[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (17)

Search Parameters:
Keywords = vapochromism

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 5151 KiB  
Article
Multi-Stimuli Responsive Viologen-Imprinted Polyvinyl Alcohol and Tricarboxy Cellulose Nanocomposite Hydrogels
by Salhah D. Al-Qahtani, Ghadah M. Al-Senani, Muneera Alrasheedi and Ard elshifa M. E. Mohammed
Sensors 2024, 24(21), 6860; https://doi.org/10.3390/s24216860 - 25 Oct 2024
Viewed by 864
Abstract
Photochromic inks have shown disadvantages, such as poor durability and high cost. Self-healable hydrogels have shown photostability and durability. Herein, a viologen-based covalent polymer was printed onto a paper surface toward the development of a multi-stimuli responsive chromogenic sheet with thermochromic, photochromic, and [...] Read more.
Photochromic inks have shown disadvantages, such as poor durability and high cost. Self-healable hydrogels have shown photostability and durability. Herein, a viologen-based covalent polymer was printed onto a paper surface toward the development of a multi-stimuli responsive chromogenic sheet with thermochromic, photochromic, and vapochromic properties. Viologen polymer was created by polymerizing a dialdehyde-based viologen with a hydroxyl-bearing dihydrazide in an acidic aqueous medium. The viologen polymer was well immobilized as a colorimetric agent into a polyvinyl alcohol (PVA)/tricarboxy cellulose (TCC)-based self-healable hydrogel. The viologen/hydrogel nanocomposite films were applied onto a paper surface. The coloration measurements showed that when exposed to ultraviolet light, the orange layer printed on the paper surface switched to green. The photochromic film was used to develop anti-counterfeiting prints using the organic hydrogel composed of a PVA/TCC composite and a viologen polymer. Reversible photochromism with strong photostability was observed when the printed papers were exposed to UV irradiation. A detection limit was monitored in the range of 0.5–300 ppm for NH3(aq). The exposure to heat (70 °C) was found to reversibly initiate a colorimetric change. Full article
(This article belongs to the Section Nanosensors)
Show Figures

Figure 1

Figure 1
<p>Preparation of tricarboxy cellulose.</p>
Full article ">Figure 2
<p>TEM analysis of viologen polymer particles at different magnifications and different positions in the tested sample (<b>a</b>–<b>f</b>).</p>
Full article ">Figure 3
<p>XRD spectrum of polymer nanoparticles.</p>
Full article ">Figure 4
<p>SEM analysis of cast hydrogel film; VP<sub>6</sub> (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 5
<p>SEM images of printed paper (VP<sub>6</sub>) at different positions on the sample surface (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 6
<p>Absorption spectra of VP<sub>6</sub> below visible (Vis) and ultraviolet (UV) irradiation.</p>
Full article ">Figure 7
<p>Reversibility of absorbance of VP<sub>6</sub> over numerous cycles of ultraviolet (598 nm) and visible (430 nm) illumination.</p>
Full article ">Figure 8
<p>Effect of viologen content on mechanical performance of stamped samples, including tensile strength, strain, and Young’s modulus.</p>
Full article ">Figure 9
<p>Effect of shearing rate on the viscosity of VP<sub>6</sub>.</p>
Full article ">Figure 10
<p>Absorption spectra of VP<sub>6</sub> at different temperatures.</p>
Full article ">Figure 11
<p>Thermochromism of VP<sub>6</sub> showing a change in color from orange (<b>a</b>) to greenish (<b>b</b>) with heating from 25 °C to 70 °C, respectively.</p>
Full article ">Figure 12
<p>Absorbance spectra of paper (VP<sub>6</sub>) under air and different concentrations of NH<sub>3(g)</sub>.</p>
Full article ">Figure 13
<p>Absorption intensity of VP<sub>6</sub> versus various concentrations of NH<sub>3(aq)</sub>.</p>
Full article ">Scheme 1
<p>Synthesis process of viologen polymer.</p>
Full article ">
12 pages, 4667 KiB  
Article
Multistimuli Luminescence and Anthelmintic Activity of Zn(II) Complexes Based on 1H-Benzimidazole-2-yl Hydrazone Ligands
by Alexey Gusev, Elena Braga, Alexandr Kaleukh, Michail Baevsky, Mikhail Kiskin and Wolfgang Linert
Inorganics 2024, 12(9), 256; https://doi.org/10.3390/inorganics12090256 - 23 Sep 2024
Viewed by 812
Abstract
Three novel Zn(II) mononuclear complexes with the general formula ZnL2Cl2 (L = 2-(4-R-phenylmethylene)benzimidazol-2-hydrazines; R-H (1), R-CH3 (2), and R-OCH3 (3)) were synthesized and fully characterized by various means. These complexes demonstrate excitation-dependent emission, which is detected by a [...] Read more.
Three novel Zn(II) mononuclear complexes with the general formula ZnL2Cl2 (L = 2-(4-R-phenylmethylene)benzimidazol-2-hydrazines; R-H (1), R-CH3 (2), and R-OCH3 (3)) were synthesized and fully characterized by various means. These complexes demonstrate excitation-dependent emission, which is detected by a change in the emission color (from blue to green) upon an increase in the excitation wavelength. Moreover complex 1 shows reversible mechanochromic luminescence behavior due to the reversible loss of solvated methanol molecules upon the intense grinding of crystals. In addition, 1 exhibits vapochromic properties, which originate from the adsorption methanol vapor on the crystal surface. The strengthening of anthelmintic activity at the transition from free hydrazones to zinc-based complexes is shown. Full article
(This article belongs to the Section Coordination Chemistry)
Show Figures

Figure 1

Figure 1
<p>Crystal and molecular structures of 1 (<b>a</b>), 2 (<b>c</b>), and 3 (<b>d</b>). Solvate molecules are omitted for clarity. Fragment of supramolecular architecture of 1 (<b>b</b>) (Colour of the atoms: grey—C; red—O; green—Cl; blue—Zn; deep blue—N).</p>
Full article ">Figure 2
<p>(<b>a</b>) Emission spectra of as-prepared crystals of 1 at different excitation. (<b>b</b>) Emission spectra of 1 after treatment of CH<sub>3</sub>OH vapor at different excitation. (<b>c</b>) Emission spectra of 1 after grinding (exc. at 430 nm) (inset emission excited at 365 nm). (<b>d</b>) Image of as-prepared crystals after treatment and after grinding under UV lamp. (<b>e</b>) Emission spectra of as-prepared crystals of 2 at different excitation. (<b>f</b>) Emission spectra of as-prepared crystals of 3 at different excitation.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Emission spectra of as-prepared crystals of 1 at different excitation. (<b>b</b>) Emission spectra of 1 after treatment of CH<sub>3</sub>OH vapor at different excitation. (<b>c</b>) Emission spectra of 1 after grinding (exc. at 430 nm) (inset emission excited at 365 nm). (<b>d</b>) Image of as-prepared crystals after treatment and after grinding under UV lamp. (<b>e</b>) Emission spectra of as-prepared crystals of 2 at different excitation. (<b>f</b>) Emission spectra of as-prepared crystals of 3 at different excitation.</p>
Full article ">Figure 3
<p>Optimized geometries of chemical structures of complexes 1–3.</p>
Full article ">Figure 4
<p>HOMO and LUMO orbitals of all molecules. <span class="html-italic">(pink and purple colours means positive and negative sign of wave function).</span></p>
Full article ">Scheme 1
<p>Ligands L1–L3 used in this study.</p>
Full article ">Scheme 2
<p>Synthesis of the Zn complexes with L1–L3.</p>
Full article ">
1 pages, 142 KiB  
Correction
Correction: Crespo-Cajigas et al. Development of a Paper-Based Sol–Gel Vapochromic Sensor for the Detection of Vapor Cross-Contamination within a Closed Container. Analytica 2024, 5, 295–310
by Janet Crespo-Cajigas, Abuzar Kabir, Joel Carrasco, Amatullah Shahid, Kenneth G. Furton and Lauryn E. DeGreeff
Analytica 2024, 5(3), 430; https://doi.org/10.3390/analytica5030027 - 3 Sep 2024
Viewed by 1000
Abstract
Addition of Two Authors [...] Full article
16 pages, 5237 KiB  
Article
Development of a Paper-Based Sol–Gel Vapochromic Sensor for the Detection of Vapor Cross-Contamination within a Closed Container
by Janet Crespo-Cajigas, Abuzar Kabir, Joel Carrasco, Amatullah Shahid, Kenneth G. Furton and Lauryn E. DeGreeff
Analytica 2024, 5(3), 295-310; https://doi.org/10.3390/analytica5030019 - 7 Jul 2024
Cited by 1 | Viewed by 1461 | Correction
Abstract
Contamination of trace levels of volatile organic compounds (VOCs) in enclosed spaces is not usually a significant cause for concern; however, it can be relevant in the case of canine scent detection training as a canine’s superior sense of smell makes them highly [...] Read more.
