Azonia aromatic heterocycles as a new acceptor unit in D-π-A+ vs D-A+ nonlinear optical chromophores

Azonia aromatic heterocycles as a new acceptor unit in D-π-A+ vs D-A+ nonlinear optical chromophores

Accepted Manuscript + + Azonia aromatic heterocycles as a new acceptor unit in D-π-A vs D-A nonlinear optical chromophores Tatiana Cañeque, Ana M. Cua...

1MB Sizes 26 Downloads 85 Views

Accepted Manuscript + + Azonia aromatic heterocycles as a new acceptor unit in D-π-A vs D-A nonlinear optical chromophores Tatiana Cañeque, Ana M. Cuadro, Raúl Custodio, Julio Alvarez-Builla, Belén Batanero, Pilar Gómez-Sal, Javier Pérez-Moreno, Koen Clays, Obis Castaño, José L. Andrés, Thais Carmona, Francisco Mendicuti, Juan J. Vaquero PII:

S0143-7208(17)30422-9

DOI:

10.1016/j.dyepig.2017.05.005

Reference:

DYPI 5967

To appear in:

Dyes and Pigments

Received Date: 28 February 2017 Revised Date:

1 May 2017

Accepted Date: 2 May 2017

Please cite this article as: Cañeque T, Cuadro AM, Custodio Raú, Alvarez-Builla J, Batanero Belé, Gómez-Sal P, Pérez-Moreno J, Clays K, Castaño O, Andrés JoséL, Carmona T, Mendicuti F, + + Vaquero JJ, Azonia aromatic heterocycles as a new acceptor unit in D-π-A vs D-A nonlinear optical chromophores, Dyes and Pigments (2017), doi: 10.1016/j.dyepig.2017.05.005. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Azonia aromatic heterocycles as a new acceptor unit in D-π-A+ vs D-A+ nonlinear optical chromophores Tatiana Cañeque,1 Ana M. Cuadro,1,* Raúl Custodio,1 Julio Alvarez-Builla,1 Belén Batanero,1 Pilar

Francisco Mendicuti4 and Juan J. Vaquero1,* 1

RI PT

Gómez-Sal,1 Javier Pérez-Moreno,2 Koen Clays,3 Obis Castaño,4 José L. Andrés,4 Thais Carmona,4

Departamento de Química Orgánica y Química Inorgánica, Universidad de Alcalá, 28871-Alcalá de

Henares, Madrid, Spain. 2Department of Physics. Skidmore College, Saratoga Springs, NY. 3Department

SC

of Chemistry, University of Leuven, Celestijnenlaan 200 D,3001 Leuven, Belgium. 4Departamento de Química Analítica, Química Física e Ingeniería Química, Universidad de Alcalá, 28871 Alcalá de

M AN U

Henares, Madrid, Spain

[email protected], [email protected]

ABSTRACT: A comparison of D-π-A+ and D-A+ cationic chromophores based on the quinolizinium system as new acceptor units is reported along with the results of studies into their linear and non-linear

TE D

optical properties and electrochemical data. Experimental and theoretical data show that quinoliziniumbased chromophores may provide a new generation of second-order non-linear materials with enhanced performance. The first hyperpolarizabilities were measured by Hyper-Rayleigh scattering experiments and the experimental data are supported by a theoretical analysis. In some chromophores the absence of a

EP

bridge (D-A+) between the donor and acceptor fragments enhances the NLO properties and the single

AC C

crystal structure of such a material has been determined by X-ray diffraction

KEYWORDS: NLO, quinolizinium acceptor unit, charged chromophores, azonia aromatic heterocycles, synthesis D-A, theoretical calculations

1

ACCEPTED MANUSCRIPT

M AN U

SC

RI PT

GRAPHICAL ABSTRACT:

Research highlight

Design and synthesis of new D-π-A+ vs D-A+ chromophores based on quinolizinium cation Unbridged D-A+ chromophores with remarkable large beta values

TE D

Cyclic voltammetry measurements compared with DFT calculations

EP

The work provides valuable hints for rational design of novel D-A+ materials

AC C

Appendix A. Supplementary information (SI)

2

ACCEPTED MANUSCRIPT 1. Introduction There is considerable interest in organic non-linear optical (NLO) materials[1] with large second-order optical non-linearities in various fields (chemistry, physics, materials) because these materials have found diverse applications in optoelectronics[2], all-optical

RI PT

data processing technology[3], bioimaging[4], sensitizers for solar cells[5], and photodynamic therapy[6] amongst others. Generally, most effort has been focused on the design of new chromophores with large first hyperpolarizability values (β)[7], which are related to an electronic intramolecular charge transfer (ICT) due to the ease with which

SC

the molecular structures, and hence the properties, can be tuned for various applications. Second-order NLO organic materials are most often based on π-conjugated molecules

M AN U

(chromophores) with strong electron donor and acceptor groups at the ends of the π-conjugated structure[8]. Although π-electron delocalization is the most widely studied and exploited form of ICT, electrons are also known to delocalize over σ bonds[9] and this topic has barely been studied. Conjugated compounds have been the focus of most research within this field, but

TE D

today it is recognized and accepted that ICT occurs through a synergistic combination of different phenomena[10]. Although a wide range of materials has been studied over the past few decades, the most recent approaches to the design of highly NLO-active systems is based on

Me2 N

Me2 N

AC C

Me2N

EP

charged molecules with the presence of cationic acceptors.

Me2 N

NMe2

N Me + PF6 S +N

+ N

PF6

NMe2

PF Me

N +

N + PF

PF

N + PF

+ PF

N

N+ PF

Figure 1. Representative examples of nonlinear optical chromophores based on azolium and pyridiniun cations as acceptor units.

Cationic chromophores are of interest because they possess an easy tunability by counter ion exchange to crystallize into non-centrosymmetric structures[11]. Furthermore, by tuning their 3

ACCEPTED MANUSCRIPT solubility, potential applications in imaging can also be achieved. Other potential advantages of charged chromophores include their greater stability and higher chromophore number densities when compared with other NLO materials. In this context, charged acceptors units are referred to benzothiazolium[12] and azinium[13] (pyridinium, quinolinium) cations. These latter having

RI PT

interest in the generation of terahertz (THz) radiation[14] as 1D and 2D chromophores represented by DAST(4-(??,N-dimethylamino)-4´-N′-methyl-stilbazoliumtosylate)/DSTMS(4(??,N-dimethyl

amino-4′-??′-methyl-stilbazolium-2,4,6-trimethylbenzenesulfonate)[15]

and

diquats/helquats [16] (Figure 1). However, azonia aromatic heterocycles (AZAH)[17] have not

SC

been explored to date as an acceptor unit, except in our work, as a marker in nonlinear optical bioimaging, exhibiting a large two-photon absorption (2PA) and as push-pull cationic

M AN U

chromophores (D-π-A+-π-D)[18] (Figure 2). These kinds of heteroaromatic cations containing a quaternary bridgehead nitrogen are well represented in nature, forming part of a wide range of alkaloids[19]. Furthermore, some derivatives of these azonia salts have proven useful in important applications such as cyanine dyes and highly fluorescent compounds used as probes

EP

TE D

for the detection of biomolecules[20].

R = OMe, NMe 2

AC C

Figure 2. Quinolizinium as acceptor unit in 1D (D-π-A+) and 2D (D-π-A+-π-D) chromophores.

As a continuation of our research on the applications of heteroaromatic cations in the NLO field[21], we explored the potential of new chromophores based on the quinolizinium cation, which may exhibit better non-linear optical properties (NLO) than other heteroaromatic cations tested to date, i.e., azinium[22]and azolium[23] salts. In addition, comparative studies on the NLO properties were carried out in order to investigate the role of quinolizinium as an acceptor and to establish the relevant structure activity relationships in simple dipolar molecules. This

4

ACCEPTED MANUSCRIPT was achieved by changing the nature of the π-bridge that connects the donor units and comparing these chromophores with unbridged ones (Figure 3). EDG

EDG

+

D

D

D-A+ EDG: electron-donating group

+

X

D-π-A+

+

+

D D-π-A+

N +

RI PT

D

SC

Figure 3. General structure of push-pull systems based on the quinolizinium cation.

M AN U

We report here our results on the use of the quinolizinium cation as an acceptor unit in push-pull chromophores in which the quinolizinium moiety is connected to an electron donor (D) through a π-conjugated bridge (D-π-A+) or directly (D-A+) by Pd-catalysed cross-coupling reactions (Figure 4). The aim was to gain a better understanding of these systems and to contribute to the recent interest in β-enhancement strategies[21a,24].

TE D

The second-order NLO properties of the selected chromophores were evaluated and compared to experimental results and theoretical calculations previously carried out on azinium cations. The resulting D-A+ and D-π-A+ chromophores with the quinolizinium acceptor were studied by

EP

hyper-Rayleigh scattering experiments and by theoretical analysis, including Density Functional Theory (DFT) and correlated Hartree–Fock-based methods (RCIS(D)). In addition, the redox

AC C

activity was evaluated by cyclic voltammetry and the linear optical properties related to the ICT characteristics of the D-A+ and D-π-A+ are included along with the X-ray structures of selected D-A+ chromophores.

2. Results and discussion 2.1 Synthesis The quinolizinium cation is the simplest model that offers four possible positions of substitution by palladium-mediated cross-coupling processes and these products exhibit markedly different

5

ACCEPTED MANUSCRIPT behaviour from an electronic point of view at the C1/C3 and C2/C4 positions. Moreover, one would also expect a different steric effect on C1/C4 and C2/C3. Our strategy for the construction of push–pull molecules was to combine quinolizinium cations – employed as acceptor units (A+) – with donor units (D) through a π-bridge using the Sonogashira[25] and

reactions

using

either

the

appropriate

arylboronic

-

-

PF6

or

potassium

M AN U

R = OMe, NMe2

acids

SC

organotrifluoroborates[27,28](Figure4) .

RI PT

Heck reactions or to connect the units directly by Stille or Suzuki−Miyaura[26] coupling

Figure 4. Synthesis of cationic (D-π-A+) and D-A+) chromophores by Pd catalysed cross-coupling reactions.

The first series of quinolizinium-based chromophores was previously obtained by us from the four isomeric bromoquinolizinium cations[29] (1a–d) (Scheme 1), which were

TE D

coupled with different arylalkynes under Sonogashira conditions[25]. This method allowed easy access for the first time to representative D-π-A+ quinolizinium-based chromophores and allowed the efficient incorporation of arylethynyl substituents into the quinolizinium system,

EP

particularly at the C2 and C3 positions (Scheme 1).

AC C

10% CuI 5% Pd2Cl2(PPh3)2

Et 3N, DMF, 60 ºC or rt

1 1a = 1-Br 1b = 2-Br 1c = 3-Br 1d = 4-Br

2a (1-) 47% 2b (2-) 78% 2c (3-) 71% 2d (4-) 61%

3b (2-) 55%

2-5 X = Br, PF6 R = OMe, NMe2, Me, Br

-

-

4b (2-) 74% 4c (3-) 76%

5b (2-) 71% 5c (3-) 70%

Scheme 1. Synthesis of D-π-A+ chromophores 2-5 by Sonogashira reaction 6

ACCEPTED MANUSCRIPT In order to study and compare the NLO properties of the quinolizinium cation as an acceptor in D-π-A+ to D-A+, we discuss here previously published results on quinolizinium[18b] linked to C1-C4 with dimethylamino- and methoxystyryl derivatives. These compounds were synthesized by Heck reactions using the corresponding vinyl quinolizinium hexafluorophosphate and aryl

RI PT

iodides as reaction partners or, alternatively, by a more classical approach involving a Knoevenagel reaction between 2-methylquinolizinium hexafluorophosphate[30] and the corresponding aryl aldehyde in higher yields as indicated in Scheme 2. I N +_

SC

R N _+

6,7

R

PF6

PF6

N _+

M AN U

R

N _+

N _+

PF6

N _+

PF6

6b: R = NMe2(42%) 7b: R = OMe (81%)

6a: R = NMe2 (37%)

R

PF6

37-81%

PF6

6c: R = NMe2(42%) 7c: R = OMe (37%)

R R 6d: R = NMe2(42%)

CHO

TE D

Me

N _+

PF6

R

MeCN Piperidin

6b: R = NMe2(92%) 7b: R = OMe (81%)

EP

Scheme 2. Synthesis of chromophores 6 and 7 by Heck and Knoevenagel reaction

AC C

Although initial studies in our laboratory showed that the Stille[26] coupling reaction was the most efficient procedure in comparison to other palladium-catalysed reactions to produce high yields of some aryl- and heteroaryl-quinolizinium cations, we prepared the new quinolizinium derivatives 9 and 10 by the Suzuki–Miyaura reaction[31], which is a greener procedure according to conditions employed by us to synthesize heteroaryl quinolizinium derivatives[27] (Scheme 3). Optimization of the reaction conditions was carried out first with the 2-bromoquinolizium salt, with different bases, catalytic systems and temperatures evaluated. It was found that the best

7

ACCEPTED MANUSCRIPT catalytic system was Pd2(dba)3 as this allowed the reaction to be carried out at room temperature in conjunction with bulky phosphines: P(o-Tol)3 or (2-biphenyl)-di-tert-butylphosphine.

