paragon Plus Environment

Nano Letters

114

phenyl-9-fluorenol core, were applied to exclude the possibility of conformation-induced

115

switching or the other systemic interference between the solvent and SMJs. Under the same

116

testing conditions used in Figure 2, RTS fluctuations did not appear in all the It trajectories

117

(see Supporting Information, Section 4), proving that the observed RTSs only originated

118

from the chemical reaction happening on the 9-phenyl-9-fluorenol core.

a

b

Current (nA)

Air

Current (nA)

0

0.2

0.4

0.6

1.0

0.8

110

80

80

50

50

Current (nA)

Current (nA)

0.6

1.0

0.8

20 0.967

0

40

80

85 55 0

50 0.968

0

90

60

60

30

30 0.6

1.0

0.8

0.20

0.5 ms

40 0

0.215

0

0.23

400

400 300

200

200

200

200

100

100 0 0.42

1.0

0.5 ms

0.425

0.44

0

0

400

400

400

300

300

200

200

200

100

100

100

0

0.2

0.4 0.6 Time (s)

1.0

0.8

30

7000

0

15

30

100

300

0 0.05

15

3500

300

0.8

2400

30

400

0.6

30

80

60

300

0.4

15

120

400

0.2

3600

1200

300

0

30

20 0.9675

90

0.4

15

1800 115

90

0.2

20 0

110

120

0

120 121 122 123 124 125

0.4

0.5 ms

0

120

0

119

0.2

0.60

0.575

120

0

75% TFA 25% HAc

0 0.55

110

0

50% TFA 50% HAc

20

20

20

20

25% TFA 75% HAc

40

30

0

0% TFA 100% HAc

c 40

40

Current (nA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 30

0.075 Time (s)

0.10

0

0

6000

2500 Counts

12000

5000

Figure 2. Real-time conductance measurements of a representative GMG-SMJ among 5 different devices. (a) Representative It trajectories in different conditions (from top to bottom: in the air, pure HAc, 25% TFA/75% HAc, 50% TFA/50% HAc, and 75% TFA/25% HAc (vol/vol)). (b) Partial It curves of the corresponding parts in (a). Insets show the enlarged views of the peaks as indicated by white arrows. (c) The corresponding histograms of (a). The blue and red lines show the gauss-shaped fit for the low and high 8 ACS Paragon Plus Environment

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

126 127 128

Nano Letters

conductance states, respectively. Insets show the enlarged views in the high conductance regions. Source-drain bias voltage (VD): 0.3 V; Gate voltage (VG): 0 V; Sampling rate: 57.6 kSa/s.

129 130

To attribute the conductance states to the corresponding molecular forms, the

131

molecular electronic structures and quantum transport properties were theoretically

132

analyzed. As shown in Figures 3a and 3b, the first-principles calculations performed by

133

Gaussian package (See Supporting Information, Section 5) showed totally different

134

electronic structures between acetate and carbocation forms, due to the variation of the

135

hybridization types of the central carbon atom: sp3 hybridization in the former and sp2

136

hybridization in the latter. As a result, the valence electrons are much more conjugated in

137

the carbocation, yielding the fact that the energy gap between the frontier molecular

138

orbitals (FMOs) is significantly decreased to ~1.34 eV (Figure 3a). Different energy gaps

139

and orbital distributions can be verified by the different UV-Vis spectroscopic behaviors

140

in Figure S15. According to the transition voltage spectroscopy (TVS) model and the

141

theoretical approach by Bâldea et al.,3234 the decreased FMO gap of the carbocation form

142

can largely lower the tunneling barrier height and increase the tunneling current, leading

143

to a higher conductance. We further calculated the charge transport properties of these

144

molecules in combination with graphene probes by using the density functional theory

145

(DFT) within the nonequilibrium Green’s function (NEGF) technique (See Supporting

146

Information, Section 5). As shown in Figure 3c, in comparison with the acetate form, for

147

the carbocation form the perturbed lowest unoccupied molecular orbital (p-LUMO) and

148

perturbed highest occupied molecular orbital (p-HOMO) have transmission peaks with

149

energy much closer to the unbiased electrode chemical potentials (EF). In the resonant 9 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