Contamination of trace levels of volatile organic compounds (VOCs) in enclosed spaces is not usually a significant cause for concern; however, it can be relevant in the case of canine scent detection training as a canine’s superior sense of smell makes them highly likely to detect low levels of contamination, contributing to inefficient training. Thus, herein, we address the need for a simple, low-cost, robust, vapochromic sensor to determine the cross-contamination of VOCs within closed containers, such as canine training aid kits. This study focuses on the development of a vapor sensor, which produces a rapid colorimetric change when a target chemical vapor is present. A pH indicator is used as the colorimetric dye and its incorporation into a sol–gel matrix on a paper substrate is confirmed via SEM characterization. The sensor’s stability and performance is tested against exposure to various levels of sunlight and temperature. The design allows the sensor to present a clear and unambiguous visible response to the release of the volatile target within a closed container. It can be readily incorporated into existing training kits and functions as a straightforward reminder of when training aids need to be changed or a new containment system should be considered. Full article
(This article belongs to the Section Sensors)
Show Figures

Figure 1

Figure 1
<p>Average reaction time for BCG sensors prepared at different dye concentrations exposed to various NH<sub>4</sub>OH dilutions (n = 3).</p>
Full article ">Figure 2
<p>Bromocresol green sensor color change from green (acid) to blue (basic).</p>
Full article ">Figure 3
<p>Comparison of pH measurements for NH<sub>4</sub>OH dilutions in headspace and in solution (n = 4).</p>
Full article ">Figure 4
<p>SEM imaging of (<b>A</b>) plain filter paper and (<b>B</b>) bromocresol green sol–gel-coated sensor.</p>
Full article ">Figure 5
<p>Progression of sensor color throughout the 10-week period for the stability experiment.</p>
Full article ">Figure 6
<p>Individual sensor RGB value trends across 10 weeks for each of the cases examined over time (normalized data; n = 3).</p>
Full article ">Figure 7
<p>Progression of sensor color throughout the 10-week period for the sunlight experiment.</p>
Full article ">Figure 8
<p>Individual sensor RGB values across 10 weeks for each of the sunlight exposure variables (normalized data; n = 2).</p>
Full article ">
21 pages, 10605 KiB  
Article
Enhancing Vapochromic Properties of Platinum(II) Terpyridine Chloride Hexaflouro Phosphate in Terms of Sensitivity through Nanocrystalization for Fluorometric Detection of Acetonitrile Vapors
by Sedigheh Barzegar, Mahmood Karimi Abdolmaleki, William B. Connick and Ghodratollah Absalan
Crystals 2024, 14(4), 347; https://doi.org/10.3390/cryst14040347 - 5 Apr 2024
Viewed by 1422
Abstract
The vapochromic properties of [Pt(tpy)Cl](PF6) crystals in the presence of acetonitrile and its effect on the crystal structure as well as the fluorescence spectrum of this complex have already been studied in the past. We synthesized nanocrystals of this compound for [...] Read more.
The vapochromic properties of [Pt(tpy)Cl](PF6) crystals in the presence of acetonitrile and its effect on the crystal structure as well as the fluorescence spectrum of this complex have already been studied in the past. We synthesized nanocrystals of this compound for the first time, and discussed different parameters and methods that affect nanocrystal structure modulation. The study demonstrates the vapochromic properties of the nanocrystals toward acetonitrile vapor by investigating the morphology and fluorescence spectra of the nanocrystals. Vapochromic studies were conducted on [Pt(tpy)Cl](PF6) nanocrystals for five cycles of absorption and desorption of acetonitrile, demonstrating shorter response times compared to regular bulk crystals. Full article
(This article belongs to the Special Issue 1D and 2D Nanomaterials for Sensor Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The chemical structure of the [Pt(tpy)Cl](PF<sub>6</sub>) [<a href="#B11-crystals-14-00347" class="html-bibr">11</a>].</p>
Full article ">Figure 2
<p>Characterization of [Pt(tpy)Cl](PF<sub>6</sub>) crystals by <sup>1</sup>HNMR.</p>
Full article ">Figure 3
<p>The X-ray powder diffractogram of the [Pt(tpy)Cl](PF<sub>6</sub>) crystals (―) and the simulated pattern from the mercury software (<span style="color:#ED0000">―</span>) [<a href="#B30-crystals-14-00347" class="html-bibr">30</a>].</p>
Full article ">Figure 4
<p>Fluorescence spectra of the crystals of Pt(II) complexes (by time) during the first absorption (<b>A</b>) and desorption (<b>B</b>) of acetonitrile vapor (λ<sub>ex</sub> = 436 nm).</p>
Full article ">Figure 5
<p>Fluorescence spectra of the crystals of Pt(II) complexes (by time) during the second (<b>A</b>,<b>B</b>), third (<b>C</b>,<b>D</b>), fourth (<b>E</b>,<b>F</b>), and fifth (<b>G</b>,<b>H</b>) absorption and desorption of acetonitrile vapor (λ<sub>ex</sub> = 436 nm).</p>
Full article ">Figure 5 Cont.
<p>Fluorescence spectra of the crystals of Pt(II) complexes (by time) during the second (<b>A</b>,<b>B</b>), third (<b>C</b>,<b>D</b>), fourth (<b>E</b>,<b>F</b>), and fifth (<b>G</b>,<b>H</b>) absorption and desorption of acetonitrile vapor (λ<sub>ex</sub> = 436 nm).</p>
Full article ">Figure 6
<p>The SEM images of the high-quality [Pt(tpy)Cl](PF<sub>6</sub>) crystals with different magnifications: (<b>A</b>) 100 µm, (<b>B</b>) 20 µm, (<b>C</b>) 20 µm, and (<b>D</b>) 10 µm.</p>
Full article ">Figure 7
<p>The microscopic images of [Pt(tpy)Cl](PF<sub>6</sub>) crystals’ color change (<b>A</b>) before and (<b>B</b>) after exposure to MeCN vapor.</p>
Full article ">Figure 8
<p>The SEM images of the [Pt(tpy)Cl](PF<sub>6</sub>) crystals after (<b>A</b>) first, (<b>B</b>) second, and (<b>C</b>) fifth absorption/desorption of MeCN vapor.</p>
Full article ">Figure 9
<p>The SEM images of the [Pt(tpy)Cl](PF<sub>6</sub>) nanocrystals synthesized under different preparation conditions: (<b>A</b>) Method 1: Deposition of dispersed [Pt(tpy)Cl](PF<sub>6</sub>) onto a silicon substrate, (<b>B</b>) Method 2: Deposition of dispersed [Pt(tpy)Cl](PF<sub>6</sub>) onto a preheated silicon substrate, (<b>C</b>) Method 3: Direct deposition of [Pt(tpy)Cl](PF<sub>6</sub>) onto a preheated silicon substrate.</p>
Full article ">Figure 10
<p>The EDX obtained from nanocrystals of [Pt(tpy)Cl](PF<sub>6</sub>), which confirms the presence of the elements of the compound.</p>
Full article ">Figure 11
<p>Effect of Pt[(tpy)Cl]PF<sub>6</sub> stock solution concentration ((<b>A</b>) 8.0 × 10<sup>−4</sup> mol L<sup>−1</sup>, (<b>B</b>) 1.6 × 10<sup>−3</sup> mol L<sup>−1</sup>, (<b>C</b>) 3.2 × 10<sup>−3</sup> mol L<sup>−1</sup>) on nanocrystal structures.</p>
Full article ">Figure 12
<p>Fluorescence spectra of the nanocrystals of [Pt(tpy)Cl](PF<sub>6</sub>) (by time) during the first (<b>A</b>) absorption and (<b>B</b>) desorption of MeCN vapor (λ<sub>ex</sub> = 436 nm).</p>
Full article ">Figure 13
<p>Fluorescence spectra of the nanocrystals of Pt(II) complexes (by time) during the second to fifth absorption (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and desorption (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) of MeCN vapor (λ<sub>ex</sub> = 436 nm).</p>
Full article ">Figure 13 Cont.
