NIR spectra simulation of thymol for better

14 downloads 0 Views 1MB Size Report
The Merck Index - An Encyclopedia of Chemicals, Drugs, and Biologicals. ... Model for Structural Parameter and Magnetic Tensors of Large Free Radicals, ...
Missing:
1 2 3 4

NIR spectra simulation of thymol for better understanding of the spectra forming factors, phase and concentration effects and PLS regression features

5

Krzysztof B. Beć,1* Justyna Grabska,1,2 Christian G. Kirchler1 and Christian W. Huck1

6 7 8 9 10

1

11

2

Institute of Analytical Chemistry and Radiochemistry, Leopold-Franzens University, Innrain 80/82, CCB-Center for Chemistry and Biomedicine, 6020-Innsbruck, Austria Faculty of Chemistry, University of Wrocław, F. Joliot-Curie 14, 50-383 Wrocław, Poland

12 13 14 15 16 17 18 19 20 21 22 23 24 25 26

*Corresponding Author: K.B. Bed, email address: [email protected]

27

1

28

Abstract

29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

Near-infrared spectroscopy (NIRS) is an effective analytical/quality control tool in various appliances, i.e. in phytopharmaceutical industry. While multivariate analysis gives NIRS the desired analytical performance level it lacks in providing deeper physical insights. Quantum mechanical (QM) simulation of NIR spectra is becoming feasible for complex molecules in the present days. The usefulness of QM simulation in NIRS surpasses that of IR region as NIR spectra are intrinsically complex due to anharmonic effects; yet QM investigations in NIR remain rare due to practical limitations. QM provides detailed band assignments, far exceeding the capability of classical spectroscopic methods, but also offers unveiling the influential factors in the chemometric models. In the present work we investigate NIR spectrum of thymol, a constituent of Thymi herba used in traditional medicine. Experimental part compares the NIR spectra of thymol in solid state, melted (neat liquid) and in CCl4 solution of different concentration. The analysis unveils the spectral regions of strongly varying sensitivity level to the sample state; two regions (6000-5600 cm-1 and 4490-4000 cm-1 ) are identified to remain largely invariant to the sample state. The band assignments and plotted density of modal contribution visualizes generalized and categorized factors which form the NIR spectrum of thymol. A principle relationship with the regression coefficients plot is established, demonstrating that the NIR bands which retain high insensitiveness to the intermolecular interactions give the most influence in the PLS regression model of thymol.

46 47 48 49

Keywords: NIRS, spectra simulation, thymol, 5-Methyl-2-(propan-2-yl)phenol, phase transition, Thymi herba, phytopharmaceuticals, overtones, combination bands.

50

2

51 52

1. Introduction

53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68

Near-infrared spectroscopy (NIRS; 10,000-4000 cm-1) grows in importance as state-of-the-art quantitative analytical method due to a number of considerable advantages.1,2,3 It offers non-invasive capability, short analysis time, high throughput and high sample volume, is applicable to a wide range of sample types and retains a relative insensitiveness to their homogeneity levels while maintaining low operational cost.4,5 High compatibility with fiber probe instrumentation and recent breakthrough in developing miniaturized spectrometers largely boosted its final worthiness.5,6,7 Consequently, NIRS is an effective analytical/quality control tool in various appliances, i.e. in phytopharmaceutical industry. At the same time a complex nature of NIR spectra, resulting from a multitude of overlaying overtone and combination bands ruled by anharmonic effects decreases the inherent selectivity of the method.1-3 Extensive calibration and chemometric methods are needed; however, in the final outcome the sum of benefits typically outweighs the deficiencies of NIRS. Nevertheless, the difficulty in the interpretation of the spectral features forms a hindrance, putting NIRS apart from IR (mid-IR; 4000-400 cm-1) spectroscopy which coincidently features typically less complex spectra and wellestablished computational approaches which yield simulated IR spectra in a straightforward way. 1 Oppositely, truly useful simulation methods for NIRS should be described as still developing in the present time.1,8

69 70 71 72 73 74 75 76 77 78

Unlike IR spectroscopy which may rely on harmonic approximation with good results,9,10 quantum mechanical (QM) simulation of NIR spectra necessarily requires anharmonic methods, in which the nature of molecular vibration (i.e. mode-mode correlations) is sufficiently described. 1,11 An overview of the theoretical background in the context of above factors may be found in recent literature;1,8,11 from the point of present introduction an essential notice is that the computational expense has been a practically limiting factor in theoretical NIRS for years. Consequently, spectral investigations utilizing QM simulations in NIR remain very rare unlike those in IR region, which contrarily can be nowadays described as routine. Substantial inherent complexity of NIR spectra makes QM simulations extremely useful, as classic spectral analysis often struggles to provide adequate explanation of the observed features.12

79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96

NIRS rapidly gains favor in the analytical applications to raw materials among which pharmaceuticals of plant origin are frequently focused analytes. 13,14 The amount of feasibility studies exploring various conditions and methods in NIR quantitative analysis of medicinal herbs has been steadily increasing in the last few years. 13,14,15,16 Coherent with these trends an evaluation of quality control performances of Thymi herba based on NIR and ATR-IR (attenuated total reflection IR) spectroscopy hyphenated with multi-variate analytical (MVA) procedures has been reported recently.17 Thymi herba is a medicinal herb featuring a variety of therapeutic properties, mostly accounted with the antioxidant capacity of phenolic compounds embodied within. The study by Pezzei et al. focused on the quantification of two key chemicals involved, rosmarinic acid and thymol; these have been forming a typical area of interest of classical analytical studies in the past highlighting their importance for the medicinal properties of the drug.18,19 The former compound has already been investigated by us using methods of spectra simulation. 20 The latter one, thymol (IUPAC: 5-Methyl-2-(propan-2-yl)phenol; Fig. 1), is an antiseptic21 (antifungal22 and antibacterial23 ) substance figuring in the European Pharmacopeia;24 a number of other therapeutic effects (antispasmodic, 25 anti-inflammatory,26 immunomodulation,27 anti-oxidant28) has been reported as well. An essential oil of Thymi herba contains 30-70% of thymol. It is a traditional natural drug, although it can be found as the active compound of numerous industrially prepared cough medicines as well.29 Thymol forms an interesting subject of study from vibrational point of view as well. While