150

tunneling mechanism described by the Landauer-Büttiker formalism, the energy gap

151

between the p-FMOs and the EF plays a crucial role in the current calculation under a low

152

bias voltage: The smaller the gap is, the more electrons can enter the Fermi window in the

153

bias range explored.23,35,36 In addition to the p-FMOs, we also notice that for the

154

carbocation form the p-HOMO-1 is also close to the EF with a distinguished transmission

155

peak, which can contribute to a large conductance. Therefore, with the same results

156

analyzed by the above-mentioned two methods and the quantum transport calculations, we

157

can conclude that the carbocation form should have a higher conductance than the acetate

158

form under a low bias voltage. The calculated scattering states of these critical transmission

159

p-FMOs are consistent with the FMOs (Figure 3b), indicating that the transmission

160

channels are mainly derived from the electronic orbitals of the molecules (Figure S16). It

161

was found that carbocation conductance varied along with solution conditions. Recent

162

studies reported that the dipole-dipole interaction between the measured molecule and

163

solution molecules might affect molecular conductance.21,37 Considering the fact that HAc

164

and TFA have the big difference in dielectric constant (~6 for acetic acid and ~39 for TFA),

165

the different polar intermolecular interactions should affect electron transport of molecular

166

junctions to some extent, thus causing the conductance variation. Another possibility is that

167

different solutions could influence the lifetime of carbocation. Due to the measurement

168

limit of the instruments used, the transition between the carbocation and acetate forms

169

could be too fast to follow without fully reaching the conductance platform of carbocation,

170

as demonstrated in Figure 2 in mixed solutions (pure HAc and 25% TFA/75% HAc).

10 ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30

a

b

0 –1

–1.71

–1.66

–5.50

–5.42

Ac

Br

–2

E (eV)

–3

LUMO

–3.64

–4 –5 –6

HOMO –4.98

R = (+)

–7 R = (+)

c Transmission

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

p-HOMO

R = Br p-LUMO p-LUMO

p-HOMO

10 0

p-HOMO-1 p-HOMO

10 –4

p-LUMO

10 –8 R= (+) R= Ac R= Br

10 –12

10 –16 –2.0

171 172 173 174 175 176 177

R = Ac

–1.5

–1.0

–0.5

0 0.5 E–EF (eV)

1.0

1.5

2.0

Figure 3. Theoretical simulations of GMG-SMJs. (a) Calculated energy levels of the molecular wire in the three forms: R = (+) (Carbocation, red), R = Ac (acetate, blue) and R = Br (Bromide, orange). The dash line shows the Fermi level of graphene electrodes: EF = –4.8 eV. (b) Corresponding calculated frontier molecular orbitals (FMOs) of the molecular wire in the three forms. (c) Transmission spectra of GMG-SMJs in the three forms around the Fermi level of graphene at a zero-bias voltage.

178 179

To analyze the reaction dynamics, RTSs in It trajectories were idealized into a two-

180

level interconversion by using a QuB software (Figure 4a). In this analysis, the lifetimes

181

(τlow/τhigh) of the acetate/carbocation forms are derived from the probability distributions of

182

the dwell times (Tlow/Thigh) of the low/high conductance states. By taking the device in

183

Figure 2 as an example, both of the probability distributions of Tlow and Thigh in the solution

184

of 75% TFA/25% HAc (vol/vol) were well fit by a single exponential decay function as

185

shown in Figures 4b and 4c, which produce the values of τlow = 5320 ± 790 μs and τhigh =

186

2540 ± 200 μs, respectively, according to a hidden Markov chain model. On the basis of

187

this fact, we initially assumed that the reversible reaction could be regarded as a simple

188

Poisson process. Their rate constants (kd and kc), corresponding to the heterolytic 11 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

189

dissociation of the acetate form (kd = 1/τlow) and the combination of the carbocation form

190

with acetic acid (kc = 1/τhigh) (insets in Figures 4b and 4c), were derived to be (1.88 ±0.28)

191

× 102 s–1 and (3.94 ± 0.29) × 102 s–1, respectively. More data in different conditions are

192

shown in the Supporting Information (Section 6). Figure 4d (solid lines) shows the

193

corresponding activation energies (Ed and Ec) calculated from the Eyring equation E =

194

RTln(kBT/hk), where R = 8.314 J/(mol·K), T is the temperature, kB is the Boltzmann

195

constant, and h is the plank constant. These values show the same tendency in comparison

196

with the macroscopic results measured by UV-Vis experiments and flash photolysis (dash

197

lines). It was worthy to notice that these statistics were proceeded from a long-term

198

observation of one molecule rather than the observation of the ensemble, and thus the

199

results were defined in a time domain, which should be comparable to the macroscopic

200

results due to the equivalence between time average and ensemble average.