<p>Fluorescence spectra of the nanocrystals of Pt(II) complexes (by time) during the second to fifth absorption (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and desorption (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) of MeCN vapor (λ<sub>ex</sub> = 436 nm).</p>
Full article ">Figure 14
<p>The SEM images of [Pt(tpy)Cl](PF<sub>6</sub>) nanocrystals (<b>A</b>) before exposure to MeCN vapor (66–122 nm), (<b>B</b>) merged together (120–450 nm) after exposure to MeCN vapor, (<b>C</b>) after 5 cycles of MeCN vapor absorption/desorption.</p>
Full article ">Figure 15
<p>The [Pt(tpy)Cl](PF<sub>6</sub>) nanocrystals response time during MeCN vapor absorption at (<b>A</b>) λ = 585 nm and (<b>B</b>) λ = 720 nm.</p>
Full article ">Figure 16
<p>The [Pt(tpy)Cl](PF<sub>6</sub>) nanocrystals response time during MeCN vapor desorption at (<b>A</b>) λ = 720 nm and (<b>B</b>) λ = 585 nm.</p>
Full article ">Scheme 1
<p>Instrumental set up for the fluorescence study of absorption and desorption of acetonitrile onto and from [Pt(tpy)Cl](PF<sub>6</sub>) crystals.</p>
Full article ">Scheme 2
<p>Synthesis of Pt[(tpy)Cl]PF<sub>6</sub> nanocrystals using [Pt(tpy)Cl](PF<sub>6</sub>) at a concentration of 1.6 × 10<sup>−3</sup> mol L<sup>−1</sup> in a mixture of acetone and water (95:5, <span class="html-italic">v</span>/<span class="html-italic">v</span>). (<b>A</b>) Method #1: dispersion of the Pt[(tpy)Cl]PF<sub>6</sub> solution in hexane, followed by deposition onto a silicon substrate. (<b>B</b>) Method #2: dispersion of the Pt[(tpy)Cl]PF<sub>6</sub> solution in hexane, followed by deposition onto a preheated silicon substrate. (<b>C</b>) Method #3: deposition of the Pt[(tpy)Cl]PF<sub>6</sub> solution onto a preheated silicon substrate.</p>
Full article ">Scheme 3
<p>The interaction of Pt-Pt orbitals to form dσ and dσ* orbitals and the possibility of MMLCT [<a href="#B10-crystals-14-00347" class="html-bibr">10</a>,<a href="#B31-crystals-14-00347" class="html-bibr">31</a>]. In detail, the closer distance of the dz<sup>2</sup> orbitals in MeCN-vapor-exposed crystals produces dσ and dσ* orbitals. The higher energy of the dσ* orbital allows easier electron transfer to ligand π* orbitals.</p>
Full article ">
12 pages, 5995 KiB  
Communication
Keto-Adamantane-Based Macrocycle Crystalline Supramolecular Assemblies Showing Selective Vapochromism to Tetrahydrofuran
by Zunhua Li, Yingzi Tan, Manhua Ding, Linli Tang and Fei Zeng
Molecules 2024, 29(3), 719; https://doi.org/10.3390/molecules29030719 - 4 Feb 2024
Cited by 3 | Viewed by 1359
Abstract
Here, we report the synthesis of adamantane-based macrocycle 2 by combining adamantane building blocks with π-donor 1,3-dimethoxy-benzene units. An unpredictable keto-adamantane-based macrocycle 3 was obtained by the oxidation of 2 using DDQ as an oxidant. Moreover, a new type of macrocyclic molecule-based CT [...] Read more.
Here, we report the synthesis of adamantane-based macrocycle 2 by combining adamantane building blocks with π-donor 1,3-dimethoxy-benzene units. An unpredictable keto-adamantane-based macrocycle 3 was obtained by the oxidation of 2 using DDQ as an oxidant. Moreover, a new type of macrocyclic molecule-based CT cocrystal was prepared through exo-wall CT interactions between 3 and DDQ. The cocrystal material showed selective vapochromism behavior towards THF, specifically, among nine volatile organic solvents commonly used in the laboratory. Powder X-ray diffraction; UV-Vis diffuse reflectance spectroscopy; 1H NMR; and single crystal X-ray diffraction analyses revealed that color changes are attributed to the vapor-triggered decomplexation of cocrystals. Full article
(This article belongs to the Special Issue Macrocyclic Compounds: Derivatives and Applications)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of <b>3</b> and DDQ.</p>
Full article ">Figure 2
<p>Partial <sup>1</sup>H NMR spectra (400 MHz, CDCl<sub>3</sub>, 298 K) of (<b>a</b>) <b>3</b> and 1.0 equiv. of DDQ, and (<b>b</b>) free <b>3</b>. [3]<sub>0</sub> = 4.0 mM. Inset: photograph showing colors of <b>3</b>, <b>3</b> + DDQ and DDQ in CHCl<sub>3</sub>.</p>
Full article ">Figure 3
<p>Single crystal structures of <b>2</b> (<b>a</b>,<b>b</b>); single crystal structures of <b>3</b> (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 4
<p>Single crystal structures of <b>3</b>@DDQ (<b>a</b>,<b>b</b>); stacking mode of <b>3</b>@DDQ (<b>c</b>).</p>
Full article ">Figure 5
<p>Photographs of <b>3</b>, <b>3</b>@DDQ and DDQ (<b>a</b>); PXRD patterns of <b>3</b>, <b>3</b>@DDQ and DDQ (<b>b</b>). I: <b>3</b>, II: DDQ, III: <b>3</b>@DDQ, IV: simulated from <b>3</b>@DDQ; (<b>c</b>) diffuse reflectance spectra of <b>3</b>, <b>3</b>@DDQ and DDQ.</p>
Full article ">Figure 6
<p>Photographs of <b>3</b>@DDQ after exposure to various vapors (<b>a</b>,<b>b</b>); (<b>c</b>) PXRD patterns of <b>3</b>@DDQ before (I) and after (II) exposure to CH<sub>2</sub>Cl<sub>2</sub>, (III) CHCl<sub>3</sub>, (IV) THF, (V) 1,4-dioxane, (VI) EtOAc, (VII) benzene, (VIII) <span class="html-italic">n</span>-hexane, (IX) EtOH and (X) ClCH<sub>2</sub>CH<sub>2</sub>Cl; (<b>d</b>) diffuse reflectance spectra of <b>3</b>@DDQ before and after exposure to various vapors.</p>
Full article ">Scheme 1
<p>The synthesis route of macrocyclic molecules <b>2</b> and <b>3</b>.</p>
Full article ">
21 pages, 4293 KiB  
Article
Synthesis of Vapochromic Dyes Having Sensing Properties for Vapor Phase of Organic Solvents Used in Semiconductor Manufacturing Processes and Their Application to Textile-Based Sensors
by Junheon Lee, Duyoung Kim and Taekyeong Kim
Sensors 2022, 22(12), 4487; https://doi.org/10.3390/s22124487 - 14 Jun 2022
Cited by 5 | Viewed by 2090
Abstract
Two vapochromic dyes (DMx and DM) were synthesized to be used for textile-based sensors detecting the vapor phase of organic solvents. They were designed to show sensitive color change properties at a low concentration of vapors at room temperature. They were applied to [...] Read more.