3

97 98 99 100 101 102

being a close derivative of phenol which has been focused in our recent study,30 the spectrum of thymol is substantially enriched with NIR features. Following these features will allow drawing generalized conclusions about the factors standing behind formation of fine effects observable in NIR spectra of close derivatives. These correlations offer shedding more light onto the relationship between NIR spectral outlines and details of chemometric calibration models in the quantification procedures of thymol. 17

103

104 105 106

Fig. 1. The optimized (DFT-B3LYP/SNST) molecular structure of thymol (refer to Supplementary Material for detailed structural parameters of the molecule).

107 108 109 110 111 112 113 114 115 116 117

118 119 120 121 122 123 124 125 126 127 128 129

The present study aims at exploring NIR spectral features of thymol and their dependence on the sample state and concentration. Solid state and melted (neat liquid) as well as diluted (CCl4 solution) thymol will be investigated. The wavenumber regions featuring various level of sensitivity to the sample property will be observed. QM simulations of NIR bands will provide detailed band assignments throughout the entire spectral region. An analysis of mode contributions in the form of density maps will be performed. Such approach grants straightforward visualization of generalized and categorized contributions of vibrational modes, highlighting the major spectra forming factors. These analyses will be confronted with the most influential determinants in the chemometric calibration models allowing to draw conclusions about the relationship between these two fundamental for NIRS aspects.

2. Materials and methods 2.1 NIR spectral measurements Thymol was purchased from Sigma (T-0501; ≥99.5 % purity) and used without further purification. The measurement of NIR spectra of solid state and melted analyte was performed on Büchi NIRFlex N-500 FT‑NIR benchtop spectrometer controlled by the manufacturer’s NIRWare 1.4.3010 software. The solid sample was measured using Büchi solids device with a rotating cylinder and operating in diffuse reflectance mode and at room temperature. The spectrum of the melted analyte was acquired at the temperature of 60˚C (333 K), stabilized by the built-in sample temperature control unit of the spectrometer and using liquid cell accessory. Sealed quartz cuvette of 1 mm pathlength was employed for the melted sample. The following recording parameters were selected for solid and melted samples; spectral resolution, spectral range and scan number were 8 cm-1 (resulting in the interpolated 4 cm -1 of data spacing and 2 cm-1 of absolute accuracy), 10,000-

4

130 131

4000 cm-1 and 32-64 (melted-solid), respectively. Spectra recordings were triplicated for each sample.

132 133 134 135 136 137 138 139 140 141 142 143 144 145

Thymol in solution was prepared using CCl4 (carbon tetrachloride) a relatively inert solvent with negligible absorption throughout the entire NIR region. Varying concentration of the analyte (10-100 mg ml-1; 0.067-0.67 mol dm-3 ) was examined to account for the eventuality of concentration effect due to intermolecular interactions. These samples were measured using the PerkinElmer Spectrum 400 FT-IR/FT-NIR spectrometer operating in transmission mode. Sealed quartz cuvette with 2 mm optical pathlength was utilized. The following measurement parameters were set; 2 cm-1 (interpolated 0.5 cm-1 data spacing), 10,000-2000 cm-1 and 16, as the spectral resolution, spectral range and scan number, respectively. Measurements were repeated three times for each sample. The spectrometer allowed extending the investigation onto the wavenumber region of 4000-3650 cm-1 , which is populated by multiple combination bands and thus by its nature remains vibrationally indistinguishable from the proper NIR region (by concept down limited to 4000 cm-1, typically).1 The spectrum of diluted (10 mg ml-1) sample was processed with baseline correction; other spectra in this work presented adequate quality for qualitative assessment with no need for preprocessing of any kind.

146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175

2.2 Quantum mechanical simulation of the NIR spectrum Since a fully anharmonic vibrational analysis induces a considerable computational cost on itself,31 it is crucial to employ a cost-effective method for the determination of the molecule’s ground state electronic structure, a preliminary step for obtaining all further properties of the system. In this role density functional theory (DFT) offers considerable advantages. In this work B3LYP (Becke, 3parameter, Lee-Yang-Parr) single-hybrid functional32 was selected, as it has repeatedly been reported to work well in anharmonic vibrational analysis.8,11,20,30 SNST basis set of triple- quality from SNS family33 was selected; it has been developed with spectroscopic appliances in mind and yields a good accuracy/cost balance for such purpose.8,31 Calculation of anharmonic frequencies is relatively sensitive to the accuracy of the approximation of the ground state structure.8 In order to obtain a desired precision, the molecular geometry optimization procedure was completed with tight convergence criteria. The resulting ground state geometry of thymol molecule which was further used in the calculations of the vibrational levels is presented in Fig. 1 and Table S1 (Supplementary Material). To increase the description quality of long-range interactions an empirical correction for dispersion was applied; in this role Grimme’s D3 damping function (GD3) 34 model was selected. Since the study focuses on solution phase a solvation model within self-consistent reaction field (SCRF) formalism35 was included. A conductor-like polarizable continuum model (CPCM) 36 solvent cavity model of CCl4 was applied; the approach has been reported to remain accurate in similar studies. 8,11 It is known that the solvent effect is easily observable in NIR spectra of comparable molecules.37,38 Relative simplicity and parameterization of SCRF framework may incorporate artificial effects on the results of vibrational analysis. A noticeable variability due to the solvent model has been observed in similar studies before;37,39 the dependence on particular molecular system inhibits effective generalization. Therefore, separate computations were performed in vacuum without implicit solvation model; a comparison and discussion of the results obtained in both ways will be provided in Section 3. Following the geometry optimization a vibrational analysis was performed, yielding harmonic frequencies and intensities of fundamental vibrations. The determined properties were subsequently used for the needs of fully anharmonic vibrational analysis by means of deperturbed second-order vibrational perturbation theory (DVPT2) 40 giving anharmonic frequencies and intensities of fundamental vibrations, as well as NIR modes - the first overtones and binary combinations. The latter ones were used for the purpose of simulation of NIR spectrum of thymol.