12 ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

201 202 203 204 205 206 207 208 209 210 211 212 213

Nano Letters

Figure 4. Statistical analyses of the GMG-SMJ device used in Figure 2. (a) Measured It trajectories (black dots) of a GMG-SMJ in the solution of HAc-TFA (25%/75%, vol/vol). The red line shows the idealised two-level interconversion by using a QUB software. (b,c) Plots of time intervals of the low (b) and high (c) states derived from It trajectories of a GMG-SMJ in the solution of HAc-TFA (25%/75%, vol/vol). The distributions were well fitted by a single-exponential decay function (blue lines). Insets show the corresponding reaction pathways. (d) Activation energies of the symmetric SN1 reaction derived from macroscopic (dash lines) and single-molecule (solid lines) experiments. (e) Combination constants of carbocation (pKR+). Column: pKR+ values derived from GMG-SMJs. Solid columns represent the respective pK values; Striped columns represent the generalized acid function J0 in the corresponding solutions. The blue and black dash lines show the J0TFA% relationship and the pKR+ values derived from bulk experiments, respectively.

214 215

Single-molecule “thermodynamic functions” should be redefined based on kinetic

216

constants derived from GMG-SMJ experiments. The equilibrium constant (K) is defined

217

as the division of the rate constants of the dissociation/combination reactions (K = kd/kc).

218

In the solution containing TFA proportions from 0% to 75%, the pK values (p means the

219

minus denary logarithm) were calculated to be 3.57 ± 0.11 (0% TFA), 2.55 ± 0.21 (25% 13 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

220

TFA), 1.73 ±0.06 (50% TFA) and 0.32 ±0.20 (75% TFA), respectively. The combination

221

constants (KR+) between the carbocation form and acetic acid were derived from the pK

222

values normalized by a carbocation-based generalized acid function (J0) (Figure 4e, blue

223

line and striped columns), which was measured by an indicator overlap method: pKR+ = J0

224

+ pK. The pKR+ values were calculated as −4.10 ± 0.11 (0% TFA), −8.11 ± 0.21 (25%

225

TFA), −9.46 ± 0.06 (50% TFA) and −9.72 ± 0.20 (75% TFA) (Figure 4e, solid columns),

226

respectively. In comparison with the macroscopic results pKR+ = ~10.50 (see Supporting

227

Information, Section 1), it seems that GMG-SMJs may improve the stability of the

228

carbocation state more or less. Two possibilities could explain this finding: (i) In

229

comparison with isolated molecules, graphene electrodes extend the conjugation range of

230

carbocation, which obviously lower the Columbic potential of carbocation; (ii)

231

Electrostatic catalysis may take effect in GMG-SMJs, that is, the strong electro-static field

232

between graphene electrodes (Es = ~0.3 V/nm) increases the stability of charge-separated

233

intermediates. In brief, the dipole (μ) of molecules is more likely to array parallel to the

234

electrostatic field, thus generating a field-induced stabilization energy μ × Es.38,39 As the

235

carbocation-anion pair is charge separated and much more polar than the acetate, it is no

236

doubt that the electrostatic field generates more stabilization energy in the carbocation

237

intermediate. Considering that the external electric field can be “counteracted” by the

238

directional alignment of solution molecules, the effect of electrostatic catalysis depends on

239

the polarity of solution. Since TFA is much more polar than HAc, it is reasonable that the

240

effect of electrostatic catalysis becomes stronger along with the decrease of TFA

241

proportion (Figure 4e), thus leading to the higher pKR+ value at 0% TFA.