DMF, r.t. Method A or B

52-80% , -

1 H2O, 65º C Method C Method C

8%mol Pd(o-tol)3

8b: R = 3-OMe; 71% (C) 9b: R = 4-OMe; 52% (A); 73% (B); 76% (C) 10b: R = 4-NMe2: 73% (b); 80% (C)

1%mol, Pd(OAc)2;

8c: R = 3-OMe; 77% (C) 9c: R = 4-OMe; 60% (B); 80% (C) 10c: R = 4-NMe2: 60% (B)

M AN U

8a: R= 3-MeO; 93% (C)

Method B 4%mol, Pd2(dba)3;

SC

Method A 2%mol, Pd2(dba)3; 4%mol Pd(o-tol)3

RI PT

9-10

1b= 2-Br 1c= 3-Br

8d: R= 3-MeO; 63% (C)

Scheme 3. Suzuki–Miyaura coupling reaction with boronic acids and potassium organotrifluoroborates

Thus, the coupling between 1b and phenylboronic acid with electron-donating groups (OMe and

TE D

NMe2) afforded the product 9b in moderate yield (52%) when Pd2(dba)3/P(o-Tol)3 in a ratio of 2%/4% (Method A) was used or in good yield for 9 and 10 with a 4%/8% ratio (Method B) in DMF in the presence of K2CO3 at room temperature for 16 hours (Scheme 3). The 3-

EP

bromoquinolizinium bromide (1c) was also subjected to Suzuki coupling in order to provide an overview of the reactivity of the C2 and C3 positions, which are electronically very different

AC C

and similar to C4 and C1, respectively. Alternatively, coupling with potassium organotrifluoroborates, (Method C) an efficient reagent that allows functionalization of quinolizinium under mild conditions by reaction with 1b, afforded chromophores 9b–10b in higher yield (Scheme 3) and similar to 8a–d, which were previously synthesized by us[27].

8

ACCEPTED MANUSCRIPT 2.2 Linear optical properties The UV-Vis spectra which are shown in the wavelength range where the main absorption bands appeared (Figures 5-8), were analyzed to evaluate the effectiveness of different electrondonating substituents (D) with different π-linkers in D-π-A+ vs D-A+ chromophores.

RI PT

The absorption spectra of a group of representative chromophores (2b–5b) bearing different electron-donating substituents (D) in position C2 of the quinolizinium (A+) with an acetylenic πlinker are represented in Figure 5. As expected, a good correlation was found between the electron-donating capabilities of the substituents and the relative position of the maximum of

SC

the ICT band (λabs,max). The electron-donating capability increased in the order Br- < Me- < MeO- < Me2N- and a corresponding bathochromic displacement of λabs,max was observed. In

M AN U

general, the λabs,max for compounds containing dimethylamino-substituents (Me2N–) shifted towards longer wavelengths (430 nm for 3b) with respect to those containing Br-substituents (341 nm for 5b).

TE D

2b λmax=378 nm 3b λmax=430 nm

0.8

4b λmax=367 nm 5b λmax=341 nm

0.6 0.4

EP

Absorbance (a.u.)

1.0

AC C

0.2

0.0 300

350

400

450

500

550

λ (nm)

Figure 5. Absorption spectra (normalized at the maximum of intensity) in the 300–580 nm range for 2– 5b in MeOH as a solvent at 25 °C.

The normalized absorption spectra of chromophores 6a–d with styryl moieties, which are collected in Figure 6, also showed the intense ICT absorption band attributed to the Me2NC6H4→QZ charge transfer. This band was significantly shifted to the red in compounds 6b (473 nm) and 6d (445 nm) when compared to chromophores 6a (425 nm) and 6c (433 nm). For the

9

ACCEPTED MANUSCRIPT former derivatives the Me2NC6H4– group was linked through a styryl moiety to the quinolizinium acceptor at the C2 and C4 positions, for the latter ones, however, the donors were located in the less activated C1 and C3 positions. This fact could be explained by a more efficient conjugation in positions 2 and 4 compared to those in 1 and 3, due to a combination of

RI PT

electronic and steric effects. A similar situation was observed for 2a–d and 8a–d, where the greatest stabilization took place for compounds b and/or d (see SI Figures S1 and S2).

The absorption spectra of chromophores 6b,c and 7b,c are also depicted in Figure 7. The ICT absorption bands were shifted to significantly longer wavelengths for compounds 6b,c

SC

substituted at C2- and C3-, when compared to 7b,c, by about 91 and 81 nm respectively. For unbridged chromophore D-A+ pairs 9b, 10b and 9c, 10c the displacements were approximately

M AN U

83 nm and 60 nm, respectively (see Figure S3 in the SI). For compounds 2b-3b, which contain acetylenic π-linkers, the maxima of the bands were displaced by around 52 nm. The stronger electron-donating character of Me2N– favors the stabilization of compounds that contain this substituent relative to those bearing MeO– regardless of the linker. The stabilization of

members of the series.

TE D

compounds b with respect to c due to the position of the substituent was also evident for all

1.0

6a λmax=425 nm

AC C

Absorbance (a.u.)

EP

6b λmax=473 nm 6c λmax=433 nm

0.8

6d λmax=445 nm

0.6 0.4 0.2

0.0 300

350

400

450

500

550

600

650

λ (nm) Figure 6. Absorption spectra (normalized at the maximum of intensity) in the 300–680 nm range for chromophores 6a–d in MeOH at 25 °C.

10

ACCEPTED MANUSCRIPT

1.0

7b λmax=382 nm

Absorbance (a.u.)

7c λmax=352 nm 0.8

6b λmax=473 nm 6c λmax=433 nm

0.6

0.2 0.0

300

350

400

450

500

550

600

RI PT

0.4

650

SC

λ (nm) Figure 7. Absorption spectra (normalized at the maximum of intensity) in the 300–650 nm range of pushpull 7b-c and 6b-c chromophores in MeOH at 25ºC.

M AN U

A good correlation was also found between λabs,max for D-A+ unbridged chromophores 8b,c and 9b,c (see Figure S4 of the SI). The λabs,max for the methoxyphenyl groups in para-positions (9b,c) showed a bathochromic displacement of the ICT band relative to 8b,c by about 13 and 5 nm, respectively, because meta-substitution leads to less effective conjugation.

TE D

Certainly, the para-substituted electron-rich MeOC6H4– and Me2NC6H4– groups at the C2 quinolizinium position (compounds 2b, 7b, 9b and 3b, 6b, 10b respectively) strongly favor DA+ conjugation stabilizing the ICT complex when compared to other series of studied

EP

compounds. For this reason, a more extensive photophysical study was performed on these compounds in order to gain a deeper insight into the linear properties and their relationship with

AC C

the non-linear ones.

The normalized absorption and emission spectra for the quinolizinium derivatives 10b, 6b and 3b, which have Me2NC6H4– as donors, are shown in Figures 8(a) and 8(b), respectively. The absorption spectrum of the olefinic chromophore 6b showed a more marked displacement of the ICT band to the red (472 nm) when compared to the acetylenic derivative 3b (447 nm) and the unbridged Me2NC6H4– chromophore 10b (431 nm). This particular trend was previously reported by Moylan et al.,[32] for a series of D-A chromophores connected by simple acetylenic, olefinic and aza linkers. Absorption spectra for the MeOC6H4- substituted 2b, 7b and

11

ACCEPTED MANUSCRIPT 9b derivatives showed a similar trend (see Figure S5 SI) although bands are less shifted to the red than their Me2NC6H4– substituted counterparts. The fluorescence spectra showed analogous features for the band displacements in both series of chromophores. The most red-shifted emission bands correspond to Me2NC6H4– donors

RI PT

containing compounds when compared to their counterparts, which have MeOC6H4- donors. However, the extent of such bathochromic changes depend on donor-linker structure. The absorption and emission maxima, fluorescence quantum yields and molar absorptivities for para-substituted MeOC6H4- and Me2NC6H4– quinolizinium derivatives are listed in Table 1.

SC

The absorption maxima λabs,max and λem,max were similar to those previously reported by us [18a,c]. The emission quantum yield and molar absorptivity values for some chromophores,

M AN U

however, showed some discrepancies but they were of the same order of magnitude. Thus, Stokes shifts of approximately 160, 123 and 171 nm for 10b, 3b and 6b and of 76, 147 and 123 nm for 9b, 2b and 7b, respectively, were found. Different behaviors may arise from the formation of TICT (Twisted intramolecular charge transfer) states in dimethylamino derivatives

TE D

whose stability in polar media depends strongly on the type of donor to the quinolizinium acceptor linker and their internal rotation number. For this reason, as previously reported, the interpretation of some of the photophysical behaviors in these systems is complex[33]. The

EP

fluorescence quantum yields were lower for the Me2NC6H4-containing derivatives than for the MeOC6H4- ones. It is well known that the fluorescence quantum yields usually decrease with

AC C

the electron substituent donating ability. The additional non-radiative deactivation excited state path arising from the dimethylamino twisting in the TICT may also contribute to this. Although with a less efficient electron donating ability, a surprisingly high fluorescence quantum yield was obtained for the unbridged MeOC6H4-containing chromophore 9b which, as expected, was decreased significantly for compounds 7b and 2b, which have olefinic and acetylenic π-linkers.

12

ACCEPTED MANUSCRIPT Table 1. Photophysical properties of the quinolizinium derivatives (see Figure 9) in methanol at 25 °C λexc (nm)

λem,max (nm)

φ

εmax (× 10-3 cm-1Mol-1L)

2b

378

360

525

0.01 ± 0.00a

31.8 ± 0.0

7b

386

360

509

0.02± 0.00a

14.9 ± 1.2

9b

360

360

436

0.35 ± 0.06a

42.1 ± 0.2

3b

447

430

570

< 0.001b

6b

472

460

643

0.006± 0000b

10b

431

420

591

0.08 ± 0.01b

RI PT

λabs,max (nm)

Comp.

37.0 ± 0.3 16.6 ± 0.2 36.4 ± 0.5

SC

Maximum absorption wavelength (λabs,max), excitation wavelength (λexc), emission maxima wavelength (λem,max), emission quantum yield (ϕ) and molar absorptivity at λabs,max (εmax). (a) Using Quinine Sulphate

TE D

0.4

b)

0.8

EP

Absorbance (a.u.)

0.8

0.6

1.0

10b 6b 3b

a)

Fluorescence (a.u.)

1.0

M AN U

(φ=0.546) in H2SO4 0.1N or (b) coumarine 153 in ethanol (φ=0.38) as standards[34].

