Next Article in Journal
Divergent Synthesis of Novel Five-Membered Heterocyclic Compounds by Base-Mediated Rearrangement of Acrylamides Derived from a Novel Isocyanide-Based Multicomponent Reaction
Previous Article in Journal
Facile Synthesis of Functionalized Spiropyrrolizidine Oxindoles via a Three-Component Tandem Cycloaddition Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Triazolobithiophene Light Absorbing Self-Assembled Monolayers: Synthesis and Mass Spectrometry Applications

by
Ghislain Tsague Kenfack
1,2,
Andreas Schinkovitz
1,
Suresh Babu
1,
Kamal Elouarzaki
2,
Marylène Dias
2,
Séverine Derbré
1,
Jean-Jacques Helesbeux
1,
Eric Levillain
2,
Pascal Richomme
1 and
Denis Séraphin
1,*
1
Laboratoire SONAS, Université d’Angers, IFR 149 QUASAV, 16 bd Daviers, 49100, Angers, France
2
Laboratoire MOLTECH-Anjou, Université d'Angers-CNRS, UMR 6200 du CNRS, 2 bd Lavoisier, 49045, Angers, France
*
Author to whom correspondence should be addressed.
Molecules 2011, 16(10), 8758-8774; https://doi.org/10.3390/molecules16108758
Submission received: 20 September 2011 / Revised: 10 October 2011 / Accepted: 17 October 2011 / Published: 19 October 2011
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
The synthesis of five light absorbing triazolobithiophenic thiols, which were utilized for producing self-assembled monolayers (SAMs) on gold surfaces, is presented. The monolayer formation was monitored by cyclic voltammetry, indicating excellent surface coverage. The new triazolobithiophenic compounds exhibited an absorption maximum around 340 nm, which is close to the emission wavelength of a standard nitrogen laser. Consequently these compounds could be used to aid ionization in laser desorption mass spectrometry (MS).

1. Introduction

Light absorbing self-assembled monolayers (SAMs) offer very interesting properties for various scientific and technical applications [1,2,3,4,5,6]. Some of the latter are photo-switch systems [5,7], light detectors [8], and light to current/energy convertors [1,9,10,11]. Furthermore, SAMs have been proposed to aid ionization in laser desorption MS. On the one hand, Mrksich combined self-assembled monolayers with matrix-assisted laser desorption/ionization mass spectrometry (SAMDI mass spectrometry) that permitted characterization of functionalized thiolates anchored to gold surfaces as a monolayer. Such a technique is therefore useful for characterizing chemical and biochemical interactions at surfaces [12,13,14]. On the other hand, SAM’s also help characterization by mass spectrometry of analytes that do not chemically interact with the monolayer [2,3,4].
With respect to the intended application, various light absorbing systems such as fullerenes [5], diphenyldiacetylene [7], and porphyrin derivatives [8,9,10,11] have been described. Among conjugated polymers, oligothiophenes have been extensively studied over the past decades due to interest in their applicability in the field of molecular electronics and linear or branched oligothiophene SAMs have been reported [15,16,17]. Despite these comprehensive studies, hardly anything is known about SAMs utilizing bithiophenes [18,19] as photo active groups [3]. 10-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)decane-1-thiol [1, Figure 1a] has been used to aid ionization in the recently developed “desorption/ionization on self-assembled monolayer surfaces” (DIAMS) method [3]. In continuation of this work, five triazolobitiophenic SAMs of general structure b [Figure 1b] have been produced. Exhibiting enhanced surface coverage and long-term stability, these SAMs further represent ideal candidates for advanced studies of the light to energy transfer during MALDI-TOF MS recordings. Indeed, earlier experiments have shown that the S-acetylated derivative of 1 [Figure 1a] exhibits an absorption maximum (λmax) of 340 nm, which is close to the emission wavelength of a standard nitrogen laser (337 nm) [3]. These specific absorption properties are primarily due to the bithiophenic moiety. With respect to the DIAMS application these properties were preserved in all the new molecules.
Figure 1. (a) 10-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)decane-1-thiol; (b) triazolo-5,5'-disubstituted-2,2’-bithiophene immobilized on a gold surface as SAMs.
Figure 1. (a) 10-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)decane-1-thiol; (b) triazolo-5,5'-disubstituted-2,2’-bithiophene immobilized on a gold surface as SAMs.
Molecules 16 08758 g001
Apart from these selective features, surface coverage together with long term stability represent crucial parameters for the quality of SAMs. The most basic bithiophenic SAM was produced by attaching a bithiophene moiety to a –(CH2)n–SH side chain of desired length, as previously described [3]. To increase coverage and stability, bithiophenic SAMs can be modified either at the terminal position or within the linker chain. It is well known that the functional group located on the surface of the monolayer, or in subterminal position significantly impacts packing arrangement. Indeed, van der Waals forces [20,21,22], dipole-dipole interactions [23,24,25], steric effect [26], hydrogen bonding [27,28,29,30,31,32,33,34,35] and π-π stacking [36] are all important factors for SAMs organization and stability. As changing the methyl group at the surface of SAMs requires a synthetic strategy with a complex desymmetrisation of the initial bithiophene, the authors decided to explore the effect of an extra aromatic ring in the subterminal part of the monolayer. In this context, the introduction of a 1,2,3-triazole ring appeared to be a good starting point, because the synthesis of the corresponding triazolobithiophene derivative could be easily achieved by a convergent click reaction. Further, the aromatic system (1,2,3-triazole ring) may exhibit π-π stacking [37,38] which could improve stability of the monolayers.
Preparations of self-assembled monolayers bearing a 1,2,3-triazole by applying a direct or a post-functionalization strategy have been previously reported [39,40], but this is the first report concerning triazolobithiophenic SAMs. The current manuscript presents the synthesis of thiols 2ac and 3a,b as well as their corresponding disulfide derivatives (Scheme 1). Subsequently free and immobilized molecules were studied by electrochemistry, and a potential application for this new group of light absorbing SAMs will be presented.
Scheme 1. Synthesis of triazolobithiophenes 2ac and 3a,b.
Scheme 1. Synthesis of triazolobithiophenes 2ac and 3a,b.
Molecules 16 08758 g004

2. Results and Discussion

The Huisgen reaction [41] consists of a 1,3-dipolar cycloaddition between a terminal alkyne and an organic azide, enabling the formation of mixture of 1,4 and 1,5-disubstituted triazole isomers [42]. Based on this approach, Sharpless et al. [43] developed the concept of click chemistry, which has become a highly recognized and frequently performed reaction in several fields of applied chemistry, such as biological, polymer and materials sciences, as well as drug discovery. Consequently the reaction appeared to be suitable for the design of bithiophenic SAMs. Considering that the location of the triazole moiety within the subterminal part of the monolayer might influence its stability [36,37], several different triazolobithiophenic derivatives were synthesized (Scheme 1). One group displayed either one, two or three carbons between the triazole ring and the bithiophene moiety (Scheme 1, triazolobithiophenes 4ac), while exhibiting a N–C connection between the triazole ring and the aliphatic side chain. Alternatively, a second group of triazolobithiophenes displaying a C–C connection between the triazole and the (CH2)8-SCOCH3 moiety were created (Scheme 1, triazolobithiophenes 5a,b). Alkynylbithiophenes 6ac and azidoalkylbithiophenes 7a,b were synthesized from 3-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)propanenitrile following a strategy proposed by Becher et al. [3,43] and underwent a copper (I) catalyzed Huisgen 1,3-dipolar cycloaddition under classical Sharpless conditions, yielding respectively high amounts of triazole linked bithiophenes 4ac and 5a,b (Scheme 1) [44].

2.1. Formation of Free Thiols and Disulfides

The self organization of SAMs on gold surfaces essentially requires the presence of free thiols or disulfides [45,46,47,48,49,50]. Consequently thioacetate derivatives 4a–c as well as 5a,b needed to be transformed into their corresponding thiols 2a–c and 3a,b or disulfides 11a–c and 12a,b. This was facilitated by mild alkaline hydrolysis as outlined in Scheme 1. Whether free thiols or disulfides were obtained primarily depended on the duration of the reaction. Short reaction times (≤2 min) selectively promoted the formation of thiols, while extended reaction times (≥90 min) led to gradual oxidation of these thiols and the formation of the corresponding disulfides.