14 ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

242

We further investigated the feasibility of observing the competitive reactions between

243

different nucleophiles (Figure 5a). To do this, bromide anions were introduced to the

244

solution to act as another nucleophile. Specifically, 10 μmol/L cetyltrimethylammonium

245

bromide was introduced to a HAc/TFA solution (25%/75%, vol/vol). We measured the

246

conductance changes of functioning GMG-SMJs in the HAc/Br−/TFA ternary solution in

247

real time. In addition to RTSs observed previously, a third conductance state was

248

occasionally detected in the It trajectory (Figures 5b5e), which did not exist before. The

249

appearance of this novel conductance state is probably related to a new competitive product:

250

the bromide form (Figure 5a). To understand its origin, in Figure 3 we compared the

251

electronic structures and transport properties of the bromide form with the formers (the

252

carbocation form and the acetate form). The calculated results showed that the bromide

253

and acetate forms have much more similarity in comparison with the carbocation form.

254

This is mainly because both bromide and acetate forms have an identical sp3 hybridization

255

type in the central carbon atom, leading to the close FMOs energy values in Figure 3a and

256

similar conjugated orbital distribution in Figure 3b. The transmission spectrum in Figure

257

3c also shows that the transmission peaks (p-FMOs) of these two sp3-hybridized structures

258

are close, while much further away from EF, in comparison with the sp2-hybridized

259

carbocation form. Furthermore, we noticed that there is a minor transmission difference

260

between these two structures as shown in Figure 3c and the p-HOMO (which dominates

261

the conductance) of the bromide form is a bit closer to EF in comparison with the acetate

262

form. According to the models and the quantum transport analysis discussed above, it is

263

reasonable to conclude that the bromide form and the acetate form should correspond to

15 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 30

264

the experimentally-observed two similar low conductance states, and the conductance of

265

the bromide form is slightly higher than the acetate form.

266 267 268 269 270 271 272 273 274

Figure 5. Competitive reactions between carbocation and different nucleophiles. (a) Schematic representation of competitive reactions of carbocation with Ac− and Br−. (b-e) Representative It trajectories and corresponding enlarged views of a GMG-SMJ, which was submerged into a HAc/Br−/TFA ternary solution (25% HAc/75% TFA (vol/vol); Br−: 10 μmol/L). Ac, Br and C+ represent the acetate form, the bromide form and the carbocation form, respectively.

275

carbocation form (C+, red), the bromide form (Br, orange) and the acetate form (Ac, blue),

276

respectively. On the basis of the attribution, we can elucidate the reaction pathway of

277

individual molecules in GMG-SMJs from real-time conductance recordings, which are

278

faithfully synchronous to the chemical reaction. By taking Figure 5e as an example, we

279

witnessed

280

Ac−(C+)−Br−(C+)−Ac−(C+)−Br−(C+)−Br−(C+)−Ac, which nicely includes the similar

281

reaction

282

Ac−(C+)−Ac−(C+)−Br−(C+)−Ac in Figure 5d. Reproducibly, the fact we found is that the

283

carbocation form is an inevitable intermediate in the transition between the bromide and

Therefore, we attribute these three conductance states (from high to low) to the

the

processes

transformation

of

Ac−(C+)−Br−(C+)−Ac

16 ACS Paragon Plus Environment

pathway

in

Figure

of

5c

and

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

284

acetate forms, which confirms the classical SN1 mechanism. It should be mentioned that

285

the carbocation form was less populated when the bromide anion was added, which is

286

consistent with the bulk experiments (Figure S6). In addition, the kinetics of competitive

287

reactions were analyzed according to the statistical approach used above. The rate constants,

288

involving the heterolytic dissociation process of the acetate/bromide forms (kd) and the

289

combination process of the carbocation form with acetic acid/bromide (kc), were calculated

290

as kd = (1.16 ± 0.18) × 102 s1 and kc = (2.73 ±0.16) ×103 s1 (Figure S18), respectively.

291

In comparison with the condition without the addition of the bromide anion, the rate of

292

heterolytic dissociation barely changed while the rate of combination increased. Such

293

results also confirm the classical SN1 mechanism, that is, the heterolytic dissociation

294

process (first step) is only related to the reactant and the combination process (second step)

295

is related to both the reactant and the nucleophile at the same time.