5

176 177 178 179 180 181 182 183

The selected approach, including combination of B3LYP functional and SNST basis set have frequently been used in VPT2 (DVPT2/GVPT2) computations, including NIR spectra simulations with good results.8,30,41,42 The final simulation of NIR spectra involved convolution procedure in order to reflect the band broadening and band shape effects. For this intention a four-parameter Lorentz-Gauss product function was employed; this model gives good agreement with experimental vibrational bands in solution phase. 43,44 All QM calculation were performed with the use of Gaussian 09 D.01 suite of programs, 45 while the spectra simulation and processing, preparation of mode density maps and assembly of figures was carried out in MATLAB R2016b. 46

184

185 186

3. Results and discussion

187 188

3.1 The experimental NIR spectrum of thymol and its correspondence with sample state and concentration level

189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207

The NIR spectra of thymol measured for solid state, melted (neat liquid; 333 K) and diluted (10-100 mg mL-1 CCl4) sample uncover a pattern of relationship to the sample state and concentration (Fig. 1). There exist a decisive difference between the spectral outlines of polycrystalline and melted thymol. The appearance of a broad structure (6800-5100 cm-1) as well as baseline elevation (5000-4000 cm-1 supposedly extending below this region) should be accounted to strong anharmonic effects, which arise in the ordered periodic structure of crystal. 47,48 These effects emerge due to highly symmetric hydrogen-bonding (HB) network of the OH group and have wellknown impact on IR spectra.47,48,49,50, The bands between 5000-4500 cm-1 wavenumbers also vary between solid and melted thymol, although more gradually. In the melted state the long-range ordering of the HB network is broken, although the concentration effect and short-range interaction through hydrogen-bonding are likely preserved. The dilution in an inert solvent introduces major difference in the regions of ca. 7000 cm -1 (2OH band) and 5000-4500 cm-1. However, once a reasonable isolation level of thymol molecules by the solvent is achieved (100 mg mL-1 CCl4 ) any further dilution (10 mg mL-1 CCl4 ) does not induce major changes. There are two spectral regions clearly independent of the sample state or concentration; 6000-5600 cm-1 and 4490-4000 cm-1 (Fig. 1). Even in the polycrystalline thymol these bands are largely invariant, although remain superimposed to the broad structures as explained earlier. The spectral features outlined in this paragraph will be further discussed with an aid of detailed band assignments and in the context of the relationship with the multi-variate quantification of thymol. 17

6

208 209 210 211

Fig. 2. A set of the experimental NIR spectra of thymol; solid state and melted (neat liquid, 333 K) as well as diluted in -1 CCl4 (100 and 10 mg mL CCl4). Highlighted are the wavenumber regions qualitatively independent of sample phase and concentration; A: 6000-5600 cm-1; B: 4490-4000 cm-1.

212 213 214

3.2 The simulated NIR spectrum of thymol; the role of solvation model, mode contributions and detailed band assignments

215 216 217 218 219 220 221 222

The present limits of anharmonic QM spectra simulations prevent feasible studies of crystalline thymol.11 We will base our modeling on the thymol molecule placed in solvent cavity for direct comparison with the corresponding experimental data (soluted analyte; 100 mg mL-1 CCl4 ). The spectral cross-analysis presented above will allow following unambiguously the differences among the measured spectra. As explained in Section 2.2 the application of solvent model usually improves the results, but this may vary depending on given molecule. In the present study an incorporation of CPCM solvent model of CCl4 enhances the agreement with the experimental spectrum, as evidenced in Fig. 3.

7

223 224 225 226

Fig. 3. Impact of implicit solvation model (SCRF/CPCM-CCl4) on the simulated (DVPT2//DFT-B3LYP/SNST+CPCM) NIR spectrum of thymol. The modeled spectral outlines compared with the experimental spectrum of thymol in solution (100 -1 mg mL CCl4).

227 228 229 230 231 232 233 234 235 236

The accuracy of simulated NIR spectrum is lower when an isolated, placed in vacuum molecule is considered (Fig. 3). A good indicator could be formed by the less accurate position of the simulated 2OH peak, which arises from a highly anharmonic and sensitive mode which has often been focused on in physicochemical NIR studies. For the purpose of qualitative analysis of the entire NIR region, however, even minor bands play a role in drawing right conclusions. On top of an appearance of improper bands (i.e. strong bands at 5550 cm-1; 4245 cm-1; 3897 cm-1) the shape of the spectral outline is generally lower when the solvation model is omitted with the exception of 5300 4500 cm-1 region. Based on these observations for further reasoning the spectrum simulated with implicit solvent approximation will be employed.