2.2. Monolayer Formation and Properties

The bithiophene chromophore represents a core structure of the synthesized molecules and has been selected due to its specific absorption maxima. Precursors 4a–c and 5a,b and their follow up products displayed an absorption maximum of 340 nm (ξmax ≈ 21,000 L·mol−1·cm−1) [3,51], which is close to the emission wavelength of a standard nitrogen laser (337 nm). Additionally the bithiophene moiety exhibited useful electrochemical properties that permitted the use of cyclic voltammetry (CV) for studying the redox behavior of any synthesized compounds. It further allowed evaluation of the surface coverage and long term stability of immobilized 2a–c and 3a,b. As expected, CV of triazolobithiophenes 4a–c and 5a,b in solution exhibited a reversible one-electron oxidation wave around 0.49 V (vs. Fc+/Fc) [3]. This characteristic one electron-transfer process confirmed the redox properties of the bithiophenic chromophore.
The formation of SAMs was then performed using freshly prepared gold electrodes, which were incubated for 30 min at 20 °C in a CH2Cl2 solution (5·10−4 M) of thiols 2a–c, 3a,b or disulfides 11a–c and 12a,b. Figure 2 displays the cyclic voltammogram of immobilized 12b, showing a defined first reversible oxidation wave, which is characteristic for a surface confined redox system. In addition, the anodic-cathodic peaks separation was close to 0 volt, confirming that the redox process of the bithiophenic system was not limited by charge transport. Further, high surface coverage (Γ) (Table 1) was observed (about 3.8 × 10−10 mol·cm−2) indicating a good organization of SAMs [49]. Monolayers were stable for at least 48 h regardless whether they originated from disulfides (11b,c, 12a,b) or their corresponding thiols 2b–c, 3a,b. Only thiol 2a and its corresponding disulfide 11a did not produce stable monolayers. Furthermore, the stability of the triazolobithiophenic SAMs is enhanced by a spacer of at least two carbons between the triazole ring and the bithiophene moiety. This may allow a better packing arrangement of SAMs on the gold surface. In accordance to that, 2a and 11a did not produce stable monolayers.
Figure 2. Cyclic voltammogram of immobilized 12b (5 × 10−4 M in CH2Cl2).
Figure 2. Cyclic voltammogram of immobilized 12b (5 × 10−4 M in CH2Cl2).
Molecules 16 08758 g002
Table 1. Redox potentials (E1ox) a and covering rate (Γ) a for SAMs of immobilized 11a–c and 12a,b.
Table 1. Redox potentials (E1ox) a and covering rate (Γ) a for SAMs of immobilized 11a–c and 12a,b.
CompoundsE1oxΓ.10−10
11a0.52Not stable
11b0.453.8 ± 0.1
11c0.413.6 ± 0.1
12a0.503.8 ± 0.3
12b0.413.8 ± 0.2
a The data of potentials (E1ox) are displayed in volt versus Fc+/Fc in CH2Cl2; using Bu4NPF6 (0.1 M) as the supporting electrolyte on gold working electrode; scan rate 500 mV/s. The covering rate (Γ) was calculated in mol·cm−2.

2.3. MS Application

Compounds 2a–c and 3a,b have been specifically designed to study the principle of energy transmission in light absorbing monolayers. In this context, DIAMS has been described as a new laser desorption MS method [2,3,49]. Desorption/ionization on self-assembled monolayer surfaces aims for the elimination of matrix noise present in classical MALDI spectra, by replacing matrix molecules with light absorbing SAMs. Hithero, 1 represents the only light absorbing molecule reported for that purpose [2,3]. Benefitting from their improved physical properties, 2b,c and 3a,b as well as 11b,c and 12a,b are currently employed for thorough MS studies on their light absorption and energy transmission principle. Figure 2 shows the time of flight mass spectrum of the alkaloid yohimbine (C21H26N2O3). The compound was directly deposited onto freshly immobilized 3b and ionized by a laser beam at 337 nm, a wavelength at which yohimbine does not absorb light [50]. Only the pseudo-molecular ion [M−H]+ at m/z 353.21, but no fragmentation products or signals from the monolayer were detected. The stability of the SAMs was also confirmed by recording a spectrum without any analyte at the same laser attenuation (70%). As previously observed for bithiophenic SAM 1 [2], at this level of laser energy, ions characteristic of the monolayer degradation are not detected.
Figure 3. Mass spectrum of yohimbine acquired by laser desorption ionization from immobilized 3b in linear positive mode (C21H26N2O3 calculated isotopic mass: 354.19, observed: [M−H]+: 353.21. Laser energy 30% (46.8 µJ)). On first sight the pseudomolecular ion [M−H]+ appears unusual, nevertheless its occurrence is essentially linked to the process of ion formation in laser desorption MS with photochemical interactions between matrix and analyte molecules [54].
Figure 3. Mass spectrum of yohimbine acquired by laser desorption ionization from immobilized 3b in linear positive mode (C21H26N2O3 calculated isotopic mass: 354.19, observed: [M−H]+: 353.21. Laser energy 30% (46.8 µJ)). On first sight the pseudomolecular ion [M−H]+ appears unusual, nevertheless its occurrence is essentially linked to the process of ion formation in laser desorption MS with photochemical interactions between matrix and analyte molecules [54].
Molecules 16 08758 g003

3. Experimental

3.1. General

All reagents and solvents were of commercial quality. DMF and MeOH were dried on molecular sieves (4Å). Column chromatography (CC) was performed on silica gel 60 (Merck, particle diameter 0.035–0.070 mm). Analytical TLC was carried out on standard silica gel 60 plates (Merck silica gel 60 F254) and compounds were visualized under UV light (254/366 nm). In addition plates were sprayed with an ethanolic solution of sulfuric acid (10%) and heated for 1 min at 100 °C. Melting points were determined using a WME-type Köffler hot stage (Wagner&Munz). 1H- and 13C-NMR spectra were obtained at room temperature on a Bruker Avance III 300 or a Bruker Avance DRX 500 spectrometer. Infrared spectra were recorded in the range of 4,000–400 cm−1 using potassium bromide disks or ATR, with a Bruker FT IR Vector 22 spectrophotometer. UV-Vis absorption was recorded on a Perkin-Elmer Lambda 19 and on a Perkin-Elmer Lambda 950 spectrometer. Exact mass measurements were performed on a LTQ-Orbitrap MS (Thermo Fisher Scientific) or on a Bruker MicrOTOF-Q II MS. The DIAMS mass spectrum of the alkaloid yohimbine was recorded on a Bruker Biflex III time-of-flight mass spectrometer (Bruker-Daltonics) equipped with a nitrogen laser (337 nm, model VSL-337i, Laser Science Inc.).

3.2. Electrochemistry

Electrochemical experiments were carried out with a Biologic SP-150 potentiostat at 20 °C. Cyclic voltammetry was performed in a three-electrode cell equipped with a platinum-plate counter electrode and an Ag/AgNO3 (0.01 M CH3CN) reference electrodes. Cyclic voltammograms were recorded in HPLC-grade methylene chloride (CH2Cl2) using Bu4NPF6 as supporting electrolyte. Gold electrodes were prepared by physical vapor deposition as previously described. In short, borosilicate glass plates were covered with a chrome layer of about 5 nm, followed by a gold layer of approximately 50 nm. Fresh electrodes were prepared just before the experiments.