296

In conclusion, this work demonstrates a single-molecule way to overcome the

297

difficulty of realizing label-free, real-time electrical measurements of fast reaction

298

dynamics with single-event sensitivity and high temporal resolution, and reveal molecular

299

mechanisms of classical chemical reactions. Both experimental and theoretical results

300

demonstrated the reversible acid-catalyzed SN1 reactions at the single-molecule level,

301

including acidity dependence and nucleophilic competitive reactions. On the basis of these

302

results, more details in SN1 reactions, such as racemization and intramolecular

303

rearrangement reactions, are feasible to investigate in the future. By rationally integrating

304

target functional groups into molecular bridges via molecular engineering, this approach

305

offers a promising tool for revealing fundamental mechanisms of general chemical 17 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

306

reactions as well as deeply understanding the basic processes of life at the molecular level

307

and developing accurate molecular diagnostics.

308

ASSOCIATED CONTENT

309

Supporting Information

310

The Supporting Information is available free of charge on the ACS Publications website at

311

DOI: 10.1021/acs.nanoletter.xxxxxxx.

312

Macroscopic experiments, molecular synthesis, device characterization and analysis,

313

characterization of control devices, detailed theoretical analysis, dynamic analysis, and

314

additional experimental results (PDF)

315

AUTHOR INFORMATION

316

Corresponding Author

317

*E-mail: [email protected]

318

*E-mail: [email protected]

319

ORCID

320

Chunhui Gu: 0000-0002-2332-4513

321

Xuefeng Guo: 0000-0001-5723-8528

322

Author contributions

323



324

Notes

325

The authors declare no competing financial interests

326

ACKNOWLEDGMENTS

C.G. and C.H. contributed equally to the work.

18 ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

327

We acknowledge primary financial supports from National Key R&D Program of China

328

(2017YFA0204901), the National Natural Science Foundation of China (grant 21727806),

329

the Natural Science Foundation of Beijing (Z181100004418003) and NSERC of Canada.

330

We thank the CalculQuebec and Compute Canada for computational facility where the

331

calculations were carried out. We are particularly grateful to Prof. Andrew M. Rappe from

332

University of Pennsylvania, Prof. K. N. Houk in University of California, Los Angeles,

333

Prof. Zhenfeng Xi from Peking University and Prof. Yiqin Gao from Peking University for

334

helpful discussions.

335

REFERENCES

336

(1) Coontz, R.; Hanson, B. Science 2004, 305, 957–957.

337

(2) Mikosch, J.; Trippel, S.; Eichhorn, C.; Otto, R.; Lourderaj, U.; Zhang, J.; Hase, W.

338 339 340

L.; Weidemueller, M.; Wester, R. Science 2008, 319, 183–186. (3) Otto, R.; Brox, J.; Trippel, S.; Stei, M.; Best, T.; Wester, R. Nat. Chem. 2012, 4, 534– 538.

341

(4) Gu, C.; Jia, C.; Guo, X. Small Methods 2017, 1, 1700071.

342

(5) Sarkar, S. K.; Bumb, A.; Mill, M.; Neuman, K. C. Cell 2013, 153, 1408.

343

(6) Roy, R.; Hohng, S.; Ha, T.; Nat. Methods 2008, 5, 507–516.

344

(7) Lu, H. P.; Xun, L.; Xie, X. S. Science 1998, 282, 1877–1882.

345

(8) Kasianowicz, J. J.; Brandin, E.; Branton, D.; Deamer, D. W. Proc. Natl. Acad. Sci.

346 347 348 349

USA 1996, 93, 13770–13773. (9) Song, L.; Hobaugh, M. R.; Shustak, C.; Cheley, S.; Bayley, H.; Gouaux, J. E. Science 1997, 274, 1859–1866. (10) Bayley, H.; Cremer, P. S. Nature 2001, 413, 226–230. 19 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

350 351 352 353

(11) Bouilly, D.; Hon, J.; Daly, N. S.; Trocchia, S.; Vernick, S.; Yu, J.; Warren, S.; Wu, Y.; Gonzalez, R. L., Jr.; Shepard, K. L.; Nuckolls, C. Nano Lett. 2016, 16, 4679–4685. (12) Goldsmith, B. R.; Coroneus, J. G.; Kane, A. A.; Weiss, G. A.; Collins, P. G. Nano Lett. 2008, 8, 189–194.

