scholarly journals Quantifying Propiconazole in Wood by Raman Microscopy

2021 ◽  
Author(s):  
Erlet Kurti

Raman spectra of wood blocks treated with different propiconazole solutions (4%, 2% and 1%) in mineral spirits were recorded using a Raman microscope, equipped with a NIR (785nm) laser. The strong propiconazole Raman band, in the uncongested region 647-693 cm-1 was chosen as the analytical band. The normalized intensity of analytical band was used to determine the propiconazole distribution in white spruce. Mapping measurements on radial face of the treated samples revealed that on average the propiconazole concentration in summerwood was much higher than concentration in springwood. CG-MS analyses were ccarriedout on methanol extractions of soaked samples milled at ~1.5 mm intervals. The depth profiles in longitudinal directions, obtained by Raman and GC-MS measurements, suggested that propiconazole tended to bloom to the surface during drying. A linear calibration plot was produced from averaged Raman normalized intensity and GC-MS-measured concentrations. By using the regression line, concentration in the longitudinal direction was predicted for another wood block soaked in 3% propiconazole solution.

2021 ◽  
Author(s):  
Erlet Kurti

Raman spectra of wood blocks treated with different propiconazole solutions (4%, 2% and 1%) in mineral spirits were recorded using a Raman microscope, equipped with a NIR (785nm) laser. The strong propiconazole Raman band, in the uncongested region 647-693 cm-1 was chosen as the analytical band. The normalized intensity of analytical band was used to determine the propiconazole distribution in white spruce. Mapping measurements on radial face of the treated samples revealed that on average the propiconazole concentration in summerwood was much higher than concentration in springwood. CG-MS analyses were ccarriedout on methanol extractions of soaked samples milled at ~1.5 mm intervals. The depth profiles in longitudinal directions, obtained by Raman and GC-MS measurements, suggested that propiconazole tended to bloom to the surface during drying. A linear calibration plot was produced from averaged Raman normalized intensity and GC-MS-measured concentrations. By using the regression line, concentration in the longitudinal direction was predicted for another wood block soaked in 3% propiconazole solution.


2001 ◽  
Vol 15 (28n30) ◽  
pp. 3865-3868 ◽  
Author(s):  
H. MIYAOKA ◽  
T. KUZE ◽  
H. SANO ◽  
H. MORI ◽  
G. MIZUTANI ◽  
...  

We have obtained the Raman spectra of TiCl n (n= 2, 3, and 4). Assignments of the observed Raman bands were made by a normal mode analysis. The force constants were determined from the observed Raman band frequencies. We have found that the Ti-Cl stretching force constant increases as the oxidation number of the Ti species increases.


2020 ◽  
Author(s):  
Ioan Bratu ◽  
Constantin Marutoiu ◽  
Antonio Hernanz ◽  
Olivia Florena Nemes ◽  
M Tiplic ◽  
...  

Abstract Three Transylvanian fragments of some 18th century Saxon tiles obtained from excavations in the Central Railway Station Square in Sibiu (Romania) have been studied by Raman microscopy, FTIR, SEM-EDX and XRF techniques. A lead-rich aluminosilicate was found to have been used as enamel/glaze for the tile fragments. An analysis of the band components of the Raman spectra of the enamel of one fragment in the 400–1200 cm− 1 spectral region indicates that their processing temperature was in the 600–800 °C range, in good agreement with the temperature obtained from FTIR spectra of the other two tile fragments.


2019 ◽  
Vol 73 (12) ◽  
pp. 1409-1419
Author(s):  
Ardian B. Gojani ◽  
Dávid J. Palásti ◽  
Andrea Paul ◽  
Gábor Galbács ◽  
Igor B. Gornushkin