Two vapochromic dyes (DMx and DM) were synthesized to be used for textile-based sensors detecting the vapor phase of organic solvents. They were designed to show sensitive color change properties at a low concentration of vapors at room temperature. They were applied to cotton fabrics as a substrate of the textile-based sensors to examine their sensing properties for nine organic solvents frequently used in semiconductor manufacturing processes, such as trichloroethylene, dimethylacetamide, iso-propanol, methanol, n-hexane, ethylacetate, benzene, acetone, and hexamethyldisilazane. The textile sensor exhibited strong sensing properties of polar solvents rather than non-polar solvents. In particular, the detection of dimethylacetamide was the best, showing a color difference of 15.9 for DMx and 26.2 for DM under 300 ppm exposure. Even at the low concentration of 10 ppm of dimethylacetamide, the color change values reached 7.7 and 13.6, respectively, in an hour. The maximum absorption wavelength of the textile sensor was shifted from 580 nm to 550 nm for DMx and 550 nm to 540 nm for DM, respectively, due to dimethylacetamide exposure. The sensing mechanism was considered to depend on solvatochromism, the aggregational properties of the dyes and the adsorption amounts of the solvent vapors on the textile substrates to which the dyes were applied. Finally, the reusability of the textile sensor was tested for 10 cycles. Full article
(This article belongs to the Topic Advanced Nanomaterials for Sensing Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of textile-based vapor sensor.</p>
Full article ">Figure 2
<p>Synthesis procedure of two vapochromic dyes.</p>
Full article ">Figure 3
<p>Color strengths at λ<sub>max</sub> (<b>a</b>) and color differences (Δ<span class="html-italic">E</span>) in cellulosic fabrics dyed with DMx and DM before and after exposure to DMAc (<b>b</b>).</p>
Full article ">Figure 4
<p>Color differences (Δ<span class="html-italic">E</span>) in cellulosic fabrics dyed with DMx and DM before and after exposure to semiconductor processing solvents.</p>
Full article ">Figure 5
<p>(<b>a</b>) Rate of color change in cellulosic fabrics dyed with DMx and DM upon exposure to DMAc, MeOH, and acetone. (<b>b</b>) Change in color spectra of dyed samples exposed to 300 ppm DMAc for 24 h. (<b>c</b>) Color coordination before and after exposure to 300 ppm DMAc.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) Rate of color change in cellulosic fabrics dyed with DMx and DM upon exposure to DMAc, MeOH, and acetone. (<b>b</b>) Change in color spectra of dyed samples exposed to 300 ppm DMAc for 24 h. (<b>c</b>) Color coordination before and after exposure to 300 ppm DMAc.</p>
Full article ">Figure 6
<p>Visualization of the concept of textile-based vapor sensor.</p>
Full article ">Figure 7
<p>(<b>a</b>) Absorption spectra DMx and DM dissolved in various solvents with different dielectric constants, (<b>b</b>) relationship between the maximum absorption wavelength and solvent dielectric constant.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>c</b>,<b>d</b>,<b>f</b>) Molecular orbital energy diagrams and isodensity surface plots of DMx and DM; (<b>b</b>,<b>e</b>) molecular orbital energy gap tendency for solvent dielectric constants.</p>
Full article ">Figure 9
<p>XRD patterns of DMx and DM before and after exposure to DMAc.</p>
Full article ">Figure 10
<p>Relationship between the adsorption amounts of solvents on cellulosic fabrics and color difference before and after exposure to solvents.</p>
Full article ">Figure 11
<p>Relationship between the rate of adsorption of solvents on fabrics and color change.</p>
Full article ">Figure 12
<p>Reusability of the textile-based vapor sensors.</p>
Full article ">
17 pages, 2788 KiB  
Article
Interrogating the Behaviour of a Styryl Dye Interacting with a Mesoscopic 2D-MOF and Its Luminescent Vapochromic Sensing
by Maria Rosaria di Nunzio, Mario Gutiérrez, José María Moreno, Avelino Corma, Urbano Díaz and Abderrazzak Douhal
Int. J. Mol. Sci. 2022, 23(1), 330; https://doi.org/10.3390/ijms23010330 - 28 Dec 2021
Cited by 7 | Viewed by 2848
Abstract
In this contribution, we report on the solid-state-photodynamical properties and further applications of a low dimensional composite material composed by the luminescent trans-4-(dicyanomethylene)-2-methyl-6-(4-dimethylaminostyryl)-4H-pyran (DCM) dye interacting with a two-dimensional-metal organic framework (2D-MOF), Al-ITQ-HB. Three different samples with increasing concentration of DCM are synthesized [...] Read more.
In this contribution, we report on the solid-state-photodynamical properties and further applications of a low dimensional composite material composed by the luminescent trans-4-(dicyanomethylene)-2-methyl-6-(4-dimethylaminostyryl)-4H-pyran (DCM) dye interacting with a two-dimensional-metal organic framework (2D-MOF), Al-ITQ-HB. Three different samples with increasing concentration of DCM are synthesized and characterized. The broad UV-visible absorption spectra of the DCM/Al-ITQ-HB composites reflect the presence of different species of DCM molecules (monomers and aggregates). In contrast, the emission spectra are narrower and exhibit a bathochromic shift upon increasing the DCM concentration, in agreeance with the formation of adsorbed aggregates. Time-resolved picosecond (ps)-experiments reveal multi-exponential behaviors of the excited composites, further confirming the heterogeneous nature of the samples. Remarkably, DCM/Al-ITQ-HB fluorescence is sensitive to vapors of electron donor aromatic amine compounds like aniline, methylaniline, and benzylamine due to a H-bonding-induced electron transfer (ET) process from the analyte to the surface-adsorbed DCM. These findings bring new insights on the photobehavior of a well-known dye when interacting with a 2D-MOF and its possible application in sensing aniline derivatives. Full article
(This article belongs to the Section Materials Science)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Photographs of the hybrid composites (under daylight) containing increasing amounts of DCM. UV-visible absorption (<b>B</b>) and emission (<b>C</b>) spectra of DCM/Al-ITQ-HB in the solid-state at different dye concentrations: 1 × 10<sup>−3</sup> M (blue solid line), 1 × 10<sup>−4</sup> M (green dashed line), and 5 × 10<sup>−6</sup> M (pink dashed-dotted line). Each absorption spectrum has been corrected by subtracting the absorption from pristine Al-ITQ-HB. For emission spectra, the excitation wavelength is 470 nm.</p>
Full article ">Figure 2
<p>Normalized (to the maximum intensity) magic-angle emission decays of DCM/Al-ITQ-HB at different initial dye concentrations: 1 × 10<sup>−3</sup> M (blue triangles), 1 × 10<sup>−4</sup> M (green squares), and 5 × 10<sup>−6</sup> M (pink circles) upon excitation at 470 nm and observing at 550–575 (<b>A</b>) and 700–750 (<b>B</b>) nm. The solid lines are from the best-fit using a multi-exponential function. IRF (black dashed line) is the instrumental response function.</p>
Full article ">Figure 3
<p>Normalized (to the maximum intensity) TRES of DCM/Al-ITQ-HB using initial dye concentration of 1 × 10<sup>−3</sup> M upon excitation at 470 nm and gating at the indicated delay times.</p>
Full article ">Figure 4
<p>(<b>A</b>) Schematic representation of the HOMO-LUMO energy levels (in gas phase) for the DCM dye and the aromatic compounds used for the quenching test (aniline, methylaniline, dimethylaniline, benzylamine, and toluene). (<b>B</b>–<b>F</b>) Emission spectra of DCM/Al-ITQ-HB ([DCM]<sub>0</sub> = 1 × 10<sup>−4</sup> M) in the absence and presence of aniline (<b>B</b>), methylaniline (<b>C</b>), benzylamine (<b>D</b>), dimethylaniline (<b>E</b>), and toluene (<b>F</b>) atmospheres. For (<b>B</b>–<b>D</b>) the percentage of emission quenching is displayed in the figure.</p>
Full article ">Scheme 1
<p>Schematic representation (not in scale) illustrating the Al-ITQ-HB framework together with the molecular structure of DCM.</p>
Full article ">Scheme 2
<p>Illustration (not in scale) of the energetic levels of DCM monomers, H- and J-aggregates. The dotted arrow corresponds to a forbidden transition, while the filled ones reflect the absorption and emission transitions together with the corresponding wavelengths and lifetimes. The scheme also illustrates the EnT happening between H-aggregates and from H-aggregates to monomers.</p>
Full article ">Scheme 3
<p>Schematic representation (not in scale) of the emission quenching mechanism of DCM/Al-ITQ-HB in presence of aniline.</p>
Full article ">
10 pages, 3424 KiB  
Article
Multifunctional Viologen-Derived Supramolecular Network with Photo/Vapochromic and Proton Conduction Properties
by Chuanqi Zhang, Huaizhong Shi, Chenghui Zhang, Yan Yan, Zhiqiang Liang and Jiyang Li
Molecules 2021, 26(20), 6209; https://doi.org/10.3390/molecules26206209 - 14 Oct 2021
Cited by 3 | Viewed by 2583
Abstract
A supramolecular network [H4bdcbpy(NO3)2·H2O] (H4bdcbpy = 1,1′-Bis(3,5-dicarboxybenzyl)-4,4′-bipyridinium) (1) was prepared by a zwitterionic viologen carboxylate ligand in hydrothermal synthesis conditions. The as-synthesized (1) has been well characterized by means [...] Read more.