354

(13) Xiang, D.; Wang, X.; Jia, C.; Lee, T.; Guo, X. Chem. Rev. 2016, 116, 4318–4440.

355

(14) Reed, M. A.; Zhou, C.; Muller, C. J.; Burgin, T. P.; Tour, J. M. Science 1997, 278,

356

252–254.

357

(15) Xu, B.; Tao, N. Science 2003, 301, 1221–1223.

358

(16) Venkataraman, L.; Klare, J. E.; Nuckolls, C.; Hybertsen, M. S.; Steigerwald, M. L.

359 360 361

Nature 2006, 442, 904–907. (17) Cui, X. D.; Primak, A.; Zarate, X.; Tomfohr, J.; Sankey, O. F.; Moore, A. L.; Moore, T. A.; Gust, D.; Harris, G.; Lindsay, S. M. Science 2001, 294, 571–574.

362

(18) Jia, C.; Ma, B.; Xin, N.; Guo, X. Acc. Chem. Res. 2015, 48, 2565–2575.

363

(19) Diez-Perez, I.; Hihath, J.; Lee, Y.; Yu, L. P.; Adamska, L.; Kozhushner, M. A.;

364 365 366 367 368 369 370

Oleynik, II; Tao, N. J. Nat. Chem. 2009, 1, 635–641. (20) Sherif, S.; Rubio-Bollinger, G.; Pinilla-Cienfuegos, E.; Coronado, E.; Cuevas, J. C.; Agrait, N. Nanotechnology 2015, 26, 291001. (21) Capozzi, B.; Xia, J.; Adak, O.; Dell, E. J.; Liu, Z.; Taylor, J. C.; Neaton, J. B.; Campos, L. M.; Venkataraman, L. Nat. Nanotech. 2015, 10, 522–527. (22) Song, H.; Kim, Y.; Jang, Y. H.; Jeong, H.; Reed, M. A.; Lee, T. Nature 2009, 462, 1039–1043.

20 ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

371

(23) Jia, C.; Migliore, A.; Xin, N.; Huang, S.; Wang, J.; Yang, Q.; Wang, S.; Chen, H.;

372

Wang, D.; Feng, B.; Liu, Z.; Zhang, G.; Qu, D.; Tian, H.; Ratner, M. A.; Xu, H.; Nitzan,

373

A.; Guo, X. Science 2016, 352, 1443–1445.

374

(24) Green, J. E.; Choi, J. W.; Boukai, A.; Bunimovich, Y.; Johnston-Halperin, E.;

375

DeIonno, E.; Luo, Y.; Sheriff, B. A.; Xu, K.; Shin, Y. S.; Tseng, H. R.; Stoddart, J. F.;

376

Heath, J. R. Nature 2007, 445, 414–417.

377

(25) Xin, N.; Guo, X. Chem 2017, 3, 373–376.

378

(26) Wen, H. M.; Li, W. G.; Chen, J. W.; He, G.; Li, L. H.; Olson, M. A.; Sue, A. C.-H.;

379 380 381 382 383

Stoddart, J. F.; Guo, X. F. Sci. Adv. 2016, 2, e1601113. (27) Guan, J.; Jia, C.; Li, Y.; Liu, Z.; Wang, J.; Yang, Z.; Gu, C.; Su, D.; Houk, K. N.; Zhang, D.; Guo, X. Sci. Adv. 2018, 4, eaar2177. (28) Olah, G. A.; Baker, E. B.; Evans, J. C.; Tolgyesi, W. S.; McIntyre, J. S.; Bastien, I. J. J. Am. Chem. Soc. 1964, 86, 1360–1373.

384

(29) Olah, G. A.; Lukas, J. J. Am. Chem. Soc. 1967, 89, 2227–2228.

385

(30) Wilson, H.; Ripp, S.; Prisbrey, L.; Brown, M. A.; Sharf, T.; Myles, D. J. T.; Blank,

386 387 388

K. G.; Minot, E. D. J. Phys. Chem. C 2016, 120, 1971–1976. (31) Sarwat, S. G.; Gehring, P.; Rodriguez Hernandez, G.; Warner, J. H.; Briggs, G. A. D.; Mol, J. A.; Bhaskaran, H. Nano Lett. 2017, 17, 3688–3693.