3.3. Synthesis

3.3.1. Alkynylbithiophenes 6a–c

3-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)propionitrile (8, 150 mg, 0.51 mmol, 1 equiv.) was dissolved in dry DMF (13 mL) and the solution was sparged with nitrogen for 45 min. Then an oxygen free methanolic solution (7 mL) of cesium hydroxide (128 mg, 0.76 mmol, 1.5 equiv.) was added and the mixture was stirred at room temperature for 1 h. Finally, 3-bromoprop-1-yne (121 mg, 1.02 mmol, 2 equiv.) was added. The solution was chilled to a temperature of −40 °C and stirred for another hour. The mixture was diluted with water (40 mL) and extracted with AcOEt (3 × 40 mL). The combined organic phases were washed with water (3 × 50 mL) and dried over MgSO4. After solvent removal, the residue was purified by isocratic CC (petroleum ether/CH2Cl2: 80/20) yielding 5-(methylthio)-5'-(prop-2-ynylthio)-2,2'-bithiophene compound (6a). Yield 86 %; green brownish oil; C12H10S4; IR (ATR): 791, 869, 964, 1070, 1196, 1225, 1313, 1420, 1497, 2918, 3066, 3288 cm−1; 1H-NMR (CDCl3/300 MHz): δ 2.31 (t, J = 2.7 Hz, 1H), 2.51 (s, 3H), 3.49 (d, J = 2.4 Hz, 2H), 6.96–7.17 (m, 4H); 13C-NMR (CDCl3/75 MHz): δ 21.9, 27.4, 72.7, 79.5, 123.8, 124.2, 131.3, 131.5, 136.1, 137.2, 138.7, 141.8. MALDI-TOF-MS positive mode [C12H10S4]+: 281.9 m/z.
Compounds 6b and 6c were prepared following procedure of 6a with slight modifications. For compound 6b, 4-bromobut-1-yne was added to the cesium hydroxide solution at 0 °C. The mixture was stirred overnight at room temperature. Compound 6c was obtained under the same conditions using 4 equiv. of 5-chloropent-1-yne. The mixture was stirred at room temperature for 24 h.
5-(But-3-ynylthio)-5'-(methylthio)-2,2'-bithiophene (6b). Yield 94%; green yellowish solid; C13H12S4; mp: 68–70 °C; IR (ATR): 791, 865, 966, 1196, 1220, 1275, 1312, 1422, 1497, 1688, 2853, 2921, 3278 cm-1; 1H-NMR (CDCl3/500 MHz): δ 2.04 (t, J = 2.5 Hz, 1H), 2.50 (dt, J = 7.0 and 2.5 Hz, 2H), 2.51 (s, 3H), 2.92 (t, J = 7.0 Hz, 2H), 6.96–7.06 (m, 4H); 13C-NMR (CDCl3/125 MHz): δ 19.4, 22.0, 37.2, 69.7, 81.9, 123.7, 124.0, 131.6, 132.1, 135.5, 137.0, 138.7, 141.1. MALDI-TOF-MS positive mode [C13H12S4]+: 295.9 m/z.
5-(Methylthio)-5'-(pent-4-ynylthio)-2,2'-bithiophene (6c). Yield 85%; green brownish oil; C14H14S4; IR (ATR): 792, 868, 964, 1196, 1212, 1343, 1423, 1497, 2852, 2918, 3067, 3294 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.85 (qu, J = 7.0 Hz, 2H), 1.97 (t, J = 2.5Hz, 1H), 2.35 (dt, J = 7.0 and 2.5 Hz, 2H), 2.51 (s, 3H), 2.91 (t, J = 7 Hz, 2H), 6.96–7.01 (m, 4H); 13C-NMR (CDCl3/125 MHz): δ 17.1, 22.0, 27.9, 37.5, 69.1, 83.1, 123.7, 123.9, 131.6, 133.4, 134.5, 136.9, 138.9, 140.4. MALDI-TOF-MS positive mode [C14H14S4]+: 310.0 m/z.

3.3.2. Azidobithiophenes 7a–b

Azidobithiophenes 7a and 7b were prepared with slight modifications according to the general procedure described for alkynylbithiophenes 6a–c. Alternatively, 2-azidoethyl-4-methylbenzenesulfonate (for 7a) or 3-azidopropyl-4-methylbenzenesulfonate (for 7b) was added to the cesium hyroxyde solution of 8. The mixtures were kept at room temperature and stirred for 4 h (7a) or 24 h (7b). Finally, compounds 7a and 7b were purified by isocratic CC (cyclohexane/CH2Cl2: 90/10).
5-(2-Azidoethylthio)-5'-(methylthio)-2,2'-bithiophene (7a). Yield 99%; green solid; C11H11N3S4; mp: 44–46 °C; IR (ATR): 794, 868, 964, 1224, 1256, 1423, 1498, 1687, 2099, 2855, 2921, 3067 cm−1; 1H-NMR (CDCl3/270 MHz): δ 2.50 (s, 3H), 2.94 (t, J = 6.9 Hz, 2H), 3.47 (t, J = 6.9 Hz, 2H), 6.97–7.26 (m, 4H); 13C-NMR (CDCl3/67.5 MHz): δ 21.9, 37.7, 50.0, 123.8, 124.2, 131.0, 131.5, 135.8, 137.2, 138.5, 141.4. MALDI-TOF-MS positive mode [C11H11N3S4]+: 312.9 m/z.
5-(3-Azidopropylthio)-5'-(methylthio)-2,2'-bithiophene (7b). Yield 76%; green oil; C12H13N3S4; IR (ATR): 792, 867, 964, 1069, 1245, 1421, 1498, 1687, 2092, 2859, 2920, 3068 cm−1; 1H-NMR (CDCl3/270 MHz): δ 1.89 (qu, J = 7.0 Hz, 2H), 2.50 (s, 3H), 2.87 (t, J = 7.0 Hz, 2H), 3.44 (t, J = 7.0 Hz, 2H), 6.97–7.01 (m, 4H); 13C-NMR (CDCl3/67.5 MHz): δ 21.9, 28.4, 35.6, 49.5, 123.7, 123.9, 131.5, 132.7, 134.8, 136.9, 138.9, 140.6. MALDI-TOF-MS positive mode [C12H13N3S4]+: 327.0 m/z.

3.3.3. 8-Azidooctyl Thioacetate (9)

8-Bromooctyl thioacetate (2.75 g, 10.3 mmol, 1 equiv.) was dissolved in dry DMF (100 mL). After the addition of sodium azide (1 g, 15.3 mmol, 1.5 equiv.), the mixture was heated to 110 °C overnight. After cooling to room temperature, distilled water (100 mL) was added. The mixture was extracted with diethyl ether (2 × 150 mL), and the combine organic layers were washed with distilled water (2 × 150 mL) and dried over MgSO4. After solvent evaporation, the residue was purified by isocratic CC (AcOEt/petroleum ether: 5/95). Yield 89%; yellowish liquid; C10H19N3OS; IR (ATR): 954, 1133, 1260, 1352, 1460, 1690, 2092, 2856, 2930 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.19–1.37 (m, 8H), 1.53–1.62 (m, 4H), 2.32 (s, 3H), 2.86 (t, J = 7.0 Hz, 2H), 3.25 (t, J = 7.0 Hz, 2H); 13C-NMR (CDCl3/125 MHz): δ 26.6, 28.6, 28.7, 28.9, 28.9, 29.0, 29.4, 30.6, 51.4, 195.9. HRMS (ESI+): calcd for [C10H19N3OS+Na]+: 252.1141, found 252.1143.

3.3.4. Dec-9-ynyl Thioacetate (10)

Dec-9-ynyl methanesulfonate (0.2 g, 0.86 mmol, 1 equiv.) was dissolved in dry DMF (10 mL). After the addition of 20 mL oxygen free potassium thioacetate solution (1.1 equiv., 0.95 mmol) in dry DMF, the mixture underwent microwave irradiation (400 W, 100 °C, 30 min). After solvent evaporation, the residue was purified by isocratic CC (cyclohexane/AcOEt: 90/10). Yield 50%; orange brownish liquid; C12H20OS; IR (ATR): 954, 1133, 1353, 1433, 1460, 1690, 2856, 2929, 3301 cm−1; 1H-NMR (CDCl3/270 MHz): δ 1.20–1.37 (m, 8H), 1.42–1.53 (m, 4H), 1.89 (t, J = 2.7 Hz, 1H), 2.09–2.15 (m, 2H), 2.27 (s, 3H), 2.81 (t, J = 7.3 Hz, 2H); 13C-NMR (CDCl3/67.5 MHz): δ 17.8, 27.8, 28.0, 28.1, 28.3, 28.4, 28.5, 28.9, 30.0, 67.6, 84.0, 195.3. MALDI-TOF-MS negative mode [C12H20OS-H]: 211.0 m/z.