A supramolecular network [H4bdcbpy(NO3)2·H2O] (H4bdcbpy = 1,1′-Bis(3,5-dicarboxybenzyl)-4,4′-bipyridinium) (1) was prepared by a zwitterionic viologen carboxylate ligand in hydrothermal synthesis conditions. The as-synthesized (1) has been well characterized by means of single-crystal/powder X-ray diffraction, elemental analysis, thermogravimetric analysis and infrared and UV-vis spectroscopy. This compound possesses a three-dimensional supramolecular structure, formed by the hydrogen bond and π–π interaction between the organic ligands. This compound shows photochromic properties under UV light, as well as vapochromic behavior upon exposure to volatile amines and ammonia, in which the electron transfer from electron-rich parts to the electron-deficient viologen unit gives rise to colored radicals. Moreover, the intensive intermolecular H-bonding networks in 1 endows it with a proton conductivity of 1.06 × 10−3 S cm−1 in water at 90 °C. Full article
(This article belongs to the Special Issue Frontiers in Coordination Polymers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The conformation of H<sub>4</sub>bdcbpy and the supramolecular layer formed by π–π type interactions (highlighted by the pink dotted lines). (<b>b</b>) The hydrogen bond interactions between the adjacent layers along <span class="html-italic">a</span> direction. (<b>c</b>) The free H<sub>2</sub>O and nitrate ions between the layers. (<b>d</b>) The O···N distance between adjacent carboxylate O atom, nitrate O atoms and bipyridinium N atom (Color modes: C—gray, N—blue, O—red and H—white. For clarity, some hydrogen atoms in (<b>a</b>,<b>b</b>,<b>d</b>) were omitted).</p>
Full article ">Figure 2
<p>(<b>a</b>) The A C impedance plots and (<b>b</b>) an Arrhenius-type plot of compound <b>1</b> at different temperatures in water.</p>
Full article ">Figure 3
<p>(<b>a</b>) Photographs, (<b>b</b>) Normalized UV−vis adsorption spectra and (<b>c</b>) EPR spectra of compound <b>1</b> before and after irradiation with a UV lamp.</p>
Full article ">Figure 4
<p>Photographs of compound <b>1</b> before and after exposure to different volatile amine vapors.</p>
Full article ">Figure 5
<p>(<b>a</b>) Normalized UV−vis absorption spectra and (<b>b</b>) EPR spectra of compound <b>1</b> before and after exposure to different volatile amine/ammonia vapors, (EA = ethylamine, DEA = diethylamine and TEA = triethylamine).</p>
Full article ">Scheme 1
<p>1,1′-bis(3,5-dicarboxybenzyl)-4,4′-bipyridinium dibromide (H<sub>4</sub>bdcbpy∙Br<sub>2</sub>).</p>
Full article ">Scheme 2
<p>A viologen-based supramolecular framework with photo/vapochromic and proton conduction properties.</p>
Full article ">
16 pages, 28540 KiB  
Article
Magnetic Switching in Vapochromic Oxalato-Bridged 2D Copper(II)-Pyrazole Compounds for Biogenic Amine Sensing
by Nadia Marino, María Luisa Calatayud, Marta Orts-Arroyo, Alejandro Pascual-Álvarez, Nicolás Moliner, Miguel Julve, Francesc Lloret, Giovanni De Munno, Rafael Ruiz-García and Isabel Castro
Magnetochemistry 2021, 7(5), 65; https://doi.org/10.3390/magnetochemistry7050065 - 12 May 2021
Cited by 7 | Viewed by 2732
Abstract
A new two-dimensional (2D) coordination polymer of the formula {Cu(ox)(4-Hmpz)·1/3H2O}n (1) (ox = oxalate and 4-Hmpz = 4-methyl-1H-pyrazole) has been prepared, and its structure has been determined by single-crystal X-ray diffraction. It consists of corrugated oxalato-bridged [...] Read more.
A new two-dimensional (2D) coordination polymer of the formula {Cu(ox)(4-Hmpz)·1/3H2O}n (1) (ox = oxalate and 4-Hmpz = 4-methyl-1H-pyrazole) has been prepared, and its structure has been determined by single-crystal X-ray diffraction. It consists of corrugated oxalato-bridged copper(II) neutral layers featuring two alternating bridging modes of the oxalate group within each layer, the symmetric bis-bidentate (μ-κ2O1,O2:κ2O2′,O1′) and the asymmetric bis(bidentate/monodentate) (μ4-κO1:κ2O1,O2:κO2′:κ2O2′,O1′) coordination modes. The three crystallographically independent six-coordinate copper(II) ions that occur in 1 have tetragonally elongated surroundings with three oxygen atoms from two oxalate ligands, a methylpyrazole-nitrogen defining the equatorial plane, and two other oxalate-oxygen atoms occupying the axial positions. The monodentate 4-Hmpz ligands alternatively extrude above and below each oxalate-bridged copper(II) layer, and the water molecules of crystallization are located between the layers. Compound 1 exhibits a fast and selective adsorption of methylamine vapors to afford the adsorbate of formula {Cu(ox)(4-Hmpz)·3MeNH2·1/3H2O}n (2), which is accompanied by a concomitant color change from cyan to deep blue. Compound 2 transforms into {Cu(ox)(4-Hmpz)·MeNH2·1/3H2O}n (3) under vacuum for three hours. The cryomagnetic study of 13 revealed a unique switching from strong (1) to weak (2 and 3) antiferromagnetic interactions. The external control of the optical and magnetic properties along this series of compounds might make them suitable candidates for switching optical and magnetic devices for chemical sensing. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Perspective view of a fragment of the neutral copper(II) layer of <b>1</b> with selected atom numbering. Thermal ellipsoids are drawn at the 30% probability level. (<b>b</b>) Projection of one corrugated oxalato-bridged copper(II) layer of <b>1</b> along the crystallographic <span class="html-italic">a</span> axis (terminal pyrazole ligands are omitted for clarity). The thinner solid lines represent the long Cu–O bond distances. (<b>c</b>) Perspective view of the crystal packing of <b>1</b> along the crystallographic <span class="html-italic">c</span> axis showing the zipper-type interpenetration of two parallel disposed layers. Hydrogen bonds are shown as dashed lines. Symmetry code: (a) = −<span class="html-italic">x</span> + 1, −<span class="html-italic">y</span> + 1, −<span class="html-italic">z</span> + 1; (b) = −<span class="html-italic">x</span> + 1, −<span class="html-italic">y</span>, −<span class="html-italic">z</span> + 1; (c) = −<span class="html-italic">x</span> + 1, <span class="html-italic">y</span> + 1/2, −<span class="html-italic">z</span> + 1/2; (d) = <span class="html-italic">x</span>, <span class="html-italic">y</span> + 1, <span class="html-italic">z</span>; (e) = <span class="html-italic">x</span>, <span class="html-italic">y</span> − 1, <span class="html-italic">z</span>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Projection of a fragment of the crystal packing of <b>1</b> along the crystallographic <span class="html-italic">c</span> axis, showing two interdigitated layers in different colors. N-H⋯Ow, N-H⋯O<sub>ox</sub> and Ow⋯O<sub>ox</sub> type H-bonds are depicted as dashed lines. (<b>b</b>) Projection of the hydrophobic interlayer region [orange box in (a)] along the crystallographic <span class="html-italic">a</span> axis, with the weak inter-pyrazole intralayer N-H⋯N H-bond in evidence. (<b>c</b>) Same as (<b>b</b>), space-filling representation. (<b>d</b>) Detailed side view of a fragment of the interlayer region [sky blue box in (<b>a</b>)] showing the existence of very small hydrophobic cavities arising from the peculiar relative orientation of the three crystallographically independent 4-Hmpz moieties in <b>1</b>. (<b>e</b>) Another zoomed-in view of the interlayer region along the crystallographic <span class="html-italic">c</span> (left) and <span class="html-italic">b</span> (right) axes, evidencing the well-defined position of the water molecules of crystallization; space-filling representation.</p>
Full article ">Figure 3
<p>(<b>a</b>) A view of the three-fold H-bonding motif involving the water molecule of crystallization in <b>1</b>. (<b>b</b>) A view of the H-bonds established between the water molecules of crystallization and each <span class="html-italic">A</span> type chain in <b>1</b>. (<b>c</b>) Projection along the crystallographic <span class="html-italic">a</span> axis of a fragment of one oxalate-copper(II) layer, with <span class="html-italic">A</span> and <span class="html-italic">B</span> type chain fragments shown in different colors for clarity, aiming at illustrating both intra- and inter-chain (intralayer) H-bonding interactions that involve the water molecule of crystallization. H-bonds are depicted as dashed lines.</p>
Full article ">Figure 4
<p>(<b>a</b>) Time profiles for the adsorption of MA (●), DMA (■), and TMA (◆) vapors by <b>1</b> at room temperature. The solid lines are only eye-guides. (<b>b</b>) XRPD of <b>1</b> (red line) and the MA adsorbates <b>2</b> (green line) and <b>3</b> (blue line). The bold black line represents the calculated XRPD of <b>1</b> from the single-crystal X-ray analysis.</p>
Full article ">Figure 5
<p>Sequential snapshots showing the color change of <b>1</b> under MA vapors.</p>
Full article ">Figure 6
<p>Temperature dependence of <span class="html-italic">χ</span><sub>M</sub><span class="html-italic">T</span> (<b>a</b>) and <span class="html-italic">χ</span><sub>M</sub> (<b>b</b>) of <b>1</b> (○) and the MA adsorbates <b>2</b> (☐) and <b>3</b> (◊). The inset in (<b>b</b>) aims to show how the maximum of <span class="html-italic">χ</span><sub>M</sub> of <b>1</b> (○) occured in the vicinity of 300 K. The solid lines are the best-fit curves (see text).</p>
Full article ">Scheme 1
<p>Illustration of the magnetic coupling model for an alternating copper(II) chain showing the relative orientation of the magnetic orbitals centered on each copper(II) ion for <b>1</b> (<b>a</b>), relative to <b>2</b> and <b>3</b> (<b>b</b>). The solid and dashed lines represent short and long metal-ligand bonds, respectively.</p>
Full article ">
12 pages, 3688 KiB  
Review
Vapochromism of Organic Crystals Based on Macrocyclic Compounds and Inclusion Complexes
by Toshikazu Ono and Yoshio Hisaeda
Symmetry 2020, 12(11), 1903; https://doi.org/10.3390/sym12111903 - 19 Nov 2020
Cited by 16 | Viewed by 3903
Abstract
Vapochromic materials, which change color and luminescence when exposed to specific vapors and gases, have attracted considerable attention in recent years owing to their potential applications in a wide range of fields such as chemical sensors and environmental monitors. Although the mechanism of [...] Read more.