389

(32) Araidai, M.; Tsukada, M. Phys. Rev. B 2010, 81, 235114.

390

(33) Lo, W. Y.; Bi, W.; Li, L.; Jung, I. H.; Yu, L. Nano Lett. 2015, 15, 958–962.

391

(34) Bâldea, I.; Xie, Z.; Frisbie, C. D. Nanoscale 2015, 7, 10465–10471.

392

(35) Huang, C.; Jevric, M.; Borges, A.; Olsen, S. T.; Hamill, J. M.; Zheng, J. T.; Yang,

393

Y.; Rudnev, A.; Baghernejad, M.; Broekmann, P.; Petersen, A. U.; Wandlowski, T.; 21 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

394

Mikkelsen, K. V.; Solomon, G. C.; Brondsted Nielsen, M.; Hong, W. Nat. Commun. 2017,

395

8, 15436.

396

(36) Xin, N.; Wang, J.; Jia, C.; Liu, Z.; Zhang, X.; Yu, C.; Li, M.; Wang, S.; Gong, Y.;

397

Sun, H.; Zhang, G.; Liu, Z.; Zhang, G.; Liao, J.; Zhang, D.; Guo, X. Nano Lett. 2017, 17,

398

856–861.

399 400

(37) Atesci, H.; Kaliginedi, V.; Celis Gil, J. A.; Ozawa, H.; Thijssen, J. M.; Broekmann, P.; Haga, M.-a.; van der Molen, S. J. Nat. Nanotech. 2018, 13, 117–121.

401

(38) Shaik, S.; Mandal, D.; Ramanan, R. Nat. Chem. 2016, 8, 1091–1098.

402

(39) Aragones, A. C.; Haworth, N. L.; Darwish, N.; Ciampi, S.; Bloomfield, N. J.;

403

Wallace, G. G.; Diez-Perez, I.; Coote, M. L. Nature 2016, 531, 88–91.

404

22 ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

405

Figure Legends

406

Figure 1. Schematic of a graphene-molecule-graphene single-molecule junction that

407

shows the formation dynamics of a 9-phenyl-9-fluorenyl cation in a SN1 reaction.

408

Figure 2. Real-time conductance measurements of a representative GMG-SMJ among 5

409

different devices. (a) Representative It trajectories in different conditions (from top to

410

bottom: in the air, pure HAc, 25% TFA/75% HAc, 50% TFA/50% HAc, and 75% TFA/25%

411

HAc (vol/vol)). (b) Partial It curves of the corresponding parts in (a). Insets show the

412

enlarged views of the peaks as indicated by white arrows. (c)

413

histograms of (a). The blue and red lines show the gauss-shaped fit for the low and high

414

conductance states, respectively. Insets show the enlarged views in the high conductance

415

regions. Source-drain bias voltage (VD): 0.3 V; Gate voltage (VG): 0 V; Sampling rate: 57.6

416

kSa/s.

417

Figure 3. Theoretical simulations of GMG-SMJs. (a) Calculated energy levels of the

418

molecular wire in the three forms: R = (+) (Carbocation, red), R = Ac (acetate, blue) and

419

R = Br (Bromide, orange). The dash line shows the Fermi level of graphene electrodes: EF

420

= –4.8 eV. (b) Corresponding calculated frontier molecular orbitals (FMOs) of the

421

molecular wire in the three forms. (c) Transmission spectra of GMG-SMJs in the three

422

forms around the Fermi level of graphene at a zero-bias voltage.

423 424 425 23 ACS Paragon Plus Environment

The corresponding

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

426

Figure 4. Statistical analyses of the GMG-SMJ device used in Figure 2. (a) Measured It

427

trajectories (black dots) of a GMG-SMJ in the solution of HAc-TFA (25%/75%, vol/vol).