0.6

0.4

0.2

0.0

0.0

AC C

0.2

300 350 400 450 500 550 600

λ (nm)

500 550 600 650 700 750 800

λ (nm)

Figure 8. Absorption (a) and emission spectra (b), both normalized to 1 at the maximum of intensity, for 10b (black), 6b (red) and 3b (blue) in methanol at 25 °C.

13

OMe ACCEPTED MANUSCRIPT

OMe

OMe

N +_ PF6

N _ + PF6

N PF6 9b

7b

2b

N(Me)2 NMe2

N(Me)2

PF6

PF6 10b

6b

3b

RI PT

N

N +_ PF6

N _ +

SC

Figure 9. Selected chromophores for the study of linear and non-linear optical properties (Tables 1 and 2).

2.3 Electrochemical properties

M AN U

The electrochemical properties of selected quinolizinium salts (Figure 9) were studied by cyclic voltammetry (CV) in acetonitrile/LiClO4 as SSE (solvent-supporting electrolyte system). The cathodic and the anodic peak potentials, measured at a scan rate of 100mV/s and summarized in Table 2, are quoted with respect to the Ag/Ag+ (sat) reference electrode. These cationic chromophores are electroreducible to the corresponding stabilized radicals in the

TE D

same potential range (Epc ∼ –1.2 V) in all cases due to the positive charge on the nitrogen atom of the quinolizinium ring. The reduction of the electrogenerated radical to the corresponding anion requires a more negative potential value and this was only observed in some cases.

EP

A relationship between the energy of the highest occupied molecular orbital (HOMO) and the

AC C

oxidation potential of a molecular organic semiconductor was described by Forrest[35]. The data

for

the

estimation

of

such

energy (EHOMO,

eV)

applied

to quinolizinium

hexafluorophosphates represented in Figure 9 are summarized in Table 2. These experimental results are consistent with theoretical computational calculations, with a very similar trend in both series of chromophores (9b>2b>7b>10b>3b>6b). The phenyl-substituted linker in these chromophores represents a conjugated system that extends the electron delocalization of the heterocycle. The HOMO is extended over the entire πsystem and mainly involves the heteroatom electron pairs of the substituent joined to the aromatic ring. For this reason, the oxidation is strongly influenced by the electron-donor 14

ACCEPTED MANUSCRIPT character of the substituents in the para-position of the phenyl group attached to the quinolizinium cation. Strongly donating substituents, such as dimethylamino (3b, 6b, 10b), lead to more negative reduction potentials (and less positive oxidation potentials) than those obtained when weaker donor substituents (2a, 7b, 9b) are present in the same position. The LUMO is

RI PT

located on the nitrogen atom in the quinolizinium cation core. This unit is only weakly influenced by the electron-donor character of substituents located at the end of the π-system and the reduction potentials for 3b, 6b, 10b, 2a, 7b and 9b are therefore very similar.

The oxidation and reduction of a functional group in a given compound proceeds more easily

SC

when it has a ‘lower’ (in absolute value) oxidation or reduction potential, respectively. Maps of the HOMO and LUMO orbitals of the chromophores under investigation show that reduction at

M AN U

the nitrogen atom in the quinolizinium cation involves the entire π-system when the reduction occurs. However, the more limited delocalization of the π-electron system in 4-methoxyderivative 9b results in a more negative reduction potential (and more positive oxidation potential) than those in 2b and 7b (see Figure 10). In turn this gives rise to a larger gap between

TE D

the Eox and Ered values for 9b in comparison with 2a and 7b. The order of the HOMO-LUMO gap should be: 9b > 2a > 7b.

Compound 3b, in which the quinolizinium salt is substituted at the 2-position with an acetylenic

EP

chain joined to a 2-(4-dimethylamino phenyl) ring, provides a cyclic voltammetry Epa value of +0.94 V, which correlates with the expected reluctance of this compound to undergo oxidation

AC C

compared to the homologous ethylenic compound 6b (Epa = +0.7 V). Stabilization of the electrogenerated radical cation in 3b is more difficult than in 6b because it cannot be highly delocalized. The oxidation of 3b should take place directly at the nitrogen atom rather than the acetylenic moiety, as indicated in Scheme 4. Cyclic voltammetry results for 9b can be compared with those for 7b. The latter alkenylbridged chromophore is much easier to oxidize. The absence of a connecting chain between the quinolizinium ring and the aryl group in 9b (Epa = +1.85 V) produces a poorly stabilized radical cation (Scheme 5) where the contribution of the resonance

15

ACCEPTED MANUSCRIPT forms with an oxygen atom supporting the positive charge is tiny. In this case the aromatic ring is oxidized but is also slightly stabilized by the alkoxy group.

7b

+1.42

9b 3b

+1.85 +0.94

6b 10b

+0.71 +0.98

+1.57

+0.87 +1.32

-1.24

-6.56

-5.97

-1.35 -1.25

-7.19 -5.92

-6.25 -5.40

-1.28 -1.47

-5.58 -5.97

-5.15 -5.47

RI PT

Table 2. Experimental values of the peak potentials E (V, vs Ag/Ag+) (± 0.03V) of selected D-π-A+ and D-A+ chromophores (Figure 9). Scan rate: 100 mV/s. Energy Energy Comp. Epa1 Epa2 Epc ∆Eb/eV HOMOa/eV HOMOb /eV +1.67 -1.27 -6.94 -5.98 -2.52 2b -2.27

-2.28 -2.33

-2.21 -2.06

SC

a:Estimated from CV data (Epa1 values) by applying the relationship by Forrest[33].

Me

Me

N

N

Me

Me N

Me

-e N

N

N

PF6-

PF6-

3b

Me

Me

N

N Me

-e

-

N

N PF6-

Me

6b

PF6-

Me

Me

Me

N

N

Me

N

N

PF6-

PF6-

TE D

PF6-

M AN U

b: B3LYP/6-31g(d) calculations including the solvent effect.

Me

OMe

-e N

9b

OMe

OMe

N

N

PF6-

PF6-

AC C

PF6-

EP

Scheme 4. Comparative delocalization D-π-A+ chromophores 3b and 6b

Scheme 5. Delocalization for chromophore 9b

The redox abilities of the aforementioned quinolizinium salts were investigated with the main focus on the first oxidation and first reduction potentials, since these values are related to the HOMO and LUMO energies, respectively. The electronic influence of individual substituents was followed in an effort to understand the relationship between redox properties and structure. The reduction of these chromophores occurs on the quinolizinium heteroatom and is

16

ACCEPTED MANUSCRIPT consequently not strongly influenced by the donor/acceptor character of the substituents. However, this process is rather dependent on the extended π-delocalized electron system involving stabilization of the electrogenerated radical intermediate. Whereas in the presence of strong donating groups these compounds are reduced more

RI PT

negatively (Er(10b) = –1.47V), the weaker donor substituents produce a shift to more positive reduction potentials (Er(9b) = –1.35V). The reduction of 9b therefore proceeds more easily than that of 10b.

The oxidation of these molecules occurs on the methoxy/amino group located at the para-

SC

position of the benzene ring, in opposite side. The oxidation potential strongly reflects the donor/acceptor character of the substituents, although the entire π-system chain is also

M AN U

considerable. On increasing the electron-donating character (Me2N related to MeO) and the extent of the conjugation in the molecule (cf. 6b and 10b), less positive oxidation potentials are

AC C

EP

TE D

registered and this is consistent with the easier oxidation of these chromophores.

Figure 10. Cyclic voltammograms of 2b, 7b and 9b in acetonitrile/LiClO4 (0.1M) as SSE, at Pt as working and auxiliary electrodes. Ag/Ag+ (sat) reference electrode. Scan rate: 100 mV/s. (see SI voltammograms of 3b, 6b and 10b).

2.4 Non-linear Optical Properties: Hyper-Rayleigh Scattering (HRS) Measurements and Computational Details.

17

ACCEPTED MANUSCRIPT The results of Femtosecond hyper-Rayleigh scattering measurements performed at 800 nm on selected compounds are compiled in Table 3 and these confirm the potential of these small ionic compounds for second-order non-linear optics. Congruent with the trends already observed in the linear optical (absorption) experiments, the first hyperpolarizability of the compounds is

RI PT

higher when the methoxyphenyl group is present as the aryl donor and the most marked effect is seen when the quinolizinium is substituted in the C-2 position with a π-linker (2b) and in the DA+ unbridged chromophore 9b, which was the chromophore with the highest first

SC

hyperpolarizability value when compared with D-π-A+ compounds.

Table 3 Experimental non-linear Optical Properties of some Chromophores 2–10 λmax

βHRS

βzzz

350

122 ± 10

300 ± 20

60 ± 4

2b

378

284 ± 14

690 ± 40

152 ± 5

2c

350

245 ± 20

600 ± 40

112 ± 6

2d

380

202 ± 18

490 ± 40

37 ± 3

4b

367

380 ± 60

918 ± 140

110 ± 20

4c

334

74 ± 3

178 ± 7

45 ± 2

5b

350

186 ± 28

350 ± 70

86 ± 12

5c

341

34 ± 8

80 ± 12

18 ± 2

6a

425

142 ± 7

343 ± 17

31 ± 2

6b

472

199 ± 6

482 ± 51

49 ± 5

6c

433

191 ± 28

462 ± 49

56 ± 6

6d

445

83 ± 6

202 ± 42

38 ± 8

7b

382

118 ± 17

285 ± 41

19 ± 2

7c

353

162 ± 16

391 ± 39

70 ± 7

AC C

EP

TE D

M AN U

Comp. . 2a

βzzz, 0

8b

347

30 ± 7

72 ± 17

15 ± 3

9b

360

500 ± 50

1190 ±120

180± 20

9c

337

61 ± 7

147 ± 17

35 ± 4

10b

431

286 ± 30

450 ± 70

50 ± 7

Maximum absorption wavelength λmax (nm) predicted by CIS(D), resonance enhanced HRS first hyperpolarizability βHRS (10–30 esu), resonance enhanced diagonal component of the molecular first hyperpolarizability βzzz (10–30 esu), off-resonance diagonal component of the molecular first hyperpolarizability βzzz,0 (10–30 esu).

18

ACCEPTED MANUSCRIPT The first hyperpolarizability value in D-π-A+ chromophores is higher in compound 2b (β = 284.10–30 esu) than in 2c and 2d, which have β values of 245.10–30 esu and 202.10–30 esu, respectively. The replacement of the donor unit (N,N-dimethylaminostyryl) at the C2 position with a double bond resulted in 6b (β = 199.10–30 esu), the chromophore with the highest first

RI PT

hyperpolarizability value when compared with compounds in which quinolizinium was substituted at the C1, C3 and C4 positions with the same substituent. In agreement with the greater influence of the position of the donor substituent at the quinolizinium acceptor, which is more marked than the relative electron-donating strength characteristics, and the better

SC

conjugation and stabilization, the highest β values correspond to the compounds with substituents at the C2 positions, thus confirming that the best conjugation and charge transfer

M AN U

from donor to acceptor is achieved in these positions.

The experimental results reported here for cationic chromophores (D-A+), when compared to the (D-π-A+) systems reported previously by us, reveal that the hyperpolarizability of (D-A+) molecules seems to be determined by the strength of the charge-transfer. Thus, the highest first

TE D

hyperpolarizability values were found for the dimethoxystyryl donor unit at the C2 (9b: β = 500·10–30 esu) and the N,N-dimethylaminostyryl donor unit at the C2 (10b: 286·10–30 esu). However, in the case of 9b, the common assumption that β can be maximized by using a

EP

conjugated bridge[36] to link D and A acceptor units is not clear because, surprisingly, the highest hyperpolarizability was found for 9b with dimethoxystyril donor unit connected directly to

AC C

quinolizinium acceptor when compared with 2b and 7b. Theoretical calculations present a similar

correlation for the resonant and off-resonant molecular hyperpolarizabilities βzzz and βzzz,0, respectively.