Vapochromic materials, which change color and luminescence when exposed to specific vapors and gases, have attracted considerable attention in recent years owing to their potential applications in a wide range of fields such as chemical sensors and environmental monitors. Although the mechanism of vapochromism is still unclear, several studies have elucidated it from the viewpoint of crystal engineering. In this mini-review, we investigate recent advances in the vapochromism of organic crystals. Among them, macrocyclic molecules and inclusion complexes, which have apparent host–guest interactions with analyte molecules (specific vapors and gases), are described. When the host compound is properly designed, its cavity size and symmetry change in response to guest molecules, influencing the optical properties by changing the molecular inclusion and recognition abilities. This information highlights the importance of structure–property relationships resulting from the molecular recognition at the solid–vapor interface. Full article
(This article belongs to the Special Issue Chemical Symmetry Breaking)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Vapochromic behavior of <b>1</b> against various alkane vapors and methanol vapors. Reproduced with permission from Reference [<a href="#B20-symmetry-12-01903" class="html-bibr">20</a>]. Copyright 2017 American Chemical Society.</p>
Full article ">Figure 2
<p>Vapochromic behavior of <b>1</b> against aliphatic aldehydes vapors. Reproduced with permission from Reference [<a href="#B22-symmetry-12-01903" class="html-bibr">22</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>a</b>) Chemical structure of <b>2</b>. (<b>b</b>) Photographs of <b>2</b> by (i) exposing structural liquid <b>2</b> to cyclohexane vapor at 25 °C for 30 min and (ii) heating the solid <b>2</b> at 80 °C under reduced pressure for 30 min. (<b>c</b>) Schematic illustration of the guest vapor-induced state change of <b>2</b>. Reproduced with permission from Reference [<a href="#B23-symmetry-12-01903" class="html-bibr">23</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 4
<p>(<b>a</b>) Chemical structure and crystal structure of <b>3</b>. (<b>b</b>) Chemical structures of ketones (C3–C8). (<b>c</b>) Schematic illustration of activation and vapochromic/vapofluorochromic behaviors of <b>3</b>. Reproduced with permission from Reference [<a href="#B24-symmetry-12-01903" class="html-bibr">24</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) Chemical structures of <b>4</b> and <b>5</b>. (<b>b</b>) Crystal structure of <b>4•5</b>. (<b>c</b>) Pictures of <b>4•5</b> and <b>4•5α</b>. (<b>d</b>) Vapochromic behavior of <b>4•5α</b> against various vapors. Reproduced with permission from Reference [<a href="#B27-symmetry-12-01903" class="html-bibr">27</a>]. Copyright 2016 Wiley-VCH.</p>
Full article ">Figure 6
<p>(<b>a</b>) Chemical structures of <b>6</b>. (<b>b</b>) Vapochromic behavior of <b>6</b> against various vapors.</p>
Full article ">Figure 7
<p>(<b>a</b>) Chemical structure of <b>7</b>. (<b>b</b>) Guest inclusion behavior of <b>7</b> against vapors of guest molecules. (<b>c</b>) Photographs of <b>7</b> and <b>7</b> with exposure of guest vapors under daylight (upper row) and under ultraviolet-light (365 nm, bottom row). Reproduced with permission from Reference [<a href="#B34-symmetry-12-01903" class="html-bibr">34</a>]. Copyright 2016 Wiley-VCH.</p>
Full article ">Figure 8
<p>(<b>a</b>) Chemical structures of <b>8</b>–<b>10</b>. (<b>b</b>) Crystal structure of <b>8</b> including anisole as a guest molecule. (<b>c</b>) Visualization of calculated voids and contact surface maps of <b>8</b>. (<b>d</b>) Vapochromic bahavior of <b>7</b> and <b>8</b>, and <b>7</b> and <b>8</b> after exposing guest vapors for 1 day.</p>
Full article ">Figure 9
<p>(<b>a</b>) Chemical structure of <b>11</b>. (<b>b</b>) Photographs of vaphochromic behavior of <b>11</b> upon uptake/release of water molecules. Reproduced with permission from Reference [<a href="#B40-symmetry-12-01903" class="html-bibr">40</a>]. Copyright 2020 Nature publishing group.</p>
Full article ">Figure 10
<p>Vapochromic behavior of <b>12</b>. Reproduced with permission from Reference [<a href="#B41-symmetry-12-01903" class="html-bibr">41</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 11
<p>(<b>a</b>) Chemical structures of <b>13</b> and <b>14</b>. (<b>b</b>) Crystal structure of acid-base complex (<b>13•14</b>) and its CH<sub>2</sub>Cl<sub>2</sub> included structure. (<b>c</b>) Schematic illustration of emission color tuning by various vapor molecules.</p>
Full article ">
4174 KiB  
Article
Fluorescent Polystyrene Films for the Detection of Volatile Organic Compounds Using the Twisted Intramolecular Charge Transfer Mechanism
by Mirko Borelli, Giuseppe Iasilli, Pierpaolo Minei and Andrea Pucci
Molecules 2017, 22(8), 1306; https://doi.org/10.3390/molecules22081306 - 6 Aug 2017
Cited by 41 | Viewed by 8401
Abstract
Thin films of styrene copolymers containing fluorescent molecular rotors were demonstrated to be strongly sensitive to volatile organic compounds (VOCs). Styrene copolymers of 2-[4-vinyl(1,1′-biphenyl)-4′-yl]-cyanovinyljulolidine (JCBF) were prepared with different P(STY-co-JCBF)(m) compositions (m% = 0.10–1.00) and molecular weights of about 12,000 g/mol. [...] Read more.