3.3.5. Triazolobithiophenes 4a–c

An aqueous solution (2.5 mL) containing CuSO4.5H2O (12 mg, 0.05 mmol, 0.12 equiv.) and sodium ascorbate (29 mg, 0.15 mmol, 0.36 equiv.) were added to a mixture of alkynylbithiophene 6a (125 mg, 0.44 mmol, 1 equiv.) and azidothioacetate 9 (107 mg, 0.47 mmol, 1.06 equiv.) in tBuOH/H2O (4 mL/1.5 mL). The mixture was stirred overnight at room temperature yielding a green brownish solid. The latter was first filtered and then washed with distilled water (4 × 25 mL) and diethyl ether (10 mL). After being vacuum dried at 50 °C for several hours, the residue was identified as triazolobithiophene 4a. Triazolobithiophenes 4b and 4c were obtained following the same procedure, but starting respectively from alkynylbithiophenes 6b and 6c.
S-8-(4-((5'-(Methylthio)-2,2'-bithiophen-5-ylthio)methyl)-1H-1,2,3-triazol-1-yl)octyl thioacetate (4a). Yield 83%; pale green brownish solid; C22H29ON3S5; mp: 85–87 °C; UV-Vis (CH2Cl2): 345 nm; IR (ATR): 791, 867, 961, 1056, 1142, 1352, 1422, 1464, 1691, 2851, 2920, 3069, 3124 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.28 (s, 8H), 1.53–1.59 (m, 2H), 1.81–187(m, 2H), 2.32 (s, 3H), 2.51 (s, 3H), 2.85 (t, J = 7.0 Hz, 2H), 4.10 (s, 2H), 4.29 (t, J = 7.0 Hz, 2H), 6.92–6.95 (m, 4H), 7.28(s, 1H); 13C-NMR (CDCl3/125 MHz): δ 21.9, 26.3, 28.6, 28.7, 28.8, 29.0, 29.4, 30.3, 30.6, 33.7, 50.3, 121.9, 123.6, 124.1, 131.5, 132.3, 135.4, 137.1, 138.7, 141.1, 196.0. HRMS (ESI+): calcd for [C22H29ON3S5+H]+: 512.0987, found 512.0984.
S-8-(4-(2-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)ethyl)-1H-1,2,3-triazol-1-yl)octyl thioacetate (4b). Yield 95%; pale green brownish solid; C23H31ON3S5; mp: 85–87 °C; UV-Vis (CH2Cl2): 344 nm; IR (ATR): 816, 864, 960, 1055, 1143, 1213, 1269, 1426, 1464, 1694, 2359, 2851, 2922, 3148 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.30 (s, 8H), 1.54 (qu, J = 7.0 Hz, 2H), 1.86–1.89 (m, 2H), 2.31 (s, 3H), 2.51 (s, 3H), 2.84 (t, J = 7.0 Hz, 2H), 3.05 (t, J = 7.0 Hz, 2H), 3.12 (t, J = 7.0 Hz, 2H), 4.31(t, J = 7.0 Hz, 2H), 6.96–7.03 (m, 4H), 7.36(s, 1H); 13C-NMR (CDCl3/125 MHz): δ 22.0, 25.9, 26.4, 28.6, 28.8, 28.9, 29.0, 29.4, 30.2, 30.6, 38.1, 50.2, 121.2, 123.8, 123.9, 131.6, 132.9, 134.8, 136.9,138.8, 140.6, 145.5, 196.0. HRMS (ESI+): calcd for [C23H31ON3S5+H]+: 526.1143, found 526.1145.
S-8-(4-(3-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)propyl)-1H-1,2,3-triazol-1-yl)octyl thioacetate (4c). Yield 70%; pale green solid; C24H33ON3S5; mp: 84–86 °C; UV-Vis (CH2Cl2): 344 nm; IR (ATR): 792, 862, 960, 1059, 1143, 1207, 1351, 1425, 1461, 1691, 2850, 2921, 3069, 3120 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.29 (s, 8H), 1.52–1.58 (m, 2H), 1.84–1.88 (m, 2H), 2.03 (qu, J = 7.0 Hz, 2H), 2.32 (s, 3H), 2.51 (s, 3H), 2.83-2.87 (m, 6H), 4.29 (t, J = 7.0 Hz, 2H), 6.96–7.01 (m, 4H); 7.25 (s, 1H); 13C-NMR (CDCl3/125 MHz): 22.0, 24.1, 26.4, 28.6, 28.7, 28.8, 28.9, 29.0, 29.39, 30.2, 30.6, 38.0, 50.2, 120.7, 123.7, 123.9, 131.6, 133.6, 134.4, 136.8, 138.8, 140.2, 195.9. HRMS (ESI+): calcd for [C24H33ON3S5+H]+: 540.1299, found 540.1296.

3.3.6. Triazolobithiophenes 5a–b

Triazolobithiophenes 5a and 5b were prepared following the procedure previously described for triazolobithiophenes 4a–c, but using azidobithiophenes 7a,b and dec-9-ynyl thioacetate 10 instead of 6a–c and 9. Moreover, the reaction time was extended to 36 h.
S-8-(1-(2-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)ethyl)-1H-1,2,3-triazol-4-yl)octyl thioacetate (5a). Yield 91%; pale green brownish solid; C23H31ON3S5; mp: 80–82 °C; UV-Vis (CH2Cl2): 343 nm; IR (KBr): 785, 866, 956, 1053, 1136, 1218, 1353, 1689, 2849, 2923, 3062, 3114 cm−1; 1H-NMR (CDCl3/270 MHz): δ 1.30 (s, 8H), 1.54 (t, J = 6.8 Hz 2H), 1.65 (m, 2H), 2.31 (s, 3H), 2.51 (s, 3H), 2.70 (t, J = 7.6 Hz, 2H), 2.84 (t, J = 7.2 Hz, 2H), 3.24 (t, J = 6.6 Hz, 2H), 4.53 (t, J = 6.6 Hz, 2H), 6.97–7.06 (m, 4H), 7.33 (s, 1H); 13C-NMR (CDCl3/67.5 MHz): δ 21.9, 25.6, 28.7, 28.9, 29.1, 29.3, 29.4, 30.6, 38.2, 49.0, 64.0, 71.4, 121.2, 123.8, 124.3, 130.8, 131.4, 135.8, 137.4, 138.2, 141.5, 148.3, 196.0. HRMS (ESI+): calcd for [C23H31ON3S5+H]+: 526.1143, found 526.1141.
S-8-(1-(3-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)propyl)-1H-1,2,3-triazol-4-yl)octyl thioacetate (5b). Yield 81%; pale green brownish solid; C24H33ON3S5; mp: 80–82 °C; UV-Vis (CH2Cl2): 343 nm; IR (KBr): 785, 865, 955, 1052, 1125, 1212, 1354, 1467, 1689, 2850, 2920, 3064, 3116 cm−1; 1H-NMR (CDCl3/270 MHz): δ 1.31 (s, 8H), 1.55 (t, J = 4.0Hz, 2H), 1.65 (m, 4H), 2.24 (m, 2H), 2.32 (s, 3H), 2.51 (s, 3H), 2.75–2.82 (m, 2H), 2.86 (t, J = 3.6 Hz, 2H), 4.51 (s, 2H), 6.97–7.03 (m, 4H), 7.33 (s, 1H); 13C-NMR (CDCl3/67.5 MHz): δ 21.9, 25.6, 28.7, 28.9, 29.0, 29.1, 29.3, 29.4, 29.5, 30.6, 35.3, 38.3, 48.1, 64.0, 120.9, 123.8, 124.1, 131.5, 132.3, 135.0, 137.2, 138.6, 140.9, 195.9. HRMS (ESI+): calcd for [C24H33ON3S5+H]+: 540.1299, found 540.1295.