237 238 239 240 241 242 243 244 245 246 247

The computational approach assumed in this work involves modeling of the first overtone and binary combination bands. As investigated recently, these bands are responsible for the vast majority (ca. 80%) of spectral information emerging in NIR region.51 Inclusion of higher order modes requires a considerable increase in the usage of computational resources; however, it does not offer a momentous improvement of the quality of simulation. 51 Higher order overtones and ternary combinations are significantly weaker bands and what’s more the corresponding spectral information is “diffused” among weak, overlapping bands. 51 A possible general exception known in literature is the second overtone of C=O stretching mode, which may appear as a well resolved NIR band,52 obviously absent in the case of thymol. 12,53 Indeed, an overview of the theoretical outline and the modal contributions as compared with the experimental spectrum confirms that the simulation reflects the majority of the spectral features observed experimentally (Fig. 4).

8

248 249 250

Fig. 4. The contributions of the first overtone and binary combination bands into the NIR spectral envelope of thymol in solution phase (100 mg mL-1 CCl4) as obtained through QM spectra simulation (DVPT2//DFT-B3LYP/SNST+CPCM).

251 252 253 254

Fig. 5. The most probable band assignments in the experimental NIR spectrum of thymol in solution phase (100 mg mL-1 CCl4) based on comparative cross-analysis with the QM simulated NIR spectrum (DVPT2//DFT-B3LYP/SNST+CPCM). Band numbers correspond to those presented in Table 1.

9

255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277

Spectra simulation provides a complete set of band assignments extending far beyond of the capabilities of classical spectroscopic methods. 8 Fig. 5 and Table 1 gather the major contributions of vibrational modes into the NIR spectrum of thymol. For sake of clarity, the experimental peaks have been ascribed following the most influential modes; high level of band overlapping (clearly evidenced in the simulation [Fig. 4]) makes elucidating concurrent minor contributions difficult. For the purpose of presenting deeper insights into the origin of the NIR spectrum of thymol a different approach will be applied as explained in Section 3.3. The overlapping is particularly noticeable in the lower NIR region, 4500-4000 cm-1; the spectral outline there may be described as a convolution of numerous weak binary combinations. Nevertheless, the model envelope resulting from the summed contributions reproduces the experimental spectrum reasonably well (Fig. 4). A considerable band overlay can be noticed between 6100-5600 cm-1 as well, where also first overtone bands appear influential. The agreement with the experimental spectrum is lower here, as the corresponding bands originate from highly anharmonic modes; the first overtone intensities are overestimated in the simulation. Similar feature has been reported for other molecules. 54 There exist some minor discrepancies in other subregions; i.e. the combination band of OH stretching and CH stretching is vastly overestimated in its height. However, these minor shortcomings may be identified effortlessly and do not inhibit comprehensive assignments (Fig. 5; Table 1). With the exception of highly anharmonic CH and CH3 stretching modes, the simulated peak wavenumbers remain remarkably correct (Table 1). The discrepancies do not exceed ca. 20 cm -1 in most cases; this translates to the error level of less than 0.5%. A relatively accurate position of the simulated 2OH band (underestimated by 9 cm-1 or 0.12 %), which is highly sensitive to the intermolecular interactions confirms that the concentration effects do not play a major role in the diluted thymol sample.

278 279 280 281

Table 1. The proposed band assignments in the experimental NIR spectrum of thymol in solution phase (100 mg mL-1 CCl4). Band numbers correspond directly to those presented in Fig. 5. The listed contributions of the respective modes are limited to the most influential ones for sake of clarity.

Wavenumber / cm-1

Band number

Experimental

Calculated

Difference

1

7056

7047

-9

2OH

2

6488

6503

15

CHalkyl + OH

3

6072

6054

-18

2s CHring, s CHring + s CHring

4

~5972

5889

-83

2as CHring

5

5920

5873

-47

as CH3, as CH3 + as CH3 

6

5872

5832

-40

as CH3

7

5748

5751

3

s CH3, s CH3 + as CH3

8

5644

5676

32

CHalkyl

9

5236

5237

1

 ipCCring + OH

Major mode contribution

10

10

5196

5186

-10

 ipCCring + OH

11

5128

5119

-9

 ipCHring + OH

12

5028

5041

13

 as CH3 + OH

13

4972

4990

18

 s CH3 + OH

14

4924

4937

13

 ipCHring + OH

15

4828

4855

27

 ipCHring + OH

16

4786, 4764

4793

-

 ipCHring + OH

17

4700

4722

22

 ipOH + OH

18

4620

4616

-4

 ip,as CCring + as CHring

19

4396

4387

-9

CH3 + as CH3 

20

4320

4315, 4294

-

 s CH3 + as CH3;  s CH3 + as CH3,  as CH3 + s CH3 

21

4260, 4236

4242

-

 s CH3 + s CH3 

22

4176

4197

21

 oopCHalkyl + CHalkyl

23

4124

4155

31

 ipOH + as CHring

24

4072

4062

-10

 rockCH3 + s CHring

25

3928

3976, 3908

-

OH + OH, CCC + OH

26

3828

3830

2

 rockCH3 + s CH3

27

3764

3761

-3

C(CH3) 2 + OH

282 283 284 285

3.3 An in-depth analysis of the origin of NIR spectrum of thymol and its relationship with PLS regression coefficient vector in quantification of thymol content