Theoretical calculations were performed using the Gaussian suite of quantum chemical programs[37]. For all of the studied cationic chemical systems, the molecular structure used to compute optical properties corresponded to an energy minimum calculated at the HF/6-31G(d) level of theory. These structures were confirmed as minima by evaluation of the harmonic vibrational frequencies, which were all real.

19

ACCEPTED MANUSCRIPT Table 4. Theoretical linear and nonlinear optical properties selected compounds λmax exp.

λmax Cal. CIS(D)

βHRS

βzzz,800

βzzz, 0

2a

350

312

83

197

79

2b

378

364

255

617

187

2c

350

323

131

-318

-126

2d

380

336

104

-276

-96

3b

447

449

233

429

297

9b

360

334

63

154

10b

431

414

209

506

RI PT

Comp.

64

152

Maximum absorption wavelength λmax (nm) predicted by CIS(D), resonance enhanced HRS first

SC

hyperpolarizability βHRS (10–30 esu), resonance enhanced diagonal component of the molecular first

hyperpolarizability βzzz (10–30 esu), off-resonance diagonal component of the molecular first

M AN U

hyperpolarizability βzzz,0 (10–30 esu).

Calculated wavelengths (λmax, in nm) were estimated by the CIS(D)[38] method, which generally provides accurate predictions of excited states that are mainly one-electron transitions from a single reference ground state. The λmax values were obtained as vertical excitations from the ground state to the first excited singlet.

TE D

As CIS(D) is one of the simplest approaches amongst the correlated excited state theories, our calculations provided good results with wavelength values that differed by less than 10% with respect to the experimental data, as predicted in other

EP

investigations[39]. A more pronounced overestimation of the excitation energies by the quantum chemical methods was observed for TDDFT calculated values.

AC C

Theoretically predicted values of the second harmonic generation (SHG) of hyperpolarizability [40] β(-2ω;ω,ω) were obtained, including the electron correlation effect at the MP2 level. Molecular components of the first hyperpolarizability were computed as dipole moment second derivatives and these values were then spatially averaged to obtain the βHRS values[41]. Details of the mathematical procedure to obtain the hyper-Rayleigh hyperpolarizability from molecular first hyperpolarizabilities are reported elsewhere[21a,b]. Both the resonance βzzz and off-resonance βzzz,0 diagonal components of the molecular first hyperpolarizability were directly calculated as dipole moment derivatives rather than being estimated from the simple two-state model. As a consequence, comparisons could not be made between these two sets of data.

20

ACCEPTED MANUSCRIPT These MOs (Figure 11) clearly show the charge transfer from the donor to the acceptor fragments of these systems. The π-π* transition bands are due to the vertical excitation between these MOs (see SI Figure 11). The discrepancies between the theoretical energy gap (∆E) values and vertical excitation (λmax) for cation 9b shows deficiencies for this HF wavefunction that are not

M AN U

SC

RI PT

completely corrected in a third order property like hyperpolarizability.

Figure 11. HOMO and LUMO of cation 3b

2.5 Crystallographic studies

The crystal and molecular-ion structures of compounds 9b and 9c were obtained by X-

TE D

ray diffraction studies (Figure 12). The PF6– counter ion in both cases is found in the crystal but omitted in tables and figures for the sake of clarity. Compound 9b crystallizes in the monoclinic P21/c space group and 9c in the triclinic P-1 space group, both of which are centrosymmetric.

EP

Selected bond distances and angles are shown in Table 5 (see SI) for both molecules. It is

AC C

interesting to note that the distance C(2)–C (11) in 9b and C (3)–C(11) in 9c are 1.485(5) and 1.481(4) Å, respectively. These are slightly shorter than the normal[42] C–C single bond (1.54(3) Å), which indicates π-conjugation between the donor and quinolizinium system. In compounds 9b and 9c the aryl group is nearly coplanar with the quinolizinium system, with angles between the planes of the donor and acceptor units of 11.87(15)° and 13.91(11)°, respectively. The practically coplanar arrangement in the chromophore corroborates the delocalization and charge transfer between the aryl and quinolizinium, which in turn explains the high hyperpolarizability values of D-A+ unbridged chromophores 9b – as predicted by the two-levels model[43]. 21

ACCEPTED MANUSCRIPT Observation of the dihedral angles between the phenyl and methoxy group for both molecules shows that for 9b the dihedral angle C(15)–C(14)–O(2)–C(17) is 1.5° and for 9c C(13)–C(14)–O(1)–C(17) is 6°. This difference probably indicates that the planarity of 9b is somewhat higher and this promotes better charge transfer (βHRS = 500±50 10–30

SC

RI PT

esu for 9b and βHRS = 61±7 10–30 esu for 9c).

TE D

M AN U

Figure 12. ORTEP drawings of compounds 9b and 9c. The PF6– counter ion has been omitted for clarity

EP

Figure 13. Weak intermolecular interactions compounds 9b and 9c

3. Experimental section

AC C

3.1 Materials and Methods

3.1.1 Chemistry. Reactions were monitored by thin-layer chromatography (TLC) using TLC silica gel-coated aluminium plates 60F-254 (Merck). Column chromatography was performed using Merck silica gel 60, 0.040–0.063 mm (230–400 mesh). NMR spectroscopy was performed on Varian UNITY-Plus 300 (1H and 75 MHz 13C) or VNMRS-500 (1H and 125 MHz 13

C) spectrometers equipped with a cryoprobe. Chemical shifts were reported as δ values (ppm)

and coupling constants (J) in Hz. The mass spectra (MS) as ESI+ and HRMS masses were recorded on a Thermo Scientific TSQ Quantum LC/MS equipped with an ESI ionization source and a TOF detector. All starting materials, namely Pd(OAc)2, CuI, PdCl2(PPh3)2, 4-methoxy22

ACCEPTED MANUSCRIPT and 4-N,N-dimethylamino- benzaldehyde, 1-ethynyl-4-methoxybenzene, and the corresponding potassium organotrifluoroborates, were purchased from Aldrich and were used without further purification. The synthesis of 1-bromo-, 2-bromo-. 3-bromo- and 4-bromoquinolizinium [28] and 2-Methylquinolizinium salts [29] were obtained by previously reported methods. For

RI PT

chromophores 2a–d, 4b,c and 5b,c see reference 24 and for the preparation of the (E)-2(arylvinyl)quinolizinium hexafluorophosphates by Heck reaction and Knoevenagel reaction see references 18b and 25a. The procedure with potassium organotrifluoroborates is described in reference 27.

SC

3.1.2 General optical properties. Absorption Spectra were recorded on a Perkin-Elmer L35 UVVis spectrophotometer in the 200–1000 nm range. Steady-state fluorescence measurements

M AN U

were performed by using an SLM 8100 AMINCO spectrofluorimeter equipped with polarizers and a double (single) concave grating monochromator in the excitation (emission) path and a cooled photomultiplier. Slit widths were set at 8 nm for excitation and emission and polarizers at the magic angle. Fluorescence decay measurements were performed on an Edinburgh

TE D

Instruments FL900 time-correlated single-photon-counting system.

The fluorescence quantum yield was calculated using equation 1 (E1):  grad i   ni2   2   grad s t   ns t 

φi = φs t 

E1

EP

where φst is the fluorescence quantum yield of the standard (st) in each case, gradi/st are the slope

AC C

of the surface plot under the full emission spectrum versus absorption at the excitation wavelength of derivative i or standard, respectively, and ni/st are the refractive indexes of methanol and water (or ethanol), respectively. 3.1.3 General voltammetric measurements. Voltammetric curves with IR compensation were obtained with a Tracelab Radiometer POL150 polarographic analyser in a Voltalab MDE150 Polarographic Stand in a three-electrode cell. Two Pt wires were used as working and auxiliary electrodes. The electrochemical properties of the chromophores indicated in Figure 9 were studied by cyclic voltammetry (CV) in acetonitrile (ACN)/LiClO4 as SSE (solvent-supporting electrolyte 23

ACCEPTED MANUSCRIPT systems). Before the voltammetric experiments, the solutions were deaerated by bubbling N2 (99.98% purity, purchased from Air Liquid) for 5 minutes. The cathodic and anodic peak potentials measured at a scan rate of 100 mV/s (summarized in Table 2) are quoted with respect to the Ag/Ag+ (sat) reference electrode.

RI PT

3.1.4 Hyper-Rayleigh scattering. General details of the hyper-Rayleigh scattering (HRS) experiments were discussed previously[44]. The HRS measurements were performed at room temperature in methanol, with crystal violet as the reference molecule and with high-frequency demodulation of the multiphoton fluorescence contribution[45].

SC

The HRS signal was analyzed towards a single major dipolar hyperpolarizability tensor element βzzz along the molecular z-axis. The dynamic on resonantly enhanced βzzz,800 value obtained at

M AN U

800 nm was reduced to the static, or off-resonance βzzz.0, value by applying the classical twolevel model[39]. From the fitting of the apparent βzzz,800 as a function of modulation frequency, a fluorescence lifetime could be obtained along with the accurate fluorescence-free hyperpolarizability value[46]. The reported β values are the averages taken from measurements

TE D

at different amplitude modulation frequencies.

Theoretical calculations: The theoretical resonant or frequency-dependent hyperpolarizabilities were obtained at the MP2/6-31G(d) level and include the electron correlation effect in this NLO

EP

property βMP2(-2ω;ω,ω). This property was estimated from the resonant or dynamic β(-2ω;ω,ω) at the HF/6-31G(d) level, and the static or non-resonant β(0;0,0) were calculated at both the

AC C

HF/6-31G(d) and MP2/6-31G(d) levels following the multiplicative approximation scheme, where the frequency dispersion is estimated to be similar at these two levels of theory[39a,47]. β MP 2 (− 2ω;ω , ω ) ≈ β HF (− 2ω;ω , ω )

β MP 2 (0;0,0) β HF (0;0,0)

The first hyperpolarizability β was calculated as a second dipole moment derivative with respect to electric fields. At the Hartree–Fock (HF) level these dipole derivatives were calculated analytically by solving the corresponding CPHF equations. When the electron correlation effect was included by means the perturbative MP2 approach, these dipole derivatives were obtained numerically. Calculated hyperpolarizabilities were obtained on 24

ACCEPTED MANUSCRIPT the molecular axes basis βijk and then transformed to the laboratory axes βXYZ by applying the expressions of Civin et al[48]. 3.1.5 Single-crystal X-ray structure determinations of 9b and 9c. Selected bond distances and angles are shown in Table 5 (see ESI) for both molecules. Details of the X-ray

RI PT

experiments, data reduction and final structure refinement calculations are summarized in Table 6 (see ESI). Single crystals suitable for X-ray diffraction were selected for data collection. The crystals were mounted on a glass fibre using an inert perfluorinated ether oil and placed in a low temperature N2 stream (200(2) K) in a Bruker-Nonius Kappa

SC

CCD single crystal diffractometer equipped with a graphite-monochromated Mo-Kα radiation source (λ = 0.71073 Å) and an Oxford Cryostream 700 unit. The structures

M AN U

were solved by direct methods (SHELXS-97) using the WINGX package[49] and completed by subsequent difference Fourier techniques and refined by using full-matrix least-squares against F2(SHELXL-97)[50]. All non-hydrogen atoms were anisotropically refined. Most of the hydrogen atoms were geometrically placed and left riding on their

TE D

parent atoms, and others were found in the Fourier difference maps. In the case of 9b the PF6– ion was found to be within the crystal with evident structural disorder, which was modelled.

EP

Crystallographic data (excluding structure factors Å) for the structures reported in this paper have been deposited with the Cambridge Crystallographic Data Centre as supplementary

AC C

publication nos. CCDC 1523732-1523733. Copies of the data can be obtained free of charge on application to CCDC, 12 Union Road, Cambridge CB2 1EZ, UK (fax: (+44)1223-336-033; email: [email protected]).