Thin films of styrene copolymers containing fluorescent molecular rotors were demonstrated to be strongly sensitive to volatile organic compounds (VOCs). Styrene copolymers of 2-[4-vinyl(1,1′-biphenyl)-4′-yl]-cyanovinyljulolidine (JCBF) were prepared with different P(STY-co-JCBF)(m) compositions (m% = 0.10–1.00) and molecular weights of about 12,000 g/mol. Methanol solutions of JCBF were not emissive due to the formation of the typical twisted intramolecular charge transfer (TICT) state at low viscosity regime, which formation was effectively hampered by adding progressive amounts of glycerol. The sensing performances of the spin-coated copolymer films (thickness of about 4 µm) demonstrated significant vapochromism when exposed to VOCs characterized by high vapour pressure and favourable interaction with the polymer matrix such as THF, CHCl3 and CH2Cl2. The vapochromic response was also reversible and reproducible after successive exposure cycles, whereas the fluorescence variation scaled linearly with VOC concentration, thus suggesting future applications as VOC optical sensors. Full article
Show Figures

Figure 1

Figure 1
<p>UV-Vis absorption spectra of 1 × 10<sup>−5</sup> M JCBF in methanol/glycerol solutions with different glycerol volume contents.</p>
Full article ">Figure 2
<p>Fluorescence (λ<sub>exc.</sub> = 410 nm) of 1 × 10<sup>−5</sup> M JCBF in methanol/glycerol solutions with different glycerol volume contents. In the inset picture, JCBF solutions were excited with a Dark Reader 46B transilluminator (~450 nm, Clare Chemical Research, Dolores, CO, USA).</p>
Full article ">Figure 3
<p>Progressive changes in the fluorescence emission (<span class="html-italic">λ<sub>exc.</sub></span> = 410 nm) of P(STY-co-JCBF)(0.34) film as a function of the exposure to chloroform vapours. The spectra were collected for a total time of 5–6 min.</p>
Full article ">Figure 4
<p>(<b>a</b>) Variation of the fluorescence maximum intensity (<span class="html-italic">λ<sub>exc.</sub></span> = 410 nm) with exposure time to chloroform vapours for all P(STY-co-JCBF)(m) films. Fluorescence was collected for a total time of 10 min in each experiment. (<b>b</b>) Comparison between P(STY-co-JCBF)(0.34) and the most sensitive julolidine-based Fluorescent Molecular Rotors (FMR) dispersed in polystyrene (PS) [<a href="#B28-molecules-22-01306" class="html-bibr">28</a>] towards chloroform vapours.</p>
Full article ">Figure 5
<p>Progressive changes in the fluorescence emission (<span class="html-italic">λ<sub>exc.</sub></span> = 410 nm) of P(STY-co-JCBF)(0.34) film as a function of the exposure to hexane vapours. The spectra were collected for a total time of 5–6 min.</p>
Full article ">Figure 6
<p>Variation of the fluorescence maximum intensity (<span class="html-italic">λ<sub>exc</sub></span><sub>.</sub> = 410 nm) of P(STY-co-JCBF)(0.34) films with exposure time to all VOCs investigated.</p>
Full article ">Figure 7
<p>Fluorescence maximum variation as a function of chloroform concentration (ppm, expressed as vol ratio of vaporized chloroform in the measuring chamber).</p>
Full article ">Scheme 1
<p>Synthesis of the 2-[4-vinyl(1,1′-biphenyl)-4′-yl]-cyanovinyljulolidine (JCBF) monomer. DMF: dimethylformamide; rt: room temperature; THF: tetrahydrofurane; FJUL: 9-formyljulolidine; CBF: of 2-[4-vinyl(1,1′-biphenyl)-4′-yl]-acetonitrile.</p>
Full article ">Scheme 2
<p>Synthesis of the P(STY-co-JCBF) copolymers. AIBN: Azobisisobutyronitrile. dd: days.</p>
Full article ">
960 KiB  
Article
Spin-Crossover Behavior of Hofmann-Type-Like Complex Fe(4,4’-bipyridine)Ni(CN)4·nH2O Depending on Guest Species
by Kazumasa Hosoya, Shin-ichi Nishikiori, Masashi Takahashi and Takafumi Kitazawa
Magnetochemistry 2016, 2(1), 8; https://doi.org/10.3390/magnetochemistry2010008 - 16 Feb 2016
Cited by 22 | Viewed by 8302
Abstract
A newly prepared metal complex Fe(4,4’-bipyridine)Ni(CN)4·nH2O, which was estimated to have a structure similar to the Hofmann-type clathrate host, changed its color from orange to deep orange and yellow on exposure to ethanol and acetone vapor, respectively, [...] Read more.
A newly prepared metal complex Fe(4,4’-bipyridine)Ni(CN)4·nH2O, which was estimated to have a structure similar to the Hofmann-type clathrate host, changed its color from orange to deep orange and yellow on exposure to ethanol and acetone vapor, respectively, and the respective samples showed thermally induced two-step and one-step spin transitions. Full article
(This article belongs to the Special Issue Spin Crossover (SCO) Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Crystal structure of Fe(py)<sub>2</sub>Ni(CN)<sub>4</sub> [<a href="#B25-magnetochemistry-02-00008" class="html-bibr">25</a>]; (<b>b</b>) the network structure of the Hofmann-type host complex M(ligand)<sub>2</sub>Ni(CN)<sub>4</sub>; and (<b>c</b>) a structure estimated for Fe(bpy)Ni(CN)<sub>4</sub>∙<span class="html-italic">n</span>H<sub>2</sub>O(<span class="html-italic">n</span> = 2.5)(<b>1</b>). (orange: Fe<sup>II</sup>; green: Ni<sup>II</sup>, gray: C; blue: N; H atoms are omitted for clarity.)</p>
Full article ">Figure 2
<p>(<b>a</b>) Fe K-edge and (<b>b</b>) Ni K-edge Extended X-ray Absorption Fine Structure (EXAFS) spectra of <b>1</b> and Fe(py)<sub>2</sub>Ni(CN)<sub>4</sub>.</p>
Full article ">Figure 3
<p>Diffuse reflectance spectra of <b>1</b>, <b>2</b> and <b>3</b>.</p>
Full article ">Figure 4
<p>χ<sub>M</sub><span class="html-italic">T vs.</span> <span class="html-italic">T</span> plots (<b>a</b>) for <b>1</b>, <b>2</b>, and <b>3</b>; (<b>b</b>) for <b>2</b> containing C<sub>2</sub>H<sub>5</sub>OH and C<sub>2</sub>D<sub>5</sub>OD; and (<b>c</b>) <b>3</b> containing (CH<sub>3</sub>)<sub>2</sub>CO and (CD<sub>3</sub>)<sub>2</sub>CO.</p>
Full article ">Figure 5
<p>Mössbauer spectra of <b>1</b>, <b>2</b> and <b>3</b> at 298 K and 77 K: (<b>a</b>,<b>b</b>) <b>1</b> at 298 K and 77 K; (<b>c</b>,<b>d</b>) <b>2</b> at 298 K and 77 K; and (<b>e</b>,<b>f</b>): <b>3</b> at 298 K and 77 K. The red and the blue line show the major and the minor component, respectively. The black line shows the sum of both components.</p>
Full article ">
495 KiB  
Article
A Novel Porphyrin-Containing Polyimide Nanofibrous Membrane for Colorimetric and Fluorometric Detection of Pyridine Vapor
by Yuanyuan Lv, Yani Zhang, Yanglong Du, Jiayao Xu and Junbo Wang
Sensors 2013, 13(11), 15758-15769; https://doi.org/10.3390/s131115758 - 19 Nov 2013
Cited by 26 | Viewed by 7283
Abstract
A novel zinc porphyrin-containing polyimide (ZPCPI) nanofibrous membrane for rapid and reversible detection of trace amounts of pyridine vapor is described. The membrane displays a distinct color change, as well as dramatic variations in absorption and fluorescent emission spectra, upon exposure to pyridine [...] Read more.