286 287 288 289 290 291 292 293 294 295

With the aim at elucidating the generalized, mode-specific spectral contributions we have developed a density map of mode influence in the form of colormap (or heatmap) plot capable of highlighting the modes of interest in a straightforward way (Fig. 6). The color range corresponds to the square rooted intensity ratio of selected simulated bands to the total intensity of modeled spectrum at any given wavenumber i. The yielded value ranges from 0 (no contribution) to 1 (the NIR spectrum is influenced by the selected modes entirely). The square root allows elucidating less pronounced contributions. The density maps determined for various selections of modes-of-interest allow unequivocal and thorough analysis of the influential determinants in NIR spectrum. The most general classification into overtones and combinations (Fig. 6) brings an already known picture (Fig. 4); however, tuning the selection to a more specific selection yields insightful correspondences. In

11

296 297 298 299 300 301 302 303 304 305 306 307 308 309

example, the combination bands involving OH stretching mode are one of the most decisive factors in forming NIR spectrum of thymol. Decomposing these into less general groups brings further details. I.e. OH + CH(ring) and OH + CH3 modes are relatively insignificant, while the combinations of OH and methyl deformations are highly influential only in a narrow wavenumber range (5080-4950 cm-1 ) and also appear between 4700 and 4600 cm -1 although with less impact. It can be easily assessed that the combinations of methyl stretching and deformation modes are largely responsible for the strong absorption region of 4450-4220 cm-1 (with peaks at 4396, 4320, 4260 and 4236 cm-1; Table 1). The distribution of the influence can be followed as well. In example, combinations of OH +CH3, CH3+CH3 and CH3+CH3 type remain highly localized in narrow wavenumber ranges. On the contrary, the combinations involving ring deformations (i.e. OH +ring and CH3+ring) can be described as diffused over wider regions along the wavenumber axis (Fig. 6). These projections allow for flexible yet straightforward generalizations and can be helpful for an indepth qualitative assessment of an NIR spectrum with highlight of categorized and grouped mode contributions.

310 311 312 313 314 315 316 317 318 319 320 321 322 323

Taking advantage of recently analyzed NIR spectrum of phenol, 30 a brief comparison with present results may be made. Only few of thymol NIR features remain analogous with those of phenol; these bands stand apart from those of aliphatic linear and cyclic alcohols. 30 In example, the major 2OH peak appears at 7056 cm-1, near that of phenol (7052 cm-1); while in the other kinds of alcohols it tends to emerge at higher wavenumbers (7103-7073 cm-1).30 Likewise, the stretching modes of CH in ring (both first overtones and binary combinations) appear above 6000 cm-1 similar to phenol.30 In this regard, the discussed molecule keeps the vibrational properties common to phenols.30 However, while being a phenol derivative thymol features a substantial increase in the number of NIR bands in its spectrum. This should be mostly attributed to the addition of the methyl groups, as these are fairly influential particularly between 4450-3900 cm-1 but also in the 5900-5700 cm-1 region (Fig. 6). In the NIR spectrum of thymol a high separation on the wavenumber axis of the stretching modes C-H in ring and C-H in the isopropyl group should be noted as well, with the latter ones appearing below 5700 cm-1 (Fig. 5; Table 1). The two above factors enrich the spectrum of thymol as opposed to a relatively featureless phenol in the discussed region (6100-5600 cm-1).30

12

324 325 326 327

-1

Fig. 6. The analysis of mode contribution into NIR spectrum of thymol (solution; 100 mg mL CCl4) based on the simulated data (DVPT2//DFT-B3LYP/SNST+CPCM). A: experimental and simulated outlines. B: contributions of selected modes as described on the figure.

13

328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347

With the information set provided by the above analyses we can finally draw more general conclusions on the spectral correspondence to the sample state and concentration (Fig. 2, Section 3.1). The upper wavenumber region which remains largely invariant of these factors is mostly populated by the combinations and overtones involving C-H and CH3 stretching modes (region A in Fig. 2). In the region B (Fig. 2) the combinations of methyl stretching and deformation modes dominate (Fig. 6) with secondary influences from the combinations which involve ring deformations. These bands feature a relative insensitiveness to the sample state, preserving their position and bandshape as well as intensity in relation to other bands (Fig. 2). An exception should be mentioned, the absorption within the region B is relatively lower than in the region A for the solid state thymol; this is likely due to the superimposition of the broadened structure appearing in the region A and originating from ordered HB network which exists in crystalline phase (Fig. 2). The highest spectral variability develops between 5400-4500 cm-1, where decisive changes can be monitored with the sample melt and dilution (Fig. 2). A strong redshift and broadening of the entirety of the bands follows the transition from diluted to neat liquid (melted) sample and with even further aggregation with a single peak appearing at ca. 4600 cm-1. The discussed region is nearly exclusively formed by the combinations of OH stretching modes coupled to ring deformation, C-H (in ring) bending and methyl deformation modes (Fig. 5; Table 1). Cross-analysis with the aid of categorized combinations (Fig. 6) allows concluding that the high spectral variability is likely influenced by the specificity of OH mode. Its sensitivity to the intermolecular interactions manifests itself strongly in this case.