3.2 Synthesis 3.2.1

General

procedure

for

the

preparation

of

arylethynyl-quinolizinium

hexafluorophosphate. A flame-dried vial was charged under argon with 50 mg (0.173 mmol) of the corresponding bromoquinolizinium salt, 10 mol % CuI (3.3 mg, 0.0173 mmol) and 5 mol % PdCl2(PPh3)2 (6.9 mg, 0.0087 mmol) in dry DMF (2.5 mL). 125

ACCEPTED MANUSCRIPT Ethynyl-4-methoxybenzene (0.208 mmol, 27 µL) and Et3N (0.259 mmol, 45 µL) were added. After heating at 60 °C for 14 h, the solution was filtered through a small pad of Celite® and washed with methanol (5 mL). The solution was concentrated, treated with saturated solution NaHCO3 (20 mL) and extracted with EtOAc (15 mL). The organic

RI PT

phase was extracted with H2O (2 × 10 mL), the aqueous phase was treated with saturated solution NH4PF6 the resulting phase was extracted with CH2Cl2 (3 × 15 mL). The organic phase was dried over Na2SO4, the solvent evaporated under reduced pressure and the

(9.6:0.4) as eluent.

hexafluoro

M AN U

2-(4-N,N-Dimethylamino-phenylethynyl)quinolizinium

SC

product was purified by column chromatography on silica gel using CH2Cl2/MeOH

phosphate

(3b).

Following the general procedure, from 4-ethynyl-N,N-dimethylaniline (0.2077 mmol, 30 mg) stirring at room temperature for 5 h and using CH2Cl2/MeOH (9:1) afforded 39.7 mg (55%) of 3b as a brown-red solid: mp 273–277 °C; IR (KBr) νmax (cm–1) 3345, 2190, 1644, 1600, 1375,

TE D

837. 1H-NMR (500 MHz, Acetone-d6) δ (ppm) 9.30 (d, J = 7.0 Hz, 1H), 8.56 (d, J = 1.5 Hz, 1H), 8.52 (d, J = 8.6 Hz, 1H), 8.43 (ddd, J = 8.6, 7.2, 1.0 Hz, 1H), 8.13–8.06 (m, 1H), 7.51 (d, J = 9.0 Hz, 1H), 6.80 (d, J = 9.0 Hz, 1H), 3.07 (s, 3H). 13C-NMR (125 MHz, acetone) δ 154.61,

EP

139.99, 139.58, 139.21, 136.67, 135.63, 129.52, 127.65, 126.03, 114.46, 108.82, 107.31, 88.16,

AC C

41.81. HRMS (ESI-TOF, MeOH) Calcd. for C19H17N2 [M+]: 273,1386 [M+] Found: 273.1384.

3.2.2 General procedure for the preparation of the quinolizinium salts (9–10). Suzuki reaction

Method A. A flame-dried 10 mL vial was charged under argon with the corresponding quinolizinium bromide (50 mg, 0.173 mmol), 2 mol % of Pd2(dba)3 (3.2 mg, 0.0035 mmol), 4 mol % P(o-Tol)3 (2.1 mg, 0.0069 mmol), 1.4 equiv. of K2CO3 (36.8 mg, 0.242 mmol), and 1.3 equiv. of the boronic acid (0.225 mmol) in dry DMF (2.5 mL). After stirring at room temperature as indicated, the reaction mixture was filtered through a small pad of Celite® and washed with MeOH (5 mL). The solution was concentrated, treated with H2O (15 mL) and 26

ACCEPTED MANUSCRIPT extracted with AcOEt (10 mL). The organic phase was extracted with H2O (2 × 10 mL), the aqueous phase was treated with saturated solution NH4PF6 and the resulting phase was extracted with CH2Cl2 (3 × 15 mL). The organic phase was dried over Na2SO4, the solvent was evaporated under reduced pressure and the product was purified by column chromatography on silica gel

RI PT

using CH2Cl2/MeOH (9.6:0.4) as eluent. Method B. 4 mol % of Pd2(dba)3 (6.3 mg, 0.0069 mmol), 8 mol % of P(o-Tol)3 (4.2 mg, 0.0138 mmol).

SC

2-(4-Methoxyphenyl)quinolizinium hexafluorophosphate (9b). From 4-methoxyphenyl boronic (34.2 mg, 0.225 mmol) and stirring the reaction mixture for 23 h afforded 34.5 mg

M AN U

(52%, method A) and 47.8 mg (73%, method B) of 9b as a white solid. mp. 230–232 °C. IR (KBr): υmax (cm–1) 3122, 2933, 1645, 1606, 1405, 1178, 836, 558. 1H NMR (300 MHz, Acetone-d6) δ (ppm) 9.38 (d, 1H, J = 7.3 Hz), 9.31 (d, 1H, J = 6.8 Hz), 8.90 (d, 1H, J = 2.0 Hz), 8.59–8.52 (m, 2H), 8.40 (td, 1H, J = 8.5, 1.1 Hz), 8.14–8.04 (m, 3H), 7.21 (d, 2H, J = 9.0 Hz),

TE D

3.93 (s, 3H). 13C NMR (75 MHz, Acetone-d6) δ (ppm) 163.3, 148.3, 144.2, 137.6, 137.4, 136.9, 130.1, 127.9, 127.1, 123.7, 122.2, 122.1, 115.9, 56.0. MS (ESI+) m/z (relative intensity) 236 (M+, 100). Anal. Calcd. For C16H14F6NOP (381.26 g/mol) C (50.41), H (3.70), N (3.67) Found:

EP

C (50.49), H (3.85), N (3.54).

3-(4-Methoxyphenyl)quinolizinium hexafluorophosphate (9c). Following Method B, from 4-

AC C

methoxyphenylboronic acid (34.2 mg, 0.225 mmol) and stirring the reaction mixture for 22 h afforded 39.6 mg (60%) of 9c as a pale yellow solid: 217–218 °C. IR (KBr): υmax (cm–1) 3113, 2942, 1609, 1509, 1401, 1292, 1257, 1203, 835, 558. 1H NMR (300 MHz, Acetone-d6) δ (ppm) 9.70 (s, 1H), 9.42 (d, 1H, J = 6.8 Hz), 8.80 (dd, 1H, J = 9.0, 1.8 Hz), 8.69 (d, 1H, J = 9.5 Hz), 8.65 (d, 1H, J = 10.1 Hz), 8.45 (ddd, 1H, J = 9.7, 7.3, 1.1 Hz), 8.21 (t, 1H, J = 6.0 Hz), 7.95 (d, 2H, J = 8.8 Hz), 7.19 (d, 2H, J = 8.8 Hz), 3.91 (s, 3H). 13C NMR (75 MHz, Acetone-d6) δ (ppm) 162.6, 142.5, 137.6, 137.5, 137.3, 136.7, 133.9, 129.9, 128.2, 128.0, 126.6, 125.3, 116.0, 55.9.

27

ACCEPTED MANUSCRIPT MS (ES+) m/z (relative intensity) 236 (M+, 100). Anal. Cal. For: C16H14F6NOP (381.26 g/mol) C (50.41), H (3.70), N (3.67) Found : C (50.64), H (3.53), N (3.78). 2-(4-N,N-Dimethylaminophenyl)quinolizinium hexafluorophosphate (10b). Procedure B, from 4-N,N-dimethylaminophenylboronic (37.1 mg, 0.225 mmol), and stirring the reaction

RI PT

mixture for 20 h, gave 49.7 mg (73%) of 10b as a red solid. mp: 242–244 °C (Acetone:Et2O). IR (KBr): υmax (cm–1) 2927, 1649, 1598, 1406, 1214, 842, 559. 1H NMR (300 MHz, Acetone-d6) δ (ppm) 9.24 (d, 1H, J = 7.3 Hz), 9.17 (d, 1H, J = 6.6 Hz), 8.77 (d, 1H, J = 2.2 Hz), 8.47 (dd,

1H, J = 7.1, 2.2 Hz), 8.44 (d, 1H, J = 10.4 Hz), 8.27 (td, 1H, J = 9.9, 1.1 Hz), 8.03 (d, 2H, J =

SC

9.2 Hz), 7.91 (td, 1H, J = 6.8, 1.0 Hz), 6.94 (d, 2H, J = 9.2 Hz), 3.12 (s, 6H). 13C NMR (75

M AN U

MHz, Acetone-d6) δ (ppm) 153.8, 149.0, 144.4, 137.3, 136.7, 136.5, 129.7, 127.7, 122.6, 121.3, 120.8, 119.5, 113.1, 40.0. MS (ESI+) m/z (relative intensity) 249 (M+, 100). Anal. Calcd. for C17H17F6N2P (394.30 g/mol) C (51.78), H (4.35), N (7.10), found C (51.64), H (4.50), N (7.22). 3-(4-N,N-Dimethylaminophenyl)quinolizinium hexafluorophosphate (10c). Method B, from 4-N,N-dimethyl aminophenylboronic acid (37.1 mg, 0.225 mmol), stirring the reaction mixture

TE D

for 30 h, gave 41.0 mg (60%) of 10c as an orange solid: mp 229–230 °C (Acetone:Et2O).IR (KBr): υmax (cm–1) 2930, 1605, 1511, 1401, 1216, 842, 559. 1H NMR (300 MHz, Acetone-d6) δ (ppm) 9.62 (s, 1H), 9.37 (d, 1H, J = 6.6 Hz), 8.77 (dd, 1H, J = 9.2, 1.8 Hz), 8.61 (d, 1H, J = 8.8

EP

Hz), 8.58 (d, 1H, J = 6.4 Hz), 8.36 (td, 1H, J = 9.7, 1.1 Hz), 8.15 (t, 1H, J = 7.1 Hz), 7.85 (d, 2H, J = 9.0 Hz), 6.94 (d, 2H, J = 9.0 Hz), 3.07 (s, 6H).

13

C NMR (75 MHz, Acetone-d6) δ

AC C

(ppm). 152.7, 141.8, 137.9, 137.1, 136.3, 135.9, 132.1, 129.0, 127.9, 127.8, 125.1, 120.4, 113.4, 40.1. MS (ES+) m/z (relative intensity) 249 (M+, 100). Anal. Calcd. for C17H17F6N2P (394.30 g/mol) C (51.78), H (4.35), N (7.10). Found: C (51.90), H (4.55), N (7.02).

3.2.3 General procedure for the preparation of the quinolizinium with organotrifluoroborates (9b and 10b) Method C. The corresponding potassium organotrifluoroborate (0.2076 mmol, 1.2 equiv) was added to a mixture of the corresponding bromoquinolizinium bromide (0.1730 mmol), K2CO3

28

ACCEPTED MANUSCRIPT (0.0857 g, 0.5190 mmol, 3 equiv), 1 mol % of Pd(OAc)2 (0.0004 g, 0.00017 mmol) in H2O (2.5 mL). The reaction mixture was heated at 60–65 °C with stirring for the time indicated (2–4.5 h). The mixture was then cooled to room temperature and treated with saturated solution NH4PF6. The resulting solid was filtered off and washed with water and diethyl ether. For a further

RI PT

purification, the solid was chromatographed on silica gel using CH2Cl2/MeOH) (9:1) as eluent. 2-(4-Methoxyphenyl)quinolizinium hexafluorophosphate (9b). Following the general procedure and using potassium 4-methoxyphenyltrifluoroborate (44.4 mg) gave 49.8 mg (76%) of 9b as a white solid.

SC

2-(4-N,N-Dimethylaminophenyl)quinolizinium hexafluorophosphate (10b). Following the general procedure and using potassium 4-(N,N-dimethylamino)phenyltrifluoro borate (47.1 mg)

M AN U

gave 54.5 mg (80%) of 10b as a red solid.

2-(4-Methoxyphenyl)quinolizinium hexafluorophosphate (9b). Following the general procedure and using potassium 4-methoxyphenyltrifluoroborate (44.4 mg), were obtained 49.8 mg (76%) of 9b as a white solid.