A novel zinc porphyrin-containing polyimide (ZPCPI) nanofibrous membrane for rapid and reversible detection of trace amounts of pyridine vapor is described. The membrane displays a distinct color change, as well as dramatic variations in absorption and fluorescent emission spectra, upon exposure to pyridine vapor. This condition allows the detection of the analyte at concentrations as low as 0.041 ppm. The vapochromic and spectrophotometric responses of the membrane are attributed to the formation of the ZPCPI-pyridine complex upon axial coordination. From surface plasmon resonance analysis, the affinity constant of ZPCPI-pyridine complex was calculated to be (3.98 ± 0.25) × 104 L·mol−1. The ZPCPI nanofibrous membrane also showed excellent selectivity for pyridine vapor over other common amines, confirming its applicability in the manufacture of pyridine-sensitive gas sensors. Full article
(This article belongs to the Section Chemical Sensors)
Show Figures


<p>Molecular structure of ZPCPI used in this study and the schematic illustration for the whole vapochromic and spectrophotometric detection of pyridine vapor.</p>
Full article ">
<p><sup>1</sup>H-NMR spectrum of ZPPAA in DMSO-<span class="html-italic">d</span><sub>6</sub>. (*) indicates solvent impurity.</p>
Full article ">
<p>FESEM micrograph (<b>A</b>) and CLSM image (<b>B</b>) (×40) of the ZPCPI nanofibrous membrane (<span class="html-italic">λ<sub>ex</sub></span> = 488 nm); FESEM micrograph (<b>C</b>) of the ZPCPI nanofibrous membrane after consecutively used for five times.</p>
Full article ">
<p>Absorption (<b>A</b>) and fluorescence emission (<b>B</b>) spectra of the ZPCPI nanofibrous membrane.</p>
Full article ">
<p>Photographs of the ZPCPI nanofibrous membrane when exposed to different vapors (30 ppm). (<b>A</b>) Blank; (<b>B</b>) Pyridine; (<b>C</b>) DEA; (<b>D</b>) TEA; (<b>E</b>) Py; (<b>F</b>) CA; (<b>G</b>) CO<sub>2</sub>; (<b>H</b>) H<sub>2</sub>O; and (<b>I</b>) recovered when puffed with N<sub>2</sub>.</p>
Full article ">
<p>Absorption spectra of the ZPCPI nanofibrous membrane when exposed to different concentrations of pyridine vapor or recovered when puffed with N<sub>2</sub>.</p>
Full article ">
<p>Fluorescence emission spectra of the ZPCPI nanofibrous membrane when exposed to different concentrations of pyridine vapor or recovered when puffed with N<sub>2</sub> (<span class="html-italic">λ<sub>ex</sub></span> = 420 nm).</p>
Full article ">
<p>Absorption spectra of the ZPCPI nanofibrous membrane exposed to different vapors (<span class="html-italic">λ<sub>ex</sub></span> = 420 nm).</p>
Full article ">
<p>SPR response induced by the ZPCPI nanofibrous membrane deposited on an SPR chip when exposed to pyridine solution with different concentrations. H<sub>2</sub>O is the buffer solution throughout the whole process.</p>
Full article ">
1356 KiB  
Article
Vapochromic Behaviour of M[Au(CN)2]2-Based Coordination Polymers (M = Co, Ni)
by Julie Lefebvre, Jasmine L. Korčok, Michael J. Katz and Daniel B. Leznoff
Sensors 2012, 12(3), 3669-3692; https://doi.org/10.3390/s120303669 - 16 Mar 2012
Cited by 30 | Viewed by 10710
Abstract
A series of M[Au(CN)2]2(analyte)x coordination polymers (M = Co, Ni; analyte = dimethylsulfoxide (DMSO), N,N-dimethylformamide (DMF), pyridine; x = 2 or 4) was prepared and characterized. Addition of analyte vapours to solid M(μ-OH2)[Au(CN)2]2 [...] Read more.
A series of M[Au(CN)2]2(analyte)x coordination polymers (M = Co, Ni; analyte = dimethylsulfoxide (DMSO), N,N-dimethylformamide (DMF), pyridine; x = 2 or 4) was prepared and characterized. Addition of analyte vapours to solid M(μ-OH2)[Au(CN)2]2 yielded visible vapochromic responses for M = Co but not M = Ni; the IR νCN spectral region changed in every case. A single crystal structure of Zn[Au(CN)2]2(DMSO)2 revealed a corrugated 2-D layer structure with cis-DMSO units. Reacting a Ni(II) salt and K[Au(CN)2] in DMSO yielded the isostructural Ni[Au(CN)2]2(DMSO)2 product. Co[Au(CN)2]2(DMSO)2 and M[Au(CN)2]2(DMF)2 (M = Co, Ni) complexes have flat 2-D square-grid layer structures with trans-bound DMSO or DMF units; they are formed via vapour absorption by solid M(μ-OH2)[Au(CN)2]2 and from DMSO or DMF solution synthesis. Co[Au(CN)2]2(pyridine)4 is generated via vapour absorption by Co(μ-OH2)[Au(CN)2]2; the analogous Ni complex is synthesized by immersion of Ni(μ-OH2)[Au(CN)2]2 in 4% aqueous pyridine. Similar immersion of Co(μ-OH2)[Au(CN)2]2 yielded Co[Au(CN)2]2(pyridine)2, which has a flat 2-D square-grid structure with trans-pyridine units. Absorption of pyridine vapour by solid Ni(μ-OH2)[Au(CN)2]2 was incomplete, generating a mixture of pyridine-bound complexes. Analyte-free Co[Au(CN)2]2 was prepared by dehydration of Co(μ-OH2)[Au(CN)2]2 at 145 °C; it has a 3-D diamondoid-type structure and absorbs DMSO, DMF and pyridine to give the same materials as by vapour absorption from the hydrate. Full article
(This article belongs to the Special Issue Molecular Devices and Machines: Cooperativity and Multifunctionality)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>Powders of Co[Au(CN)<sub>2</sub>]<sub>2</sub> (<b>top left</b>), Co(μ-OH<sub>2</sub>)<sub>2</sub>[Au(CN)<sub>2</sub>]<sub>2</sub> (<b>top middle</b>), Co[Au(CN)<sub>2</sub>]<sub>2</sub>(pyridine)<sub>4</sub> (<b>top right</b>), Co[Au(CN)<sub>2</sub>]<sub>2</sub>(DMF)<sub>2</sub> (<b>bottom left</b>) and Co[Au(CN)<sub>2</sub>]<sub>2</sub>(DMSO)<sub>2</sub> (<b>bottom right</b>). The latter three were synthesized by vapour absorption by solid Co[Au(CN)<sub>2</sub>]<sub>2</sub>.</p>
Full article ">
<p>Extended 2-D structure of Zn[Au(CN)<sub>2</sub>]<sub>2</sub>(DMSO)<sub>2</sub>. (<b>a</b>) Local geometry of Zn, showing thermal ellipsoids; (<b>b</b>) A single 2-D sheet, viewed down the face of the sheet; (<b>c</b>) A pair of 2-D sheets, viewed perpendicular to the sheet face, long Au(1)-Au(2*) interactions 3.4943(6) Å represents the closest contact between sheets (DMSO molecules excluded for clarity).</p>
Full article ">
<p>Comparison between the powder X-ray diffractograms predicted for Zn[Au(CN)<sub>2</sub>]<sub>2</sub>(DMSO)<sub>2</sub> (Zn, green) and the blue Cu[Au(CN)<sub>2</sub>]<sub>2</sub>(DMSO)<sub>2</sub> polymorph (Cu, blue), with the diffractogram obtained experimentally for Ni[Au(CN)<sub>2</sub>]<sub>2</sub>(DMSO)<sub>2</sub> (Ni, orange), prepared via solution methods.</p>
Full article ">
<p>Comparison between the powder X-ray diffractogram determined experimentally for the Co[Au(CN)<sub>2</sub>]<sub>2</sub>(DMSO)<sub>2</sub> (Co, purple) and the diffractogram predicted for Mn[Au(CN)<sub>2</sub>]<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub> from the single crystal structure (Mn, blue) [<a href="#b54-sensors-12-03669" class="html-bibr">54</a>].</p>
Full article ">
<p>Comparison between the powder X-ray diffractogram (generated from single-crystal data) for Co[Au(CN)<sub>2</sub>]<sub>2</sub>(DMF)<sub>2</sub> (Co, purple) [<a href="#b50-sensors-12-03669" class="html-bibr">50</a>], the experimental diffractogram for Ni[Au(CN)<sub>2</sub>]<sub>2</sub>(DMF)<sub>2</sub> (Ni, orange) and the diffractogram predicted by the proposed structural model (Ni*, black).</p>
Full article ">
<p>Structural model proposed for Ni[Au(CN)<sub>2</sub>]<sub>2</sub>(DMF)<sub>2</sub>: (<b>A</b>). A 2-D square grid array with DMF molecules coordinated on both sides of the Ni[Au(CN)<sub>2</sub>]<sub>2</sub> layer; (<b>B</b>) and (<b>C</b>). Side-view of a layer showing the relative position of the [Au(CN)<sub>2</sub>]<sup>−</sup> units with respect to the Ni(II) centres.</p>
Full article ">
<p>Comparison between the powder X-ray diffractogram (generated from single-crystal data) for Cu[Au(CN)<sub>2</sub>]<sub>2</sub>(pyridine)<sub>2</sub> (Cu, orange) [<a href="#b35-sensors-12-03669" class="html-bibr">35</a>], the experimental diffractogram of Co[Au(CN)<sub>2</sub>]<sub>2</sub>(pyridine)<sub>2</sub> (Co, purple), and the diffractogram predicted by its structural model (Co*, black).</p>
Full article ">
<p>Structural model proposed for Co[Au(CN)<sub>2</sub>]<sub>2</sub>(pyridine)<sub>2</sub>: (<b>A</b>) 2-D square-grid array with pyridine molecules on both sides of the grid; (<b>B</b>) Side-view of a 2-D layer, showing the position of the [Au(CN)<sub>2</sub>]<sup>−</sup> units with respect to the Co(II) centres.</p>
Full article ">
<p>Extended 3-D structure of Co[Au(CN)<sub>2</sub>]<sub>2</sub>, looking down the <span class="html-italic">c</span> axis. Dashed lines represent Au-Au interactions.</p>
Full article ">
Back to TopTop