428

The red line shows the idealised two-level interconversion by using a QUB software. (b,c)

429

Plots of time intervals of the low (b) and high (c) states derived from It trajectories of a

430

GMG-SMJ in the solution of HAc-TFA (25%/75%, vol/vol). The distributions were well

431

fitted by a single-exponential decay function (blue lines). Insets show the corresponding

432

reaction pathways. (d) Activation energies of the symmetric SN1 reaction derived from

433

macroscopic (dash lines) and single-molecule (solid lines) experiments. (e) Combination

434

constants of carbocation (pKR+). Column: pKR+ values derived from GMG-SMJs. Solid

435

columns represent the respective pK values; Striped columns represent the generalized acid

436

function J0 in the corresponding solutions. The blue and black dash lines show the J0TFA%

437

relationship and the pKR+ values derived from bulk experiments, respectively.

438

Figure 5. Competitive reactions between carbocation and different nucleophiles. (a)

439

Schematic representation of competitive reactions of carbocation with Ac− and Br−. (b-e)

440

Representative It trajectories and corresponding enlarged views of a GMG-SMJ, which

441

was submerged into a HAc/Br−/TFA ternary solution (25% HAc/75% TFA (vol/vol); Br−:

442

10 μmol/L). Ac, Br and C+ represent the acetate form, the bromide form and the

443

carbocation form, respectively.

444 445 446

24 ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

447

Nano Letters

Figure 1

448 449 450 451 452 453 454 455 456 457

25 ACS Paragon Plus Environment

Nano Letters

458

Figure 2 a

b

Current (nA)

Air

Current (nA)

0

0.2

0.4

0.6

0.8

1.0

Current (nA)

Current (nA)

0

0

80

80

50

50

50

20 0.967

20

40

0.2

0.4

0.6

1.0

0.8

0.5 ms

0.9675

80

85

0.968

55 0

0 120

90

90

90

60

60

30

30 0.2

0.4

0.6

0.8

1.0

0.20

40 0

0.215

0

0.23

400

400 300

200

200

200

200

100

100 0 0.42

1.0

0.5 ms

0.425

0.44

0

0

400

400

400

300

300

200

200

200

100

100

100

0

0.2

0.4 0.6 Time (s)

1.0

0.8

0 0.05

460 461 462 463 464 465 466 467 468 469 26 ACS Paragon Plus Environment

30

7000

0

15

30

100

300

0

15

3500

300

0.8

2400

30

400

0.6

30

80

60

300

0.4

15

120

400

0.2

30

3600

1200

300

0

15

1800 115

120 0.5 ms

20 0

110

120

0

459

0.60

0.575

110

0

75% TFA 25% HAc

0 0.55

110

0

50% TFA 50% HAc

20

20

20

20

25% TFA 75% HAc

40

30

0

0% TFA 100% HAc

c 40

40

Current (nA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 30

0.075 Time (s)

0.10

0

0

6000

2500 Counts

12000

5000

Page 27 of 30

470

Figure 3

a

b

0 –1

–1.71

–1.66

–5.50

–5.42

Ac

Br

–2

E (eV)

–3

LUMO

–3.64

–4 –5 –6

HOMO –4.98

R = (+)

–7 R = (+)

c Transmission

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

p-HOMO

R = Ac p-LUMO p-LUMO

p-HOMO

10 0

R = Br

p-HOMO-1 p-HOMO

10 –4

p-LUMO

10 –8 R= (+) R= Ac R= Br

10 –12

10 –16 –2.0

–1.5

–1.0

–0.5

471

0 0.5 E–EF (eV)

472 473 474 475 476 477 478 479 480 481 482 483 27 ACS Paragon Plus Environment

1.0

1.5

2.0

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

484

Figure 4

485 486

28 ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30

Figure 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

487

29 ACS Paragon Plus Environment

Nano Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

488 489

Keywords: Single-molecule detection, reaction dynamics, molecular electronics, nucleophilic substitution

490 491 492 493 494 495 496 497

TOC Graphic: Through molecular engineering, a single molecular wire containing a functional center of 9-phenyl-9-fluorenol was covalently wired into nanogapped graphene electrodes to form stable single-molecule junctions. Both experimental and theoretical studies consistently demonstrate and interpret direct measurements of the dynamic processes of detailed molecular transformations of a SN1 reaction at the singlemolecule/single-event level.

498

30 ACS Paragon Plus Environment

Page 30 of 30