348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372

A recent study has demonstrated the feasibility of quantification of thymol content in Thymi herba by employing multi-variate calibration based on NIR spectra of the natural sample.17 The constructed PLS regression coefficient vector has indicated a number of bands which introduce the highest influence on the calibration model for thymol. 17 These bands fall with no exception into the two “invariant” regions established in the present work. The influential wavenumbers [cm-1] have been reported as: 5860, 5760 (falling into the region A; Fig. 2) and 4476, 4418, 4392, 4368, 4220, 4128, 4092 (belonging in the region B; Fig. 2). 17 The major origin of these two groups can be summarized as overtones and combinations of CH3 and aromatic (region A) and combinations involvingCH3 (region B) modes. It can be concluded that the spectral information corresponding to the otherwise highly influential (in the sense of spectrum forming factor; Fig. 6) OH mode has not been recognized as being decisive in the quantification by PLS model.17 This includes a discrimination of the 5300-4500 cm-1 region. It may be argued that a “diffused” character (being spread along wide wavenumber regions) of these bands plays a role here. It should also be mentioned, that the impact of quantified molecule on its chemical surrounding through intermolecular interactions (i.e. HB or aromatic ring stacking in the case of thymol) creates a set of relevant spectral correlation which may likely be identified in the regression vector. Further studies could be aimed at simulation of these effects. The literature knows the concept of incorporating the band assignment in the chemometrics for improvement of the analytical performance of NIRS.55 However, the hindrance in utilization of such approaches has typically been imposed by lack of ability to obtain a detailed decomposition of an NIR spectrum; a barrier which may largely be removed by using spectra simulation. The correspondence of the linear spectra and the regression plots can also be found in the literature;56 however, most often such comparisons are facing similar limitations in their interpretation as the case described in the past sentence. The current preliminary study demonstrates the potential of QM in bringing deeper understanding of the relationship between the chemometric models and molecular and physiochemical background of the quantified sample.

14

373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393

4. Summary QM simulation forms a major support tool as NIR spectra are inherently complex with strong band overlay; the anharmonic effects and sensitiveness to molecular interactions further inhibit the applicability of classical spectroscopic methods for an in-depth analysis of NIR data. By employing anharmonic calculations a complete decomposition of the NIR spectrum of thymol was possible, yielding comprehensive band assignments throughout the entire NIR region. The experimental study uncovered a patterned spectral variability in dependence of the sample state (solid and melted, neat liquid) and concentration (neat liquid and diluted in an inert solvent) . Spectral regions featuring high invariance to the sample property were identified. Further insights could be provided with the use of modal density plots; these allowed generalization, categorization and assessment of the spectrum forming factors and their straightforward visualization. The influential wavenumbers could be confronted with the major features of the regression coefficient vector in PLS quantification model of thymol content as reported recently. It was found that all the wavenumbers which have been identified as decisive in the PLS regression fall into the two “invariant” NIR spectral sub-regions of thymol as established in the present work. These regions are populated by the bands arising from CH3 and aromatic CH modes. Surprisingly, otherwise the most influential for the spectrum forming OH mode has been largely discriminated in the chemometric model. It was suggested that the sensitiveness of the OH mode to the intermolecular interactions, leading to a strong broadening of the corresponding bands may be accounted to the observed regularity. The presented study demonstrates the potential of gaining decisive advance in the capability of spectral analysis in the NIR region through utilization of QM spectra simulations.

394 395

Acknowledgement

396 397

Calculations have been carried out in Wrocław Centre for Networking and Supercomputing (http://www.wcss.pl), under grant no. 375.

398 399

References 1

Y. Ozaki, C.W. Huck, K.B. Bed. Near Infrared Spectroscopy and its Applications in V.P. Gupta (Ed.)

Molecular and Laser Spectroscopy, Elsevier, 2017. 2

C.W. Huck. Infrared Spectroscopy in Near-infrared/Infrared Bioanalysis Including Imaging, John

Wiley & Sons, Encyclopedia of Analytical Chemistry, 2016. 3

E.W. Ciurczak, J.K. Drennen. III Pharmaceutical and Medical Applications of Near-Infrared

Spectroscopy, CRC Press: Boca Raton, 2002. 4

C.W. Huck. Near-Infrared (NIR) Spectroscopy in Natural Product Research in K. Hostettman, S. Chen,

A. Marston, H. Stuppner (Eds.) Handbook of Chemical and Biological Plant Analytical Methods, John Wiley & Sons, 2014. 5

S. Saranwong, S. Kawano, A. Ikehata, G. Noguchi, S. Park, K. Sashida, T. Okura, R. Haff. Development

of a low-cost NIR instrument for minced meat analysis: Part 1 - Spectrophotometer and sample presentations. Am. J. Agric. Sci. Technol., 2013, 2, 61-68.

15

6

L. Herberholz, O. Kolomiets, H.W. Siesler. Quantitative analysis by a portable near infrared

spectrometer: Can it replace laboratory instrumentation for in situ analysis? NIR News, 2010, 21, 1-8. 7

R. Henn, A. Schwab, C.W. Huck. Evaluation of Benchtop versus Portable Near-Infrared Spectroscopic

Method Combined with Multivariate Approaches for the Fast and Simultaneous Quantitative Analysis of Main Sugars in Syrup Formulations, Food Contr., 2016, 68, 97-104. 8

K.B. Bed, Y. Futami, M.J. Wójcik, Y. Ozaki. A spectroscopic and theoretical study in the near-infrared

region of low concentration aliphatic alcohols, Phys. Chem. Chem. Phys., 2016, 18, 13666-13682. 9

L. Dymioska, J. Janczak, K.S.M. Sheweshen, J. Lorenc, J. Hanuza. Crystal and molecular structures,

temperature dependence of the IR and Raman spectra and vibrational dynamics of aquo 4,6 dimethyl-5H-[1,2,3]triazolo[4,5-c]pyridine in a new zwitterionic form. J. Mol. Struct., 2017, 1144, 482495. 10

E. Kucharska, J. Hanuza, J. Lorenc. Conformation of azo-bridge in 3,3′-dinitro-2,2′-azobipyridine and

its 4,4′(or 5,5′ or 6,6′)-dimethyl-derivatives: Vibrational studies and DFT quantum chemical calculations. Spectrochim. Acta A, 2014, 127, 370-380. 11

K.B. Bed, J. Grabska, Y. Ozaki. Advances in Anharmonic Methods and Their Applications to

Vibrational Spectroscopies. In M.J. Wojcik, H. Nakatsuji, B. Kirtman, Y. Ozaki (Eds.) Frontiers of Quantum Chemistry. Springer, 2017. 12

L.G. Weyer, S.C. Lo. Spectra-Structure Correlations in the Near-Infrared. In J. M. Chalmers, P. R.