TE D

2-(4-N,N-Dimethylaminophenyl)quinolizinium hexafluorophosphate (10b). Following the general procedure and using potassium 4-(N,N-dimethylamino )phenyltrifluoro borate (47.1

4. Conclusions

EP

mg), were obtained 54.5 mg (80%) of 10b as a red solid.

AC C

In summary, we have prepared a series of D-A+ and D-π-A+ chromophores based on quinolizinium using diffferent palladium-promoted C–C bond formation reactions. The optical properties of these chromophores were investigated by absorption and emission spectroscopy. The spectra of compounds with electron-rich substituents at the C2 position of the quinolizinium show a strong effect on the stabilization energy of the chromophores and a bathochromic effect is observed in the absorption spectrum, which is consistent with a better conjugation or a lower energy gap. The fluorescence spectra also presented displacements with a similar trend.

29

ACCEPTED MANUSCRIPT The electrochemical behaviour of the chromophores is mainly affected by the electronic nature of the donor/acceptor character of the substituents, although the entire π-system chain is also considerable. On increasing the electron-donating character and the extent of conjugation of the molecule, less positive oxidation potentials were registered. This finding is consistent with the

RI PT

easier oxidation of these compounds. The calculated HOMO/LUMO energies and their differences correlate with the electrochemical measurements obtained by cyclic voltammetry. The experimental first hyperpolarizabilities (βHRS) of the series of charged chromophores show that compounds 2b, 3b, 9b and 10b have significantly higher values than the rest of the

SC

synthesized cationic systems. The low transition energy and high CT level are the decisive factors that provide a higher first hyperpolarizability in compounds substituted at C2.

M AN U

Theoretical calculations support the experimental results and also predict a similar trend. However, chromophore 9b showed some discrepancies between the theoretical calculation and experimental results. Further investigation of 9b by X-ray analysis revealed significant planarization.

TE D

In comparison to previously reported D-π-A+ molecules[21a], the lack of a conjugated bridge in these D-A+ chromophores does not seem to be significant in terms of obtaining higher βHRS values, with the magnitude of β in this kind of chromophore being dominated by additive

EP

contributions from the donors and the acceptors. The results highlight an attractive strategy in guiding the design of new NLO materials, with unbridged D-A+ chromophores allowing a

AC C

remarkable degree of control of the linear and non-linear optical response by changing the position and electronic nature of the substituents and, more importantly, the nature of the linker.

Acknowledgements

Financial support from the Spanish Ministerio de Ciencia y Competitividad (projects MICINN/CTQ2011-24715 and MINECO/ CTQ2014-52488-R and CTQ2015-64425-C2-1-R and CTQ2016-80600-P). Instituto de Salud Carlos III (FEDER funds, ISCIII RETIC REDINREN RD012/20021/0014) are gratefully acknowledged

30

ACCEPTED MANUSCRIPT References

AC C

EP

TE D

M AN U

SC

RI PT

[1] (a) Dini D, Calvete MJF, Hanack M. Nonlinear Optical Materials for the Smart Filtering of Optical Radiation. Chem Rev. 2016;116:13043-233. (b) Yang H, Tang R, Wu W, Liu W, Guo Q, Liu Y, et al. A series of dendronized hyperbranched polymers with dendritic chromophore moieties in the periphery: convenient synthesis and large nonlinear optical effects. Polym Chem. 2016;7:4016-24. (c) Cambre S, Campo J, Beirnaert C, Verlackt C, Cool P, Wenseleers W. Asymmetric dyes align inside carbon nanotubes to yield a large nonlinear optical response. Nat Nanotechnol. 2015;10:248-52. (d) Kuzyk MG. Using fundamental principles to understand and optimize nonlinear-optical materials. J Mater Chem. 2009;19:7444-65. (e) Nonlinear Optical Properties of Matter: From Molecules to Condensed Phases; Papadopoulos MG, Leszczynski J, Sadlej AJ Eds. Springer: Dordrecht, 2006. (f) Marder SR. Organic nonlinear optical materials: where we have been and where we are going. Chem Commun. 2006:131-4. [2] (a) Wojciechowski, A.; Raposo, M. M. M.; Castro, M. C. R.; Kuznik, W.; Fuks-Janczarek, I.; Pokladko-Kowar, M.; Bures, F. Nonlinear optoelectronic materials formed by push-pull (bi)thiophene derivatives functionalized with di(tri) cyanovinyl acceptor groups. J Mater Sci: Mater Electron. 2014; 25:1745-50. [b] Elder, D. L.; Benight, S. J.; Song, J.; Robinson, B. H.; Dalton, L. R. Matrix-Assisted Poling of Monolithic Bridge- Disubstituted Organic NLO Chromophores. Chem. Mater. 2014; 26:872-74. [c] Dalton LR, Sullivan PA, Bale DH. Electric Field Poled Organic Electro-optic Materials: State of the Art and Future Prospects. Chem Rev. 2010;110:25-55. [3] (a) Hales JM, Barlow S, Kim H, Mukhopadhyay S, Bredas J, Perry JW, et al. Design of Organic Chromophores for All-Optical Signal Processing Applications. Chem Mater. 2014;26:549-60. [b] Gieseking RL, Mukhopadhyay S , Risko C, Marder SR, Brédas JL. Design of Polymethine Dyes for AllOptical Switching Applications: Guidance from Theoretical and Computational Studies. Adv Mater. 2014; 26:68–84 [c] He GS, Zhu J, Baev A, Samoc M, Frattarelli DL, Watanabe N, Facchetti A, Ågren H, Marks TJ, Prasad PN.Twisted π-System Chromophores for All-Optical Switching. J Am. Chem Soc 2011;133:6675–80. [4] (a) Qi J, Qiao W, Wang ZY. Advances in organic near-infrared and emerging applications. Chem Rec 2016;16:1531–48. (b) Jana A, Bai L, Li X, Agren H, Zhao Y. Morphology Tuning of Self-Assembled Perylene Monoimide from Nanoparticles to Colloidosomes with Enhanced Excimeric NIR Emission for Bioimaging. ACS Appl Mater Interfaces. 2016;8:2336-47. (c) Kim D, Ryu HG, Ahn KH. Recent development of two-photon fluorescent probes for bioimaging. Org Biomol Chem. 2014;12:4550-66. (d) Wang ZY. Near-Infrared Organic Materials and Emerging. Applications, CRC Press, Boca Raton, 2013. [5] (a) Prachumrak N, Sudyoadsuk T,Thangthong A,Nalaoh P, Jungsuttiwong S, Daengngern R, Namuangruk S, Pattanasattayavong P, PromarakV. Improvement of D–π–A organic dye-based dye-sensitized solar cellperformance by simple triphenylamine donor substitutions on the π-linker of the dye. Mater Chem Front. 2017, Advance Article DOI: 10.1039/c6qm00271d. (b) Eom YK, Kang SH, Choi IT, Yoo Y, Kim J, Kim HK. Significant light absorption enhancement by a single heterocyclic unit change in the π-bridge moiety from thieno[3,2-b]benzothiophene to thieno[3,2b]indole for high performance dye-sensitized and tandem solar cells. J Mater Chem A. 2017;5:2297-308. (c) Lee C, Lin RY, Lin L, Li C, Chu T, Sun S, et al. Recent progress in organic sensitizers for dye-sensitized solar cells. RSC Adv. 2015;5:23810-25. (d) Malytskyi V, Simon J, Patrone L, Raimundo J. Thiophene-based push-pull chromophores for small molecule organic solar cells (SMOSCs). RSC Adv. 2015;5:354-97. (e) Yen Y, Chou H, Chen Y, Hsu C, Lin JT. Recent developments in molecule-based organic materials for dyesensitized solar cells. J Mater Chem. 2012;22:8734-47. [6] (a) Lucky SS, Soo KC, Zhang Y. Nanoparticles in Photodynamic Therapy. Chem Rev. 2015;115:1990-2042. (b) Ji S, Ge J, Escudero D, Wang Z, Zhao J, Jacquemin D. Molecular Structure-Intersystem Crossing Relationship of Heavy-Atom-Free BODIPY Triplet Photosensitizers. J Org Chem. 2015;80:5958-63. (c) Zhao J, Wu W, Sun J, Guo S. Triplet photosensitizers: from molecular design to applications. Chem Soc Rev. 2013;42:5323-51. [7] (a) Li M, Zhang H, Zhang Y, Hou B, Li C, Wang X, Zhang J, Xiao HL, Cui Z, Ao Y. Facile synthesis of benzothiadiazole-based chromophores for enhanced performance of second-order nonlinear optical materials. J Mater Chem C. 2016;4:9094-102. (b) Castro MCR, Belsley M, Raposo MMM. Synthesis and characterization of push-pull bithienylpyrrole NLOphores with enhanced hyperpolarizabilities. Dyes Pigm. 2016;131:333-9. (c) Y. Yang, H. Wang, F. Liu, D. Yang, S. Bo, L. Qiu, Z. Zhen and X. Liu. The synthesis of new doubledonor chromophores with excellent electro-optic activity by introducing modified bridges. Phys Chem Chem Phys. 2015;17:5776-84. [8] (a) M. Klikar, P. Solanke, J. Tydlitat and F. Bures, Alphabet-Inspired Design of (Hetero)Aromatic Push-Pull Chromophores. Chem. Rec. 2016;16:1886–1905; (b) F. Bures, Fundamental aspects of property tuning in push–pull molecules. RSC Adv. 2014;4;58826–51;