Spatial heterodyne spectroscopy (SHS) is used for quantitative analysis and classification of liquid samples. SHS is a version of a Michelson interferometer with no moving parts and with diffraction gratings in place of mirrors. The instrument converts frequency-resolved information into a spatially resolved one and records it in the form of interferograms. The back-extraction of spectral information is done by the fast Fourier transform. A SHS instrument is constructed with the resolving power 5000 and spectral range 522–593 nm. Two original technical solutions are used as compared to previous SHS instruments: the use of a high-frequency diode-pumped solid-state laser for excitation of Raman spectra and a microscope-based collection system. Raman spectra are excited at 532 nm at the repetition rate 80 kHz. Raman shifts between 330 cm−1 and 1600 cm−1 are measured. A new application of SHS is demonstrated: for the first time, it is used for quantitative Raman analysis to determine concentrations of cyclohexane in isopropanol and glycerol in water. Two calibration strategies are employed: univariate based on the construction of a calibration plot and multivariate based on partial least squares regression. The detection limits for both cyclohexane in isopropanol and glycerol in water are at a 0.5 mass% level. In addition to the Raman–SHS chemical analysis, classification of industrial oils (biodiesel, poly(1-decene), gasoline, heavy oil IFO380, polybutenes, and lubricant) is performed using the Raman–fluorescence spectra of the oils and principal component analysis. The oils are easily discriminated showing distinct non-overlapping patterns in the principal component space.


1982 ◽  
Vol 65 (1) ◽  
pp. 172-177
Author(s):  
John H O'Keefe ◽  
Leo F Sharry

Abstract A wide-line proton magnetic resonance technique (PMR), in which hydrogen response is used as the measure of wax in dried greasy wool, is a viable and preferred alternative to the commonly used Soxhlet procedure in which extracted weight is measured. The validity of the PMR method depends on wax hydrogen composition (inter- and intra-breed) being constant, and on background contribution from non-wax hydrogen (in suint, wool fiber, and vegetable matter) being small and constant. The experimentation reported in this paper shows that both requirements are met. The PMR method for wax is precise (better than 0.1% at mid-range); rapid (200 samples per 8 h day); sensitive (better than 15 mg); and yields a linear calibration graph over the entire range (0-100%) for a 5 g sample. A comparison of the methods shows that the linear regression line for wax by Soxhlet (y) against wax by PMR (x) is given by y = 1.0257 (SE 0.2277) + 0.9894 (SE 0.0142)x, with r2 = 0.99, and n = 20.


Materials ◽  
2020 ◽  
Vol 13 (23) ◽  
pp. 5453
Author(s):  
Min Wang ◽  
Changhao Wang ◽  
Jian Wang ◽  
Liming Lu ◽  
Xiaoye Gong ◽  
...  

In situ high-temperature Raman spectra of polycrystalline KBi(MoO4)2 were recorded from room temperature to 1073 K. Thermal stability of the monoclinic KBi(MoO4)2 was examined by temperature-dependent XRD. The monoclinic phase transformed into the scheelite tetragonal structure at 833 K, and then to the monoclinic phase at 773 K. Quantum chemistry ab initio calculation was performed to simulate the Raman spectra of the structure of KBi(MoO4)2 high-temperature melt. The experimental Raman band at 1023 K was deconvoluted into seven Gaussian peaks, and the calculated results were in good agreement with the experimental data. Therefore, the vibrational modes of Raman peaks of molten KBi(MoO4)2 were assigned. It was confirmed that the isolated structure of [Bi(MoO4)2]− monomer, consisting of Mo6+ centers and Bi3+ sub-centers connected by edge-sharing, mainly exists in the melt of KBi(MoO4)2.


2000 ◽  
Vol 15 (3) ◽  
pp. 583-585 ◽  
Author(s):  
G. Chen ◽  
W. Hao ◽  
Y. Shi ◽  
Y. Wu ◽  
S. Perkowitz

For the first time, we measured Raman spectra from Li(Al1-xCox)O2 (x = 0.5 to 0.9), a new cathode material for lithium batteries. Whereas LiCoO2 sintered at 400 °C develops a spinel structure, Li(Al1-xCox)O2 sintered at 380 °C is amorphous, as shown by its single broad Raman band. Li(Al1-xCox)O2 sintered at 700 or 900 °C shows Raman peaks independent of x that coincide with those from LiCoO2, indicating that Li(Al1-xCox)O2 has the α–NaFeO2 structure (space group R3m). Traces of the impurity phase Co3O4 appear in samples treated at 900 °C but not at 700 °C. The Raman peak widths exceed those in LiCoO2, suggesting that replacement of Co by Al increases disorder among the Li ions.