Griffiths (Eds.). Handbook of Vibrational Spectroscopy. Wiley, Chichester, 2002; Vol. 3, 1817. 13

P. Wang, Z. Yu. Species authentication and geographical origin discrimination of herbal medicines

by near infrared spectroscopy: A review. J. Pharmaceut. Anal. 2015, 5, 277-284. 14

S. Türker-Kaya, C.W. Huck. A Review of Mid-Infraraed and Near-Infrared Imaging: Principles,

Concepts and Applications in Plant Tissue Analysis. Molecules, 2017, 22, 168-188. 15

C.W. Huck. Miniaturized NIR and MIR Sensors for Medicinal Plant Quality Control. Spectroscopy,

2017, 32/s8, 8-15. 16 G. Tang, Q. Cao, D. Wang, S. Ji. Application of Near-Infrared Spectroscopy in Determination of Schaftoside Acid in Desmodium styracifolium (Osb.) Merr. Merr. Spectral Analysis Reviews, 2018, 6, 33-42. 17

C.K. Pezzei, S.A. Schönbichler, S. Hussain, C.G. Kirchler, V.A. Huck-Pezzei, M. Popp, J. Krolitzek, G.K.

Bonn, C.W. Huck. Near-infrared and Mid-infrared Spectroscopic Techniques for a Fast and Nondestructive Quality Control of Thymi herba, Planta Medica, 2017, 83, 1-8. 18

I. Fecka I, S. Turek. Determination of polyphenolic compounds in commercial herbal drugs and

spices from Lamiaceae: thyme, wild thyme and sweet marjoram by chromatographic techniques. Food Chem. 2008, 108, 1039–1053.

16

19

S.J. Lee, K. Umano, T. Shibamoto, K.G. Lee. Identification of volatile components in basil (Ocimum

basilicum L.) and thyme leaves (Thymus vulgaris L.) and their antioxidant properties. Food Chem., 2005, 91, 131–137. 20

C.G. Kirchler, C.K. Pezzei, K.B. Bed, S. Mayr, M.Ishigaki, Y. Ozaki, C.W. Huck. Critical evaluation of

spectral information of benchtop vs. portable near-infrared spectrometers: Quantum chemistry and two-dimensional correlation spectroscopy for a better understanding of PLS regression models of the rosmarinic acid content in Rosmarini folium. Analyst, 2017, 142, 455-464. 21

M.J. O'Neil (ed.). The Merck Index - An Encyclopedia of Chemicals, Drugs, and Biologicals.

Cambridge, UK: Royal Society of Chemistry, 2013., p. 1741. 22

M. Segvid-Klarid, I. Kosalec, J. Mastelid, E. Piecková, S. Pepeljnak. Antifungal activity of thyme

(Thymus vulgaris L.) essential oil and thymol against moulds from damp dwellings. Lett. Appl. Microbiol., 2007, 44, 36–42. 23

N. Didry, L. Dubreuil, M. Pinkas. Activity of thymol, carvacrol, cinnamaldehyde and eugenol on oral

bacteria. Pharm. Acta Helv., 1994, 69, 25–28. 24

Council of Europe. European Pharmacopoeia, 8th ed 2013, English: Subscription to Main volume +

Supplement 1 + Supplement 2, 8th ed. Stuttgart: Deutscher Apotheker Verlag; 2013. 25

A. Astudillo, E. Hong, R. Bye, A. Navarrete. Antispasmodic activity of extracts and compounds of

Acalypha phleoides. Cav. Phytother. Res., 2004, 18, 102–106. 26

P.C. Braga, M. Dal Sasso, M. Culici, T. Bianchi, L. Bordoni, L. Marabini. Antiinflammatory activity of

thymol: inhibitory effect on the release of human neutrophil elastase. Pharmacology, 2006, 77, 130– 136. 27

Y. Suzuki, H. Furuta. Stimulation of guinea pig neutrophil superoxide anion-producing system with

thymol. Inflammation, 1988, 12, 575–584. 28

S. Fujisawa, Y. Kadoma. Effect of phenolic compounds on the polymerization of methyl

methacrylate. Dent. Mater. 1992, 8, 324–326. 29

T. Dingermann, K. Hiller, G. Schneider, I. Zündorf. Schneider – Arzneidrogen. Heidelberg: Springer

Spektrum, 2004. 30

K.B. Bed, J. Grabska, M.A. Czarnecki. Spectra-structure correlations in NIR region: Spectroscopic

and anharmonic DFT study of n-hexanol, cyclohexanol and phenol. Spectrochim. Acta A, 2018, 197, 176-184. 31

K.B. Bed, M.J. Wojcik, T. Nakajima. Quantum chemical calculations of basic molecules: alcohols and

carboxylic acids. NIR news, 2016, 27, 15-21. 32

A.D. Becke. Density‐functional thermochemistry. III. The role of exact exchange. J.Chem.Phys.,

1993, 98, 5648-5652.