3.3.7. Thiols 2a–c and 3a,b

Triazole 4c (20 mg, 37 mmol, 1 equiv.) was dissolved in dry DMF (2 mL) ventilated with nitrogen for 45 min. Then, 1 mL of a methanolic cesium hydroxide solution (22 mg, 0.13 mmol, 3.5 equiv.) was added. The mixture was stirred for 2 min at room temperature. The reaction was quenched by addition of 0.2% HCl solution (5 mL). The solid was filtered and washed with distilled water (5 mL). In a second purification step, the residue was solubilized in CH2Cl2 (20 mL), and washed with distilled water (4 × 20 mL). Finally the organic phase was dried using MgSO4. After solvent removal, 2c was obtained.
Following the same procedure, the hydrolysis of 4a,b and 5a,b yielded respectively 2a,b and 3a,b.
8-(4-((5'-(Methylthio)-2,2'-bithiophen-5-ylthio)methyl)-1H-1,2,3-triazol-1-yl)octane-1-thiol (2a). Yield 55%; pale yellow solid; C20H27N3S5; mp: 80–82 °C; IR (ATR): 789, 863, 1055, 1141, 1227, 1425, 1459, 1600, 2853, 2919, 3073, 3124 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.25–1.37 (m, 8H), 1.53–1.63 (m, 2H), 1.81–188 (m, 2H), 2.50 (s, 3H), 2.51 (pseudo q, J = 7.5 Hz, 2H), 4.10 (s, 2H), 4.29 (t, J = 7.2 Hz, 2H), 6.90–6.95 (m, 4H), 7.26 (s, 1H); 13C-NMR (CDCl3/75 MHz): δ 21.9, 26.3, 28.3, 28.8, 28.9, 29.1, 30.3, 33.7, 38.9, 50.3, 121.9, 123.6, 124.1, 131.5, 132.2, 135.4, 137.1, 138.6, 141.1, 144.1. HRMS (ESI): calcd for [C20H27N3S5-H]: 468.0736, found 468.0736.
8-(4-(2-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)ethyl)-1H-1,2,3-triazol-1-yl)octane-1-thiol (2b). Yield 44%; brown yellowish solid; C21H29N3S5; mp: 79–81 °C; IR (ATR): 796, 865, 1035, 1099, 1211, 1257, 1314, 1427, 1632, 2850, 2919, 3071, 3143 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.25–1.34 (m, 8H), 1.54–1.64 (m, 2H), 1.85–1.90 (m, 2H), 2.51 (pseudo q, J = 7.5 Hz, 2H), 2.51 (s, 3H), 3.01–3.15 (m, 4H), 4.31 (t, J = 7.2 Hz, 2H), 6.95–7.03 (m, 4H), 7.35 (s, 1H); 13C-NMR (CDCl3/75 MHz): δ 22.0, 25.9, 26.4, 28.3, 28.8, 28.9, 29.0, 30.3, 38.1, 38.9, 50.2, 121.2, 123.8, 123.9, 131.6, 132.9, 134.8, 136.9, 138.8, 140.6, 145.5. HRMS (ESI-): calcd for [C21H29N3S5-H]: 482.0892, found 482.0892.
8-(4-(3-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)propyl)-1H-1,2,3-triazol-1-yl)octane-1-thiol (2c). Yield 41%; brown yellowish solid; C22H31N3S5; mp: 76–78 °C; IR (KBr): 792, 861, 993, 1059, 1157, 1205, 1424, 1459, 1552, 2849, 2921, 3067, 3119 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.24–1.35 (m, 8H), 1.54-1.63 (m, 2H), 1.84–1.88 (m, 2H), 2.02 (qu, J = 7.2 Hz, 2H), 2.50 (s, 3H), 2.51 (pseudo q, J = 7.5 Hz, 2H), 2.82-2.87 (m, 4H), 4.28 (t, J = 7.2 Hz, 2H), 6.96–7.01 (m, 4H), 7.23 (s, 1H); 13C-NMR (CDCl3/75 MHz): δ 22.0, 24.1, 26.4, 28.3, 28.8, 28.9, 29.0, 29.1, 30.3, 38.0, 38.9, 50.2, 120.7, 123.7, 123.9, 131.6, 133.7, 134.4, 136.8, 138.9, 140.2, 146.8. HRMS (ESI): calcd for [C22H31N3S5-H]: 496.1049, found 496.1048.
8-(1-(2-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)ethyl)-1H-1,2,3-triazol-4-yl)octane-1-thiol (3a). Yield 44%; brownish solid; C21H29N3S5; mp: 76–78 °C; IR (ATR): 790, 865, 988, 1058, 1105, 1221, 1316, 1426, 1550, 2852, 2922, 3066, 3113 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.24–1.34 (m, 8H), 1.57–1.68 (m, 4H), 2.51 (pseudo q, J = 7.5 Hz, 2H), 2.51 (s, 3H), 2.64–2.72 (m, 2H), 3.24 (t, J = 6.9 Hz, 2H), 4.53 (t, J = 6.9 Hz, 2H), 6.96–7.07 (m, 4H), 7.32 (s, 1H); 13C-NMR (CDCl3/75 MHz): 21.9, 25.6, 28.4, 29.1, 29.3, 29.4, 29.7, 32.7, 38.3, 39.1, 49.1, 121.2, 123.9, 124.3, 130.8, 131.5, 135.9, 137.5, 138.3, 141.5, 148.3. HRMS (ESI-): calcd for [C21H29N3S5-H]: 482.0892 , found 482.0892.
8-(1-(3-(5'-(Methylthio)-2,2'-bithiophen-5-ylthio)propyl)-1H-1,2,3-triazol-4-yl)octane-1-thiol (3b). Yield 43%; pale yellow solid; C22H31N3S5; mp: 74–76 °C; IR (ATR): 792, 860, 907, 1037, 1170, 1211, 1425, 1495, 1548, 2850, 2921, 3066, 3116 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.30–1.35 (m, 8H), 1.57–1.64 (m, 4H), 2.17–2.24 (m, 2H), 2.51 (s, 3H), 2.51 (pseudo q, J = 7.5 Hz, 2H), 2.64–2.71 (m, 2H), 2.76 (t, J = 6.9 Hz, 2H), 4.46 (t, J = 6.9 Hz, 2H), 6.97–7.03 (m, 4H), 7.23 (s, 1H); 13C-NMR (CDCl3/75 MHz): δ 18.5, 22.1, 25.7, 28.5, 29.2, 29.3, 29.5, 29.6, 35.4, 39.2, 48.2, 58.6, 120.9, 123.9, 124.2, 131.6, 132.3, 135.2, 137.3, 138.6, 141.0, 148.6. HRMS (ESI-): calcd for [C22H31N3S5-H]: 496.1049, found 496.1050.