31

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

[9] (a) Orimoto Y, Aoki Y. Strong electron correlation effects on first- and second-order hyperpolarizabilities in zwitterionic σ -conjugated systems: its dependence on substituents, conformations, spacer size, and basis sets. J Phys Chem A. 2007;111:8241-9. (b) Sitha S, Rao JL, Bhanuprakash K, Choudary BM. Enhanced Nonlinear Optical Response in Zwitterionic Molecules: A Computational Study on the Role of Orbital Interactions through σ Bonds. J Phys Chem A. 2001;105:8727-33. [10] (a) Kim S, Cho Y, Lee A, Son H, Han W, Cho DW, Kang SO. Influence of π-conjugation structural changes on intramolecular charge transfer and photoinduced electron transfer in donor–π–acceptor dyads. Phys Chem Chem Phys. 2017;19:426-35. (b) Inouchi T, Nakashima T, Kawai T. Charge Transfer Emission of T-Shaped π-Conjugated Molecules: Impact of Quinoid Character on the Excited State Properties. J Phys Chem A. 2014; 118: 2591-98. [c] Feixas F, Matito E, Poater J, Sola M. Understanding Conjugation and Hyperconjugation from Electronic Delocalization Measures. J Phys Chem A. 2011;115:13104-13. [11] (a) Bakalska R, Todorova M, Sbirkova H, Shivachev B, Kolev T. Comparing of the crystal structure and spectroscopic properties of some stilbazolium dyes with enlarged π-conjugated system I. Chromophores with pdimethylamino group. Dyes Pigm. 2017;136:919-29. (b) Guo D, Ren Y, Liu J, Tao X. Synthesis and crystal structures of imidazolium salts for quadratic optical nonlinearities. Dyes Pigm. 2013;98:384-92. (c) Todorova M, Bakalska R, Kolev T. Synthesis, crystal structure, and spectroscopic properties of new stilbazolium salt with enlarged π-conjugated system. Spectrochim Acta, Part A. 2013;108:211-22. (d) Glavcheva Z, Umezawa H, Okada S, Nakanishi H. New trimethyammonium derivatives toward second-order nonlinear optics. J Nonlinear Opt Phys Mater. 2005;14:357-65. [12] (a) Zhang Y, Zhang X, Xu X, Fu X, Hou H, Liu Y. Rigid Organization of Fluorescence-Active Ligands by Artificial Macrocyclic Receptor to Achieve the Thioflavin T-Amyloid Fibril Level Association. J Phys Chem B. 2016;120:3932-40. (b) Cigan M, Gaplovsky A, Sigmundova I, Zahradnik P, Dedic R, Hromadova M. Photostability of D-π-A nonlinear optical chromophores containing a benzothiazolium acceptor. J Phys Org Chem. 2011;24:450-9. (c) Coe BJ, Harris JA, Hall JJ, Brunschwig BS, Hung S, Libaers W, et al. Syntheses and Quadratic Nonlinear Optical Properties of Salts Containing Benzothiazolium Electron-Acceptor Groups. Chem Mater. 2006;18:590718. [13] (a) Bures F, Cvejn D, Melanova K, Benes L, Svoboda J, Zima V, Pytela O, Mikysek T, Ruzickova Z, Kityk IV, Wojciechowski A, AlZayed N. Effect of intercalation and chromophore arrangement on the linear and nonlinear optical properties of model aminopyridine push-pull molecules. J Mater Chem C. 2016;4:468-78. (b) Hao WH, Yan P, Li G, Wang ZY. Short-conjugated zwitterionic cyanopyridinium chromophores: Synthesis, crystal structure, and linear/nonlinear optical properties. Dyes Pigm. 2014;111:145-55. (c) Li L, Cui H, Yang Z, Tao X, Lin X, Ye N, et al. Synthesis and characterization of thienyl-substituted pyridinium salts for second-order nonlinear optics. CrystEngComm. 2012;14:1031-7. [14] (a) Lee SC, Kang BJ, Koo MJ, Lee SH, Han JH, Choi JY, et al. New Electro-Optic Salt Crystals for Efficient Terahertz Wave Generation by Direct Pumping at Ti:Sapphire Wavelength. Adv Opt Mater. 2017. 1600758. DOI: 10.1002/adom.201600758. (b) Lee S, Jazbinsek M, Hauri CP, Kwon OP. Recent progress in acentric core structures for highly efficient nonlinear optical crystals and their supramolecular interactions and terahertz applications. CrystEngComm. 2016;18:7180-203. (c) Lee S-H, Koo M-J, Lee K-H, Jazbinsek M, Kang B-J, Rotermund F, Kwon,OP. Quinolinium-based organic electro-optic crystals: Crystal characteristics in solvent mixtures and optical properties in the terahertz range. Mater Chem Phys. 2016;169:62-70. (d) Chen H, Ma Q, Zhou Y, Yang Z, Jazbinsek M, Bian Y, Ye N, Wang D, Cao H, He W. Engineering of Organic Chromophores with Large Second-Order Optical Nonlinearity and Superior Crystal Growth Ability. Cryst Growth Des. 2015;15:5560-7. http://www.rainbowphotonics.com [15] (a) Somma C, Folpini G, Gupta J, Reimann K, Woerner M, Elsaesser T. Ultra-broadband terahertz pulses generated in the organic crystal DSTMS. Opt Lett. 2015;40:3404-7. (b) Teng B, Wang S, Feng K, Cao L, Zhong D, You F, et al. Crystal growth, quality characterization and THz properties of DAST crystals. Cryst Res Technol. 2014;49:943-7. [16] (a) Buckley LER, Coe BJ, Rusanova D, Sanchez S, Jirasek M, Joshi VD, et al. Ferrocenyl helquats: unusual chiral organometallic nonlinear optical chromophores. Dalton Trans. 2017;46(4):1052-64. (b) Coe BJ, Rusanova D, Joshi VD, Sanchez S, Vavra J, Khobragade D, Severa L, Císařová I, Šaman D, Pohl R, Clays K, Depotter G, Brunschwig BS, Teplý F. Helquat Dyes: Helicene-like Push-Pull Systems with Large Second-Order Nonlinear Optical Responses. J Org Chem. 2016;81:1912-20. (c) Reyes-Gutierrez PE, Jirasek M, Severa L, Novotna P, Koval D, Sazelova P, et al. Functional helquats: helical cationic dyes with marked, switchable chiroptical properties in the visible region. Chem Commun. 2015;51:15836. (d) Coe BJ, Fielden J, Foxon SP, Asselberghs I, Clays K, Van Cleuvenbergen S, et al. Ferrocenyl Diquat Derivatives: Nonlinear Optical Activity, Multiple Redox States, and Unusual Reactivity. Organometallics. 2011;30(21):5731-43.

32

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

(e) Coe BJ, Fielden J, Foxon SP, Helliwell M, Brunschwig BS, Asselberghs I, et al. Quadratic and Cubic Nonlinear Optical Properties of Salts of Diquat-Based Chromophores with Diphenylamino Substituents. J Phys Chem A. 2010;114:12028-41. [17] Sucunza D, Cuadro AM, Alvarez-Builla J, Vaquero JJ. Recent Advances in the Synthesis of Azonia Aromatic Heterocycles. J Org Chem. 2016;81:10126-35. [18] (a) Marcelo G, Pinto S, Caneque T, Mariz IFA, Cuadro AM, Vaquero JJ, et al. Nonlinear Emission of Quinolizinium-Based Dyes with Application in Fluorescence Lifetime Imaging. J Phys Chem A. 2015;119:235162. (b) Caneque T, Cuadro AM, Alvarez-Builla J, Perez-Moreno J, Clays K, Castano O, et al. Novel charged NLO chromophores based on quinolizinium acceptor units. Dyes Pigm. 2014;101:116-21. (c) Macoas E, Marcelo G, Pinto S, Caneque T, Cuadro AM, Vaquero JJ, et al. A V-shaped cationic dye for nonlinear optical bioimaging. Chem Commun. 2011;47:7374-6. [19] The quinolizinium, as acharged moiety or as a quinolizinium ylide, is present in the structure of relevant natural alkaloids such as coralyne, the family of berberines (more than 60 alkaloids), sempervirine, flavopereirine, afrocurarine, neooxygambirtannine and flavocarpine, inter alia. For example, see: (a) Malik TA, Kamili AN, Chishti MZ, Tanveer S, Ahad S, Johri RK. In vivo anticoccidial activity of berberine [18, 5,6-dihydro-9,10dimethoxybenzo(g)-1,3-benzodioxolo(5,6-a) quinolizinium] - An isoquinoline alkaloid present in the root bark of Berberis lycium. Phytomedicine 2014;21:663-69. (b) Grycova L, Dostal J, Marek R. Quaternary protoberberine alkaloids. Phytochemistry, 2007;68:150-75. [20] (a) Zacharioudakis E, Caneque T, Custodio R, Muller S, Cuadro AM, Vaquero JJ, et al. Quinolizinium as a new fluorescent lysosomotropic probe. Bioorg Med Chem Lett. 2017;27:203-7. (b) Suarez RM, Bosch P, Sucunza D, Cuadro AM, Domingo A, Mendicuti F, et al. Targeting DNA with small molecules: a comparative study of a library of azonia aromatic chromophores. Org Biomol Chem. 2015;13:52738. (c)Yue X, Armijo Z, King K, Bondar MV, Morales AR, Frazer A,. Mikhailov I A , Przhonska OV, Belfield K. D. Steady-State and Femtosecond Transient Absorption Spectroscopy of New Two-Photon Absorbing FluoreneContaining Quinolizinium Cation Membrane Probes. ACS Appl Mater Interfaces. 2015;7:2833-46. (d) Granzhan A, Ihmels H, Tian M. The benzo[b]quinolizinium ion as a water-soluble platform for the fluorimetric detection of biologically relevant analytes. ARKIVOC. 2015(6Spec.Issue):494-523. (e) Faulhaber K, Granzhan A, Ihmels H, Otto D, Thomas L, Wells S. Studies of the fluorescence light-up effect of amino-substituted benzo[b]quinolizinium derivatives in the presence of biomacromolecules. Photochem Photobiol Sci. 2011;10:1535-45. [21] (a) Ramirez MA, Custodio R, Cuadro AM, Alvarez-Builla J, Clays K, Asselberghs I, Mendicuti F, Castaño O, Andrés JL, Vaquero JJ. Synthesis of charged bis-heteroaryl donor-acceptor (D-A+) NLO-phores coupling (πdeficient-π-excessive) heteroaromatic rings. Org Biomol Chem. 2013;11:7145-54. (b) Ramirez MA, Cuadro AM, Alvarez-Builla J, Castano O, Andres JL, Mendicuti F, Clays K, Asselbergh I, Vaquero JJ. Donor-(π-bridge)-azinium as D-π-A+ one-dimensional and D-π-A+-π-D multidimensional V-shaped chromophores. Org Biomol Chem. 2012;10:1659-69. (c) Ramirez MA, Caneque T, Cuadro AM, Mendicuti F, Clays K, Asselbergh I, Vaquero JJ. Novel linear and Vshaped D-π-A+-π-D chromophores by Sonogashira reaction. ARKIVOC. 2011:140-55. (d) Caneque T, Cuadro AM, Alvarez-Builla J, Perez-Moreno J, Clays K, Marcelo G, Mendicuti F, Castaño O, Andrés JL, Vaquero JJ. Heteroaromatic Cation-Based Chromophores: Synthesis and Nonlinear Optical Properties of Alkynylazinium Salts. Eur J Org Chem. 2010:6323,6330. [22] (a) Jeong JH, Kim J, Campo J, Lee SH, Jeon WY, Wenseleers W, Jazbinsek M, Yun H, Kwon OP. NMethylquinolinium derivatives for photonic applications: Enhancement of electron-withdrawing character beyond that of the widely-used N-methylpyridinium. Dyes Pigm. 2015;113:8-17. (b) Lee SH, Jazbinsek M, Yun H, Kim JT, Lee YS, Kwon OP. New quinolinium polymorph with optimal packing for maximal off-diagonal nonlinear optical response. Dyes Pigm. 2013;96:435-9. (c) Coe BJ, Beljonne D, Vogel H, Garin J, Orduna J. Theoretical Analyses of the Effects on the Linear and Quadratic Nonlinear Optical Properties of N-Arylation of Pyridinium Groups in Stilbazolium Dyes. J Phys Chem A. 2005;109:10052-7. [23] (a) Hrobarik P, Sigmundova I, Zahradnik P, Kasak P, Arion V, Franz E, et al. Molecular Engineering of Benzothiazolium Salts with Large Quadratic Hyperpolarizabilities: Can Auxiliary Electron-Withdrawing Groups Enhance Nonlinear Optical Responses?. J Phys Chem C. 2010;114:22289-302. (b) Quist F, Vande Velde CML, Didier D, Teshome A, Asselberghs I, Clays K, et al. Push-pull chromophores comprising benzothiazolium acceptor and thiophene auxiliary donor moieties: Synthesis, structure, linear and quadratic non-linear optical properties. Dyes Pigm. 2009;81:203-10. [24] (a) Farsadpour S, Taghizadeh Ghoochany L, Kaiser C, von Freymann G. New class of hyperpolarizable pushpull organic chromophores by applying a novel and convenient synthetic strategy. Dyes Pigm. 2016;127:73-7. (b) Ziemann EA, Freudenreich N, Speil N, Stein T, Van Steerteghem N, Clays K, Heck J. Synthesis, structure and NLO properties of a 1,3,5-substituted tricationic cobaltocenium benzene complex. J Organomet Chem. 2016;820:125-9. (c) Liu J, Gao W, Liu X, Zhen Z. Benefits of the use of auxiliary donors in the design and preparation of NLO chromophores. Mater Lett. 2015;143:333-5. (d) Zhang Y, Blau WJ. Nonlinear optics Dipoles align inside a nanotube. Nat Nanotechnol. 2015;10:205-6.