17

33

V. Barone, P. Cimino, E. Stendardo. Development and Validation of the B3LYP/N07D Computational

Model for Structural Parameter and Magnetic Tensors of Large Free Radicals, J. Chem. Theory Comput., 2008, 4, 751-764. 34

S. Grimme, J. Antony, S. Ehrlich, H. Krieg. A consistent and accurate ab initio parameterization of

density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys., 2010, 132, 154104. 35

J. Tomasi, B. Mennucci, R. Cammi. Quantum mechanical continuum solvation models. Chem. Rev.,

2005, 105, 2999-3093. 36

V. Barone, M. Cossi. Quantum calculation of molecular energies and energy gradients in solution by

a conductor solvent model, J. Phys. Chem. A, 1998, 102, 1995–2001. 37

T. Gonjo, Y. Futami, Y. Morisawa, M.J. Wójcik, Y. Ozaki. Hydrogen Bonding Effects on the

Wavenumbers and Absorption Intensities of the OH Fundamental and the First, Second and Third Overtones of Phenol and 2,6-Dihalogenated Phenols Studied by Visible/Near-Infrared/Infrared Spectroscopy and Density Functional Theory Calculations. J. Phys. Chem. A, 2011, 115, 9845-9853. 38

M.A. Czarnecki, K. Orzechowski. Effect of temperature and concentration on self -association of

octan-3-ol studied by vibrational spectroscopy and dielectric measurements. J. Phys. Chem. A, 2003, 107, 1119-1126. 39

Y. Mikami, A. Ikehata, C. Hashimoto, Y. Ozaki. Near-Infrared (NIR) Study of Hydrogen Bonding of

Methanol Molecules in Polar and Nonpolar Solvents: An Approach from Concentration-Dependent Molar Absorptivity. Appl. Spectrosc., 2014, 68, 1181-1189. 40

V. Barone. Anharmonic vibrational properties by a fully automated second-order perturbative

approach, J. Chem. Phys., 2005, 122, 014108-10. 41

M. Hodecker, M. Biczysko, A. Dreuw, V. Barone. Simulation of Vacuum UV Absorption and

Electronic Circular Dichroism Spectra of Methyl Oxirane: The Role of Vibrational Effects. J. Chem. Theory Comput., 2016, 12, 2820-33. 42

J. Grabska, M.A. Czarnecki, K.B. Bed, Y. Ozaki. Spectroscopic and Quantum Mechanical Calculation

Study of the Effect of Isotopic Substitution on NIR Spectra of Methanol. J. Phys. Chem. A, 2017, 121, 7925–7936. 43

J. P. Hawranek. On the numerical description of asymmetric absorption bands. Acta Phys. Pol. B,

1971, 40, 811−814. 44

J. Grabska, M. Ishigaki, K.B. Bed, M.J. Wójcik, Y. Ozaki. Correlations between structure and near-

infrared spectra of saturated and unsaturated carboxylic acids. An insight from anharmonic DFT calculations, J. Phys. Chem. A, 2017, 121, 3437–3451. 45

M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V.

Barone, B. Mennucci, G.A. Petersson, et al. Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2013.

18

46 47

MATLAB, The MathWorks, Inc., Natick, Massachusetts, United States. M. Boczar, Ł. Boda, M.J. Wójcik. Theoretical modeling of infrared spectra of hydrogen-bonded

crystals of salicylic acid. Spectrochim. Acta A, 2006, 64, 757-760. 48

H.T. Flakus, M. Chelmecki. Infrared spectra of the hydrogen bond in benzoic acid crystals:

Temperature and polarization effects. Spectrochim. Acta A, 2002, 58, 179-196. 49

J. Hanuza, P. Godlewska, E. Kucharska, M. Ptak, M. Kopacz, M. Mączka, K. Hermanowicz, L. Macalik.

Molecular structure and vibrational spectra of quercetin and quercetin-5’-sulfonic acid. Vib. Spectrosc., 2017, 88, 94-105. 50

E. Kucharska, I. Bryndal, T. Lis, J. Lorenc, J. Hanuza. Influence of methyl and nitro group

substitutions on the structure and vibrational characteristics of the hydrazo-bridge in 6,6′-dimethyl3,3′,5,5′-tetranitro-2,2′-hydrazobipyridine. Vib. Spectrosc., 2016, 83, 70-77. 51

J. Grabska, M.A. Czarnecki, K.B. Bed, Y. Ozaki. Spectroscopic and Quantum Mechanical Calculation

Study of the Effect of Isotopic Substitution on NIR spectra of Methanol, J. Phys. Chem. A, 2017, 121, 7925-7936. 52

Y. Chen, Y. Morisawa, Y. Futami, M.A. Czarnecki, H.-S. Wang, Y. Ozaki. Combined IR/NIR and

Density Functional Theory Calculations Analysis of the Solvent Effects on Frequencies and Intensities of the Fundamental and Overtones of the C=O Stretching Vibrations of Acetone and 2 -Hexanone. J. Phys. Chem. A 2014, 118, 2576−2583. 53

J. Workman, Jr., L. Weyer. Practical Guide and Spectral Atlas for Interpretive Near-Infrared

Spectroscopy, Second Edition. CRC Press, 2012. 54

J. Grabska, K.B. Bed, Y. Ozaki, C.W. Huck. Temperature Drift of Conformational Equilibria of Butyl

Alcohols Studied by Near-Infrared Spectroscopy and Fully Anharmonic DFT. J. Phys. Chem. A, 2017, 121, 1950-1961. 55

F. Westad, A. Schmidt, M. Kermit. Incorporating chemical band-assignment in near infrared

spectroscopy regression models, J. Near Infrared Spectrosc., 2008, 16, 265-273. 56

M. Saleem, M. Bilal, S. Anwar, A. Rehman, M. Ahmed. Optical diagnosis of dengue virus infection in

human blood serum using Raman spectroscopy. Laser Phys. Lett., 2013, 10, 035602-035607.

19

Suggest Documents