3.3.8. Disulfides 11a–c, 12a,b

The reaction was carried out described for thiols 2a–c, and 3a,b under slightly modified conditions. The reaction time was increased to 90 min and 3 equiv. cesium hydroxide (19 mg, 0.11 mmol) were used. These conditions finally yielded disulfides 11a–c and 12a,b, respectively.
1,2-Bis(8-(4-((5'-(methylthio)-2,2'-bithiophen-5-ylthio)methyl)-1H-1,2,3-triazol-1-yl)octyl)disulfide (11a). Yield 44%; pale yellow solid; C40H52N6S10; mp: 122–124 °C; IR (KBr): 791, 867, 1057, 1226, 1312, 1342, 1423, 1462, 1497, 1549, 2849, 2918, 3069, 3122 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.29 (s, 8H), 1.58–1.65 (m, 2H), 1.81–186 (m, 2H), 2.50 (s, 3H), 2.66 (t, J = 6.9 Hz, 2H), 4.10 (s, 2H), 4.29 (t, J = 6.9 Hz, 2H), 6.92–6.95 (m, 4H), 7.26 (s, 1H); 13C-NMR (CDCl3/75 MHz): δ 21.9, 26.3, 28.3, 28.8, 28.9, 29.1, 30.3, 33.7, 38.9, 50.3, 121.9, 123.6, 124.1, 131.5, 132.2, 135.4, 137.1, 138.6, 141.1, 144.1. HRMS (ESI+): calcd for [C40H52N6S10+H]+: 937.1540, found 937.1544.
1,2-Bis(8-(4-(2-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)ethyl)-1H-1,2,3-triazol-1-yl)octyl)disulfide (11b). Yield 50%; brown yellowish solid; C42H56N6S10; mp: 118–120 °C; IR (KBr): 792, 865, 1055, 1144, 1214, 1268, 1316, 1427, 1498, 1551, 2850, 2920, 3067, 3146 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.25–1.36 (m, 8H), 1.57–1.67 (m, 2H), 1.87–1.90 (m, 2H), 2.51 (s, 3H), 2.65 (t, J = 7.5 Hz, 2H), 3.06 (t, J = 7 Hz, 2H), 3.13 (t, J = 6.5 Hz, 2H), 4.32 (t, J = 7.5 Hz, 2H), 6.96–7.03 (m, 4H), 7.38 (s, 1H); 13C-NMR (CDCl3/125 MHz): δ 22.0, 25.9, 26.4, 28.3, 28.8, 28.9, 29.0, 30.3, 38.1, 38.9, 50.2, 121.2, 123.8, 123.9, 131.6, 132.9, 134.8, 136.9, 138.8, 140.6, 145.5 ppm. HRMS (ESI+) calcd for [C42H56N6S10+H]+: 965.1846, found 965.1855.
1,2-Bis(8-(4-(3-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)propyl)-1H-1,2,3-triazol-1-yl)octyl)disulfide (11c). Yield 40%; brown yellowish solid; C44H60N6S10; mp: 118–120 °C; IR (KBr): 790, 865, 1026, 1059, 1150, 1211, 1424, 1460, 1497, 1550, 2849, 2920, 3068, 3123 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.31 (s, 8H), 1.58–1.68 (m, 2H), 1.84–1.89 (m, 2H), 2.02 (qu, J = 7.2 Hz, 2H), 2.50 (s, 3H), 2.66 (t, J = 7.5 Hz, 2H), 2.82–2.88 (m, 4H), 4.28 (t, J = 7.2 Hz, 2H), 6.96–7.01 (m, 4H), 7.23 (s, 1H); 13C-NMR (CDCl3/75 MHz): δ 22.0, 24.1, 26.4, 28.3, 28.8, 28.9, 29.0, 29.1, 30.3, 38.0, 38.9, 50.2, 120.7, 123.7, 123.9, 131.6, 133.7, 134.4, 136.8, 138.9, 140.2, 146.8. HRMS (ESI+) calcd for [C44H60N6S10+H]+: 993.2160, found 993.2161.
1,2-Bis(8-(1-(2-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)ethyl)-1H-1,2,3-triazol-4-yl)octyl)disulfide (12a). Yield 66%; brown yellowish solid; C42H56N6S10; mp: 125–127 °C; IR (KBr): 794, 865, 1052, 1096, 1212, 1261, 1423, 1464, 1496, 1548, 2850, 2921, 3062, 3119 cm−1; 1H-NMR (CDCl3/300 MHz): δ 1.25–1.35 (m, 8H), δ 1.61–1.71 (m, 4H), 2.52 (s, 3H), 2.64–2.72 (m, 4H), 3.24 (t, J = 6.9 Hz, 2H), 4.53 (t, J = 6.9 Hz, 2H), 6.96–7.06 (m, 4H), 7.31 (s, 1H); 13C-NMR (CDCl3/75 MHz): 21.9, 25.6, 28.4, 29.1, 29.3, 29.4, 29.7, 32.7, 38.3, 39.1, 49.1, 121.2, 123.9, 124.3, 130.8, 131.5, 135.9, 137.5, 138.3, 141.5, 148.3. HRMS (ESI+) calcd for [C42H56N6S10+H]+: 965.1846, found 965.1847.
1,2-Bis(8-(1-(3-(5'-(methylthio)-2,2'-bithiophen-5-ylthio)propyl)-1H-1,2,3-triazol-4-yl)octyl)disulfide (12b). Yield 48%; pale yellow solid; C44H60N6S10; mp: 119–121 °C; IR (KBr): 790, 863, 1051, 1123, 1213, 1328, 1426, 1465, 1497, 1551, 2849, 2919, 3062, 3113 cm−1; 1H-NMR (CDCl3/500 MHz): δ 1.28–1.39 (m, 8H), 1.64–1.70 (m, 4H), 2.19–2.25 (m, 2H), 2.51 (s, 3H), 2.65–2.72 (m, 4H), 2.76 (t, J = 7 Hz, 2H), 4.47 (t, J = 6.5 Hz, 2H), 6.96–7.03 (m, 4H), 7.24 (s, 1H); 13C-NMR (CDCl3/125 MHz): δ 18.5, 22.1, 25.7, 28.5, 29.2, 29.3, 29.5, 29.6, 35.4, 39.2, 48.2, 58.6, 120.9, 123.9, 124.2, 131.6, 132.3, 135.2, 137.3, 138.6, 141.0, 148.6. HRMS (ESI+): calcd for [C44H60N6S10+H]+: 993.2160, found 993.2165.

4. Conclusions

In summary, two series of triazolobithiophenes exhibiting either N–C (compounds 2a-c) or C–C connections (compounds 3a,b) to the aliphatic side chain were synthesized. Each group contained derivatives of varying spacer length (Cn, n = 1, 2, 3 carbons) between the triazole and the bithiophene moiety. Except for 2a and 11a, all the thiols or their corresponding disulfides yielded stable monolayers with high surface coverage. Due to their excellent durability, the new molecules represent ideal candidates for thorough studies on the energy transmitting principle in DIAMS MS [55]. Consequently, the click chemistry appeared to be suitable for the design of new SAMs bearing a 1,2,3-triazole and various chromophors that exhibit, as the pioneering bithiophene 1, an absorption maximum which is close to the emission wavelength of a standard nitrogen laser.

Supplementary Materials

Supplementary materials can be accessed at https://www.mdpi.com/1420-3049/16/10/8758/s1.

Acknowledgments

This work was supported by the Centre National de la Recherche Scientifique (CNRS France), the Agence Nationale de la Recherche (ANR France), and the Région des Pays de la Loire (France).

Conflict of Interest

The authors declare no conflict of interest.