1980 ◽  
Vol 58 (12) ◽  
pp. 1220-1228 ◽  
Author(s):  
Mohsen Jaber ◽  
François Bertin ◽  
Germaine Thomas-David

A study by Raman spectrometry of the System Fe3+–H2C2O4 in aqueous solution shows very clearly the successive formation of the three complexes: Fe(C2O4)+, Fe(C2O4)2−, Fe(C2O4)33−. Each species is characterised by a specific Raman spectrum.The evolution of Raman band intensities with the range of exciting line, and the comparison between ir and Raman spectra of Fe(III) and Al(III) complexes, show for Fe(III) species a pre-resonance Raman effect.


2020 ◽  
Author(s):  
Ana Magalhaes ◽  
Pola Goldberg Oppenheimer ◽  
Tim Overton ◽  
Kevin Wright

<p>Biofilm development in industrial settings can prove costly to manufacturing and consumer health. The presence of contaminants in raw materials, finished products and on process contact surfaces negatively impacts on product quality and safety. Rapid and accurate identification of spoilage and pathogenic microorganisms is crucial to implement effective biofilm control strategies that enhance product safety. The application of confocal Raman microscopy (CRM) for non-invasive and rapid characterisation of clinical and food isolates has been reported. The question remains whether the technique can be used as an online monitoring tool for real-time measurement of biofilm build-up in dynamic manufacturing conditions.</p> <p>In this study we investigated if CRM could be used in the manufacturing environment as an alternative microbiological quality control method. We assessed if this technology is able to differentiate between bacterial species and their growth phenotype, as well as detect contaminants from process samples.</p> <p>Laboratory and industrial isolates grown under different culture conditions (planktonic, agar plates, and CDC grown biofilms), and formulated products were analysed using a confocal Raman microscope (Thermofisher DXR2xi) under optimised settings. Reference and experimental Raman spectra were collected and analysed for all test conditions [1]. Spectral similarities were evaluated by developing a microbial multivariate predictive model using a two-way orthogonal partial least squares (O2PLS) regression for cluster analysis [2].</p> <p>Optimal spectra for microbes were obtained in the fingerprint region at approximately 600 - 1800 cm<sup>-1 </sup>where characteristic peaks could be assigned to different biological macromolecules (nucleic acids, proteins, lipids and carbohydrates). Cluster analysis showed good group separation with low variation within but high variation between bacterial strains, enabling bacterial differentiation. It also highlighted the variations observed in the spectral fingerprint for planktonic, agar and biofilm growth modes. Comparative studies suggest that the peak associated with nucleotide ring stretching (~ 700 cm<sup>-1</sup>) could be used as a microbial marker for contamination in formulation.</p> <p>Our findings indicate that confocal Raman microscopy can be used for at-line monitoring of contamination in product streams. Raman spectra provide biochemical data for microbial characterisation but variations in the spectra are often difficult to observe and interpret. Multivariate statistical methods permit rapid interrogation of spectral data, with the potential to improve microbial identification. In combination with multivariate analysis, CRM can be used as an analytical tool for rapid identification of industrial isolates and differentiation of their growth phenotype.</p> <p> </p> <p><strong>References</strong></p> <p>[1] Beier, B.D., R.G. Quivey, and A.J. Berger,Raman microspectroscopy for species identification and mapping within bacterial biofilms.AMB Express, 2012. <strong>2</strong>(1): p. 35-35.</p> <p>[2] Zou, X., et al., Automatic Spectroscopic Data Categorization by Clustering Analysis (ASCLAN): A Data-Driven Approach for Distinguishing Discriminatory Metabolites for Phenotypic Subclasses.Analytical Chemistry, 2016. <strong>88</strong>(11): p. 5670-5679.</p>


Sign in / Sign up

Export Citation Format

Share Document