33

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

(e) Beverina L, Pagani GA. π-Conjugated Zwitterions as Paradigm of Donor-Acceptor Building Blocks in Organic-Based Materials. Acc Chem Res. 2014;47:319-29. (f) Zhou X, Luo J, Davies JA, Huang S, Jen AKY. Push-pull tetraene chromophores derived from dialkylaminophenyl, tetrahydroquinolinyl and julolidinyl moieties: optimization of second-order optical nonlinearity by fine-tuning the strength of electron-donating groups. J Mater Chem. 2012;22:16390-8. (g) Beverina L, Sanguineti A, Battagliarin G, Ruffo R, Roberto D, Righetto S, et al. UV absorbing zwitterionic pyridinium-tetrazolate: exceptional transparency/optical nonlinearity trade-off. Chem Commun. 2011;47:292-4. (h) Andreu R, Galan E, Orduna J, Villacampa B, Alicante R, Navarrete JTL, et al. Aromatic/Proaromatic Donors in 2-Dicyanomethylenethiazole Merocyanines: From Neutral to Strongly Zwitterionic Nonlinear Optical Chromophores. Chem - Eur J. 2011;17:826-38. (i) Abbotto A, Beverina L, Manfredi N, Pagani GA, Archetti G, Kuball H-G, Wittenburg C, Heck J, Holtmann J. Second-order nonlinear optical activity of dipolar chromophores based on pyrrole-hydrazono donor moieties. Chem - Eur J. 2009;15:6175-85. (j) Beaudin AMR, Song N, Bai Y, Men L, Gao JP, Wang ZY, et al. Synthesis and Properties of Zwitterionic Nonlinear Optical Chromophores with Large Hyperpolarizability for Poled Polymer Applications. Chem Mater 2006;18:1079-84. (k) Bai Y, Song N, Gao JP, Sun X, Wang X, Yu G, Wang ZY. A New Approach to Highly Electrooptically Active Materials Using Cross-Linkable, Hyperbranched Chromophore-Containing Oligomers as a Macromolecular Dopant. J Am Chem Soc. 2005;127:2060-1. (l) Song N, Men L, Gao JP, Bai Y, Beaudin AMR, Yu G, Wang ZY. Cross-Linkable Zwitterionic Polyimides with High Electrooptic Coefficients at Telecommunication Wavelengths. Chem Mater 2004;16:3708-13 (m) Szablewski M, Thomas PR, Thornton A, Bloor D, Cross GH, Cole JM, et al. Highly Dipolar, Optically Nonlinear Adducts of Tetracyano-p-quinodimethane: Synthesis, Physical Characterization, and Theoretical Aspects. J Am Chem Soc 1997;119:3144-54. [25] Garcia D, Cuadro AM, Alvarez-Builla J, Vaquero JJ. Sonogashira Reaction on Quinolizium Cations. Org Lett. 2004;6:4175-8. [26] (a) Garcia-Cuadrado D, Cuadro AM, Barchin BM, Nunez A, Caneque T, Alvarez-Builla J, et al. PalladiumMediated Functionalization of Heteroaromatic Cations: Comparative Study on Quinolizinium Cations. J Org Chem. 2006;71:7989-95. (b) Garcia-Cuadrado D, Cuadro AM, Alvarez-Builla J, Sancho U, Castano O, Vaquero JJ. First Synthesis of Biquinoliziniun Salts: Novel Example of a Chiral Azonia Dication. Org Lett. 2006;8:5955-8. (c) Garcia D, Cuadro AM, Alvarez-Builla J, Vaquero JJ. Sonogashira Reaction on Quinolizium Cations. Org Lett. 2004;6:4175-8. (d) Barchin BM, Valenciano J, Cuadro AM, Alvarez-Builla J, Vaquero JJ. Use of the Stille Coupling Reaction on Heteroaromatic Cations: Synthesis of Substituted Quinolizinium Salts. Org Lett. 1999;1:545-7. [27] Caneque T, Cuadro AM, Alvarez-Builla J, Vaquero JJ. Efficient functionalization of quinolizinium cations with organotrifluoroborates in water. Tetrahedron Lett. 2009;50:1419-22. [28] For leading references, see: (a) Molander GA, Ellis N. Organotrifluoroborates: Protected Boronic Acids That Expand the Versatility of the Suzuki Coupling Reaction. Acc Chem Res. 2007;40:275-86. (b)Molander GA, Figueroa R. Organotrifluoroborates: Expanding organoboron chemistry. Aldrichimica Acta. 2005;38:49-56. (c) Darses S, Genet J. Potassium Organotrifluoroborates: New Perspectives in Organic Synthesis. Chem Rev 2008;108:288-325. [29] For the synthesis of 1-bromo-, 3-bromo- and 4-bromoquinolizinium, see: (a) Sanders GM, Van Dijk M, Van der Plas HC. Synthesis and deuteration of some halogenoquinolizinium bromides. Heterocycles 1981;15(1):213-223. (b) For the synthesis of 2-bromoquinolizinium see: (b) Fozard A, Jones G. Quinolizines. V. The synthesis and properties of some hydroxyquinolizinium salts. J Chem Soc 1963:2203-2209. [30] (a) Miyadera T, Iwai I. The studies on quinolizinium salts. I. Synthesis of quinolizinium and 1-, 2-, 3-, and 4methylquinolizinium bromides. Chem Pharm Bull 1964;12:1338-43. (b) Sato K, Okazaki S, Yamagishi T, Arai S. The synthesis of azoniadithia[6]helicenes. J Heterocycl Chem. 2004;41:443-7. [31] Ishiyama T, Murata M, Miyaura N. Palladium(0)-Catalyzed Cross-Coupling Reaction of Alkoxydiboron with Haloarenes: A Direct Procedure for Arylboronic Esters. J Org Chem. 1995;60:7508-10. [32] Moylan CR, Twieg RJ, Lee VY, Swanson SA, Betterton KM, Miller RD. Nonlinear optical chromophores with large hyperpolarizabilities and enhanced thermal stabilities. J Am Chem Soc. 1993;115:12599-600. [33] (a) Jedrzejewska B, Pietrzak M, Paczkowski J. Solvent effect on the spectroscopic properties of styryl quinolinium dyes series. Dyes Pigm. 2013;97:278-85; (b) Li D, Yu D, Zhang Q, Li S, Zhou H, Wu J, TianY. Synthesis, crystal structure and third-order nonlinear optical properties in the near-IR range of a novel stilbazolium dye substituted with flexible polyether chains. Dyes Pigm. 2013;99:673-85; (c) Jędrzejewska B, Pietrzak M, Pączkowski J. Solvent Effects on the Spectroscopic Properties of Styrylquinolinium. Dyes Series. J. Fluoresc. 2010; 20:73-86. (d)Molecular Fluorescence: Principles and Applications; Valeur B. Wiley-VCH: Weinheim, 2002, Chapter 3, p.62 [34] Brouwer AM. Standards for photoluminescence quantum yield measurements in solution Pure Appl Chem. 2011;83:2213-28.

34

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

[35] D'Andrade BW, Datta S, Forrest SR, Djurovich P, Polikarpov E, Thompson ME. Relationship between the ionization and oxidation potentials of molecular organic semiconductors. Org Electron. 2005;6:11-20. [36] (a) Alain V, Thouin L, Blanchard-Desce M, Gubler U, Bosshard C, Gunter P, et al. Molecular engineering of push-pull phenylpolyenes for nonlinear optics. Improved solubility, stability, and nonlinearities. Adv Mater. 1999;11(14):1210-4. (b) Rao VP, Jen AKY, Wong KY, Drost KJ. Dramatically enhanced second-order nonlinear optical susceptibilities in (tricyanovinyl)thiophene derivatives. J Chem Soc , Chem Commun. 1993:1118-20. (c) Cheng LT, Tam W, Stevenson SH, Meredith GR, Rikken G, Marder SR. Experimental investigations of organic molecular nonlinear optical polarizabilities. 1. Methods and results on benzene and stilbene derivatives. J Phys Chem. 1991;95:10631-43. [37] Gaussian 09, Revision B.1. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, Bloino J,.Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Montgomery JA, Peralta JE Jr, Ogliaro F, Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD , Farkas O, Foresman JB, Ortiz JV, Cioslowski J, Fox DJ. Gaussian, Inc., Wallingford CT, 2014 For details of these calculations, see reference 20a,b. [38] Head-Gordon M, Rico RJ, Oumi M, Lee TJ. A doubles correction to electronic excited states from configuration interaction in the space of single substitutions. Chem Phys Lett. 1994;219:21-9. [39] Grimme S, Izgorodina EI. Calculation of 0-0 excitation energies of organic molecules by CIS(D) quantum chemical methods. Chem Phys. 2004;305:223-30. [40] (a)Johnson, L E.; Dalton, Larry R.; Robinson, Bruce H. Optimizing Calculations of Electronic Excitations and Relative Hyperpolarizabilities of Electrooptic Chromophores. Acc Chem Res. 2014;47:3258-65. (b)Castet F. Bogdan E. Plaquet A; Ducasse L, Champagne B, Rodriguez V. Reference molecules for nonlinear optics: a joint experimental and theoretical investigation. J Chem Phys. 2012;136: 024506/1-024506. [41] (a)Jacquemin D, Champagne B, Hattig C. Correlated frequency-dependent electronic first hyperpolarizability of small push-pull conjugated chains. Chem Phys Lett. 2000;319:327-34. (b)Sekino H, Bartlett RJ. Frequency-dependent hyperpolarizabilities in the coupled-cluster method: the Kerr effect for molecules. Chem Phys Lett. 1995;234:87-93. (c)Rice JE, Handy NC. The calculation of frequency-dependent hyperpolarizabilities including electron correlation effects. Int J Quantum Chem. 1992;43:91-118. [42] Wilberg N, Holeman AF, Wiberg E, Holleman-Wiberg E. Inorganic Chemistry. Eds. Walter de Gruyter GmbH & Co. Berlin, 2001;1757-59. [43] (a) Isborn CM, Leclercq A, Vila FD, Dalton LR, Bredas JL, Eichinger BE, Robinson, B. H. Comparison of Static First Hyperpolarizabilities Calculated with Various Quantum Mechanical Methods. J Phys Chem A. 2007;111:1319-27. (b) Carella A, Centore R, Fort A, Peluso A, Sirigu A, Tuzi A. Tuning second-order optical nonlinearities in push-pull benzimidazoles. Eur J Org Chem. 2004;12:2620-6. (c) Oudar JL, Chemla DS. Hyperpolarizabilities of the nitroanilines and their relations to the excited state dipole moment. J Chem Phys. 1977;66:2664-8. [44] (a) Wostyn K, Binnemans K, Clays K, Persoons A. Hyper-Rayleigh scattering in the Fourier domain for higher precision: Correcting for multiphoton fluorescence with demodulation and phase data. Rev Sci Instrum. 2001;72:3215-20. (b) Hendrickx E, Clays K, Persoons A. Hyper-Rayleigh Scattering in Isotropic Solution. Acc Chem Res. 1998;31:675-83. (c) Clays K, Persoons A. Hyper-Rayleigh scattering in solution. Phys Rev Lett. 1991;66:2980-3. [45] Olbrechts G, Clays K, Wostyn K, Persoons A. Fluorescence-free hyperpolarizability values by near-infrared, femtosecond hyper-Rayleigh scattering. Synth Met. 2000;115:207-11 [46] Franz E, Harper EC, Coe BJ, Zahradnik P, Clays K, Asselberghs I. Benzathiazoliums and pyridiniums for second-order nonlinear optics. Proc SPIE-Int Soc Opt Eng. 2008; 6999 (Organic Optoelectronics and Photonics III), 699923/1-699923/11. doi:10.1117/12.781194 [47] Zyss J, Dhenaut C, Chauvan T, Ledoux I. Quadratic nonlinear susceptibility of octupolar chiral ions. Chem Phys Lett. 1993;206:409-14. [48] Civin SJ, Rauch JE, Decius J. C. Theory of hyper-Raman effects (nonlinear inelastic light scattering). Selection rules and depolarization ratios for the second-order polarizability. J Chem Phys. 1965; 43:4083-95 [49] Farrugia LJ. WinGX suite for small-molecule single-crystal crystallography. J Appl Crystallogr. 1999; 32: 837-8. [50] Sheldrick GM. A short history of SHELX. Acta Crystallogr, Sect A: Found Crystallogr. 2008;64:112-22.

35