References and Notes

  1. Plunkett, K.N.; Mohraz, A.; Haasch, R.T.; Lewis, J.A.; Moore, J.S. Light-regulated electrostatic interactions in colloidal suspensions. J. Am. Chem. Soc. 2005, 127, 14574–14575. [Google Scholar] [CrossRef]
  2. Bounichou, M.; Sanguinet, L.; Elouarzaki, K.; Alévêque, O.; Dias, M.; Levillain, E.; Rondeau, D. Evaluation of a new matrix-free laser desorption/ionization method through statistic studies: Comparison of the DIAMS (desorption/ionization on self-assembled monolayer surface) method with the MALDI and TGFA-LDI techniques. J. Mass. Spectrom. 2008, 43, 1618–1626. [Google Scholar] [CrossRef] [Green Version]
  3. Sanguinet, L.; Alévêque, O.; Blanchard, P.; Dias, M.; Levillain, E.; Rondeau, D. Desorption/ionization on self-assembled monolayer surfaces (DIAMS). J. Mass. Spectrom. 2006, 41, 830–833. [Google Scholar] [CrossRef] [Green Version]
  4. Nayak, R.; Knapp, D.R. Matrix-free LDI mass spectrometry platform using patterned nanostructured gold thin film. Anal. Chem. 2010, 82, 7772–7778. [Google Scholar] [CrossRef]
  5. Li, Y.; Liu, Y.; Wang, N.; Li, Y.; Liu, H.; Lu, F.; Zhuang, J.; Zhu, D. Self-assembled monolayers of C60-perylenetetracarboxylic diimide-C60 triad on indium tin oxide surface. Carbon 2005, 43, 1968–1975. [Google Scholar] [CrossRef]
  6. Nie, H.-Y.; McIntyre, N.S.; Lau, W.M.; Feng, J.M. Optical properties of an octadecylphosphonic acid self-assembled monolayer on a silicon wafer. Thin Solid Films 2008, 517, 814–818. [Google Scholar] [CrossRef]
  7. Chan, Y.H.; Lin, J.T.; Chen, I.W.; Chen, C.H. Monolayers of diphenyldiacetylene derivatives: Tuning molecular tilt angles and photopolymerization efficiency via electrodeposited Ag interlayer on Au. J. Phys. Chem. B 2005, 109, 19161–19168. [Google Scholar] [CrossRef]
  8. Neilands, O.; Kirichenko, N.; Muzikante, I.; Fonavs, E.; Gerca, L.; Jursenas, S.; Valiokas, R.; Karpicz, R.; Valkunas, L. Detection of blue light by self-assembled monolayer of dipolar molecules. In UV Solid-State Light Emitters and Detectors; Shur, M.S., Zukauskas, A., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2004; NATO Science Series II Mathematics Physics and Chemistry; Volume 144, pp. 261–269. [Google Scholar]
  9. Kondo, M.; Nakamura, Y.; Fujii, K.; Nagata, M.; Suemori, Y.; Dewa, T.; Lida, K.; Gardiner, A.T.; Cogdell, R.J.; Nango, M. Self-assembled monolayer of light-harvesting core complexes from photosynthetic bacteria on a gold electrode modified with alkanethiols. Biomacromolecules 2007, 8, 2457–2463. [Google Scholar]
  10. Imahori, H.; Hasobe, T.; Yamada, H.; Nishimura, Y.; Yamazaki, I.; Fukuzumi, S. Concentration effects of porphyrin monolayers on the structure and photoelectrochemical properties of mixed self-assembled monolayers of porphyrin and alkanethiol on gold electrodes. Langmuir 2001, 17, 4925–4931. [Google Scholar]
  11. Li, Y.; Li, Y.; Liu, H.; Wang, S.; Wang, N.; Zhuang, J.; Li, X.; He, X.; Zhu, D. Self-assembled monolayers of porphyrin perylenetetracarboxylic diimide [60] fullerene on indium tin oxide electrodes: Enhancement of light harvesting in the visible light region. Nanotechnology 2005, 16, 1899–1904. [Google Scholar] [CrossRef]
  12. Mrksich, M. Mass spectroscopy of self-assembled monolayers: A new tool for molecular surface science. ACS Nano 2008, 2, 7–18. [Google Scholar] [CrossRef]
  13. Patrie, S.M.; Mrksich, M. Self-assembled monolayers for MALDI-TOF mass spectrometry for immunoassays of human protein antigens. Anal. Chem. 2007, 79, 5878–5887. [Google Scholar] [CrossRef]
  14. Plutowski, U.; Richert, C. A direct glimpse of cross-hybridization: background-passified microarrays that allowmass-spectrometric detection of captured oligonucleotides. Angew. Chem. Int. Ed. 2005, 44, 621–625. [Google Scholar] [CrossRef]
  15. Zotti, G.; Vercelli, B.; Berlin, A. Monolayers and multilayers of conjugated polymers as nanosized electronic components. Acc. Chem. Res. 2008, 41, 1098–1109. [Google Scholar] [CrossRef]
  16. Zhu, L.; Tang, H.; Harima, Y.; Yamashita, K.; Aso, Y.; Otsubo, T. Enhanced hole injection in organic light-emitting diodes consisting of self-assembled monolayer of tripod-shaped π-conjugated thiols. J. Mater. Chem. 2002, 12, 2250–2254. [Google Scholar]
  17. De Boer, B.; Meng, H.; Perepichka, D.F.; Zheng, J.; Frank, M.M.; Chabal, Y.J.; Bao, Z. Synthesis and characterization of conjugated mono- and dithiol oligomers and characterization of their self-assembled monolayers. Langmuir 2003, 19, 4272–4284. [Google Scholar] [CrossRef]
  18. Tour, J.M.; Leroy, J.; Pearson, D.L.; Lamba, J.J.S.; Burgin, T.P.; Whitesides, G.M.; Allara, D.L.; Parikh, A.T.; Atre, S.V. Self-assembled monolayers and multilayers of conjugated thiols, α,ω-dithiols, and thioacetyl-containing adsorbate. Understanding attachments between potential molecular wires and gold surfaces. J. Am. Chem. Soc. 1995, 117, 9529–9534. [Google Scholar]
  19. Berlin, A.; Zotti, G.; Schiavon, G.; Zecchin, S. Adsorption of carboxyl-terminated dithiophene and terthiophene molecules on ITO electrodes and their electrochemical coupling to polymer layers. The influence of molecular geometry. J. Am. Chem. Soc. 1998, 120, 13453–13460. [Google Scholar] [CrossRef]
  20. Clegg, R.S.; Hutchison, J.E. Control of monolayer assembly structure by hydrogen bonding rather than by adsorbate-substrate templating. J. Am. Chem. Soc. 1999, 121, 5319–5327. [Google Scholar] [CrossRef]
  21. Valiokas, R.; Svedhem, S.; Svensson, S.C.T.; Liedberg, B. Self-assembled monolayers of oligo(ethylene glycol)-terminated and amide group containing alkanethiolates on gold. Langmuir 1999, 15, 3390–3394. [Google Scholar] [CrossRef]
  22. Valiokas, R.; Malysheva, L.; Onipko, A.; Lee, H-H.; Ruželė, Ž.; Svedhem, S.; Svensson, S.C.T.; Gelius, U.; Liedberg, B. On the quality and structural characteristics of oligo(ethylene glycol) assemblies on gold: an experimental and theoretical study. J. Electron. Spectr. Relat. Phenom. 2009, 172, 9–20. [Google Scholar] [CrossRef]
  23. Evans, S.D.; Goppert-Berarducci, K.E.; Urankar, E.; Gerenser, L.J.; Ulman, A. Monolayers having large in-plane dipole moments: Characterization of sulfone-containing self-assembled monolayers of alkanethiols on gold by Fourier transform infrared spectroscopy, X-ray photoelectron spectroscopy and wetting. Langmuir 1991, 7, 2700–2709. [Google Scholar] [CrossRef]
  24. Evans, S.D.; Urankar, E.; Ulman, A.; Ferris, N. Self-assembled monolayers of alkanethiols containing a polar aromatic group: Effects of the dipole position on molecular packing, orientation, and surface wetting propertie. J. Am. Chem. Soc. 1991, 113, 4121–4131. [Google Scholar]
  25. Kang, J.F.; Ulman, A.; Liao, S.; Jordan, R. Mixed self-assembled monolayers of highly polar rigid biphenyl thiols. Langmuir 1999, 15, 2095–2098. [Google Scholar] [CrossRef]
  26. Barrena, E.; Palacios-Lidón, E.; Munuera, C.; Torrelles, X.; Ferrer, S.; Jonas, U.; Salmeron, M.; Ocal, C. The role of intermolecular and molecule-substrate interactions in the stability of alkanethiol nonsaturated phases on Au(111). J. Am. Chem. Soc. 2004, 126, 385–395. [Google Scholar]
  27. Lewis, P.A.; Smith, R.K.; Kelly, K.F.; Bumm, L.A.; Reed, S.M.; Clegg, R.S.; Gunderson, J.D.; Hutchison, J.E.; Weiss, P.S. The role of buried hydrogen bonds in self-assembled mixed composition thiols on Au(111). J. Phys. Chem. B 2001, 105, 10630–10636. [Google Scholar]
  28. Boal, A.K.; Rotello, V.M. Intra- and inter-monolayer hydrogen bonding in amide functionalized alkanethiol SAMs on gold nanoparticles. Langmuir 2000, 16, 9527–9532. [Google Scholar]
  29. Lenk, T.J.; Hallmark, V.M.; Hoffmann, C.L.; Rabolt, J.F.; Castner, D.G.; Erdelen, C.; Ringsdorf, H. Structural investigation of molecular organization in self-assembled monolayers of a semifluorinated amidethiol. Langmuir 1994, 10, 4610–4617. [Google Scholar]
  30. Tam-Chang, S.-W.; Biebuyck, H.A.; Whitesides, G.M.; Nooh, J.; Nuzzo, R.G. Self-assembled monolayers on gold generated from alkanethiols with the structure RNHCOCH2SH. Langmuir 1995, 11, 4371–4382. [Google Scholar]
  31. Sabapathy, R.C.; Bhattacharyya, S.; Leavy, M.C.; Cleland, W.E.; Hussey, C.L. Electrochemical and spectroscopic characterization of self-assembled monolayers of ferrocenylalkyl compounds with amide linkages. Langmuir 1998, 14, 124–136. [Google Scholar] [CrossRef]
  32. Clegg, R.S.; Hutchison, J.E. Hydrogen-bonding, self-assembled monolayers: Ordered molecular films for study of through-peptide electron transfer. Langmuir 1996, 12, 5239–5243. [Google Scholar] [CrossRef]
  33. Clegg, R.S.; Reed, S.M.; Hutchison, J.E. Self-assembled monolayers stabilized by three-dimensional networks of hydrogen bonds. J. Am. Chem. Soc. 1998, 120, 2486–2487. [Google Scholar] [CrossRef]
  34. Clegg, R.S.; Reed, S.M.; Smith, R.K.; Barron, B.L.; Rear, J.A.; Hutchison, J.E. The interplay of lateral and tiered interactions in stratified self-organized molecular assemblies. Langmuir 1999, 15, 8876–8883. [Google Scholar]
  35. Clegg, R.S.; Hutchison, J.E. Control of monolayer assembly structure by hydrogen bonding rather than by adsorbate-substrate templating. J. Am. Chem. Soc. 1999, 121, 5319–5327. [Google Scholar] [CrossRef]
  36. Dou, R.-F.; Ma, X.-C.; Xi, L.; Yip, H.L.; Wong, K.Y.; Lau, W.M.; Jia, J.-F.; Xue, Q.-K.; Yang, W.-S.; Ma, H.; Jen, A.K.-Y. Self-assembled monolayers of aromatic thiols stabilized by parallel-displaced π-π stacking interactions. Langmuir 2006, 22, 3049–3056. [Google Scholar]
  37. Yin, G.; Zhang, Y.; Li, B.; Zhang, Y. Syntheses, structures and luminescent properties of a dimer and an one-dimensional chain coordination polymer with the flexible bis(triazole) and hydroxybenzoate ligands. J. Mol. Struct. 2007, 837, 263–268. [Google Scholar] [CrossRef]
  38. Yu, S.; Chai, X.; Hu, H.; Yan, Y.; Guan, Z.; Zou, Y.; Sun, Q.; Wu, Q. Synthesis and antifungal evaluation of novel triazole derivatives as inhibitors of cytochrome P450 14α-demethylase. Eur. J. Med. Chem. 2010, 45, 4435–4445. [Google Scholar] [CrossRef]
  39. Chelmowski, R.; Käfer, D.; Köster, S.D.; Klasen, T.; Winkler, T.; Terfort, A.; Metzler-Nolte, N.; Wöll, C. Postformation modification of SAMs: Using click chemistry to functionalize organic surfaces. Langmuir 2009, 25, 11480–11485. [Google Scholar]
  40. Lee, J.K.; Chi, Y.S.; Choi, I.S. Reactivity of acetylenyl-terminated self-assembled monolayers on gold: Triazole formation. Langmuir 2004, 20, 3844–3847. [Google Scholar]
  41. Himo, F.; Lovell, T.; Hilgraf, R.; Rostovtsev, V.; Noodleman, L.; Sharpless, K.B.; Fokin, V.V. Copper(I)-catalyzed synthesis of azoles. DFT study predicts unprecedented reactivity and intermediates. J. Am. Chem. Soc. 2005, 127, 210–216. [Google Scholar]
  42. Rostovtsev, V.V.; Green, L.G.; Fokin, V.V.; Sharpless, K.B. A stepwise Huisgen cycloaddition process: Copper(I)-catalyzed regioselective ligation of azides and terminal alkynes. Angew. Chem. Int. Ed. 2002, 41, 2596–2599. [Google Scholar] [CrossRef]
  43. Blanchard, P.; Jousselme, B.; Frère, P.; Roncali, J. 3- and 3,4-bis(2-cyanoethylsulfanyl)thiophenes as building blocks for functionalized thiophene-based π-conjugated systems. J. Org. Chem. 2002, 67, 3961–3964. [Google Scholar] [CrossRef]
  44. Tornøe, C.W.; Christensen, C.; Meldal, M. Peptidotriazoles on solid phase: (1,2,3)-Triazoles by regiospecific copper(I)-catalyzed 1,3-dipolar cycloadditions of terminal alkynes to azides. J. Org. Chem. 2002, 67, 3057–3064. [Google Scholar]
  45. Love, J.C.; Estroff, L.A.; Kriebel, J.K.; Nuzzo, R.G.; Whitesides, G.M. Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chem. Rev. 2005, 105, 1103–1169. [Google Scholar]
  46. Nuzzo, R.G.; Allara, D.L. Adsorption of bifunctional organic disulfides on gold surfaces. J. Am. Chem. Soc. 1983, 105, 4481–4483. [Google Scholar] [CrossRef]
  47. Noh, J.; Murase, T.; Nakajima, K.; Lee, H.; Hara, M. Nanoscopic investigation of the self-assembly processes of dialkyl disulfides and dialkyl sulfides on Au(111). J. Phys. Chem. B 2000, 104, 7411–7416. [Google Scholar]
  48. Biebuyck, H.A.; Bain, C.D.; Whitesides, G.M. Comparison of organic monolayers on polycrystalline gold spontaneously assembled from solutions containing dialkyl disulfides or alkanethiols. Langmuir 1994, 10, 1825–1831. [Google Scholar]
  49. Bain, C.D.; Biebuyck, H.A.; Whitesides, G.M. Comparison of self-assembled monolayers on gold: coadsorption of thiols and disulfides. Langmuir 1989, 5, 723–727. [Google Scholar] [CrossRef]
  50. Strong, L.; Whitesides, G.M. Structures of self-assembled monolayer films of organosulfur compounds adsorbed on gold single crystals: Electron diffraction studies. Langmuir 1988, 4, 546–558. [Google Scholar] [CrossRef]
  51. Guyard, L.; Hapiot, P.; Jouini, M.; Lacroix, J.-C.; Lagrost, C.; Neta, P. Oxidative coupling of small oligothiophenes and oligopyrroles in water in the presence of cyclodextrin pulse radiolysis investigations. J. Phys. Chem. A 1999, 103, 4009–4015. [Google Scholar]
  52. Bounichou, M.; Alévêque, O.; Breton, T.; Dias, M.; Sanguinet, L.; Levillain, E.; Rondeau, D. Self-assembled monolayer-assisted mass spectrometry. J. Mater. Chem. 2009, 19, 8032–8039. [Google Scholar]
  53. Zanolari, B.; Ndjoko, K.; Ioset, J.-R.; Marston, A.; Hostettmann, K. Qualitative and quantitative determination of yohimbine in authentic yohimbe bark and in commercial aphrodisiacs by HPLC-UV-API/MS methods. Phytochem. Anal. 2003, 14, 193–201. [Google Scholar] [CrossRef]
  54. Zenobie, R.; Knochenmuss, R. Ion Formation in MALDI Mass Spectrometry. Mass Spectrom. Rev. 1998, 17, 337–366. [Google Scholar] [CrossRef]
  55. Schinkovitz, A.; Tsague-Kenfack, G.; Levillain, E.; Dias, M.; Helesbeux, J.-J.; Derbré, S.; Séraphin, D.; Richomme, P. Free and immobilized matrix molecules: impairing ionization by quenching secondary ion formation in laser desorption MS. J. Mass. Spectrom. 2011, 46, 884–890. [Google Scholar] [CrossRef] [Green Version]
  • Sample Availability: Samples of the compounds are available from the authors.

Share and Cite

MDPI and ACS Style

Kenfack, G.T.; Schinkovitz, A.; Babu, S.; Elouarzaki, K.; Dias, M.; Derbré, S.; Helesbeux, J.-J.; Levillain, E.; Richomme, P.; Séraphin, D. Triazolobithiophene Light Absorbing Self-Assembled Monolayers: Synthesis and Mass Spectrometry Applications. Molecules 2011, 16, 8758-8774. https://doi.org/10.3390/molecules16108758

AMA Style

Kenfack GT, Schinkovitz A, Babu S, Elouarzaki K, Dias M, Derbré S, Helesbeux J-J, Levillain E, Richomme P, Séraphin D. Triazolobithiophene Light Absorbing Self-Assembled Monolayers: Synthesis and Mass Spectrometry Applications. Molecules. 2011; 16(10):8758-8774. https://doi.org/10.3390/molecules16108758

Chicago/Turabian Style

Kenfack, Ghislain Tsague, Andreas Schinkovitz, Suresh Babu, Kamal Elouarzaki, Marylène Dias, Séverine Derbré, Jean-Jacques Helesbeux, Eric Levillain, Pascal Richomme, and Denis Séraphin. 2011. "Triazolobithiophene Light Absorbing Self-Assembled Monolayers: Synthesis and Mass Spectrometry Applications" Molecules 16, no. 10: 8758-8774. https://doi.org/10.3390/molecules16108758

Article Metrics

Back to TopTop