Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Development of a high-resolution, broadband spatial heterodyne Raman spectrometer based on field-widened grating-echelle structure

Open Access Open Access

Abstract

We propose a spatial heterodyne Raman spectrometer (SHRS) based on a field-widened grating-echelle (FWGE). A normal grating is combined with an echelle grating in a conventional spatial heterodyne spectrometer to eliminate ghost images without using masks, and prevents interference among the spatial frequencies of different diffraction orders. Mathematical expressions and derivation processes are given for the spectral parameters in the FWGE-SHRS and a verification breadboard system is fabricated. The FWGE-SHRS measures Raman spectra of single chemicals and mixed targets with different integration times, laser powers, concentrations, and transparent containers. The results of the experiments demonstrate that the FWGE-SHRS is suitable for high-resolution, broadband Raman measurements for a wide range of applications.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

As one of accurate and nondestructive methods for material identification and component analysis [1,2], Raman spectroscopy has been applied widely in fields including medicine [3,4], geology [5,6], physics [7,8], biology [9,10], and chemistry [11,12]. In recent years, the spatial heterodyne spectrometer (SHS) has emerged as a novel type of Fourier interferometer. The SHS offers advantages that include a one-time measurement capability, high stability, large throughput, and broadband or high spectral resolution, which makes the spatial heterodyne Raman spectrometer (SHRS) highly suitable for accuracy Raman spectrum measurement applications.

However, because of the restrictions of the Nyquist sampling theorem, it is difficult for the SHRS to meet requirements for high spectral resolution and a broadband spectral range simultaneously [13]. Some new structures have been proposed to solve this restriction in conventional SHS equipment, with particular focus on replacement of the normal grating with a multi-grating structure consisting of many sub-gratings with different groove densities [1416]. However, direct replacement of the normal grating with a multi-grating will reduce the signal light throughput. Additionally, the intensity of the spectrum within each wave band will decrease as the number of sub-gratings increases. Researchers have also focused on the replacement of normal gratings with echelle gratings with multiple orders. With its mixture of different diffraction orders, the double-echelle SHS (DESHS) can realize high-resolution and broadband measurement capabilities [17]. However, to eliminate the effects of ghosts and shadows, the DESHS structure requires a mask to be inserted at the focal plane of the imaging optics to select the spectral information for one specific order of the two echelle gratings [1820], which then prevents the overall spectrum from being obtained in a single measurement process. However, ghosts and shadows can still be detected in the spectral information when using this structure. Replacing one echelle grating with a mirror (mirror-echelle SHS, or MESHS) can solve these problems in the DESHS [21]. However, the introduction of the mirror reduces the spectral resolution by half, thus losing the advantage of the system. In addition, the field-widening theories for the echelle grating and the mirror are not suitable, and this indicates that the field-of-view (FOV) of the MESHS cannot be improved as compared with the conventional SHS.

In response to the issues noted above, we proposed the concept of the grating-echelle SHS (GESHS) structure, which replaces one normal grating in the conventional SHS with an echelle grating. The proposed GESHS structure can not only use the characteristics of multiple diffraction orders to achieve high resolution and broadband measurements in a single measurement process without using a mask or sacrificing the spectral resolution, but also can increase the FOV by introducing field-widening prisms that correspond to the normal grating and the echelle grating based on field-widening theory. As a result of the different diffraction characteristics of the normal grating and the echelle grating, the mathematical expressions for the Littrow wavenumber and the other spectral parameters in the GESHS are different to the corresponding expressions for the conventional SHS.

In this article, the mathematical derivation process for the principles and the expressions of the GESHS is presented. The calibration theory is also discussed in detail. On this basis, using the field-widening theory, we have designed and manufactured suitable field-widening prisms for both the normal grating and the echelle grating and applied them in the Raman experiment to broaden the FOV of the spectrometer; this is why we ultimately called the proposed structure the field-widened grating-echelle spatial heterodyne Raman spectrometer (FWGE-SHRS). In experiments, the detailed Raman spectra of chemicals acquired at different integration times, laser powers, and concentrations are studied in detail. The Raman spectra of mixtures of inorganic powders in plastic bags of various thicknesses and the Raman spectra of a mixture of organic solutions in a glass bottle are also presented.

2. Principle

2.1 Basic theory

Figure 1 illustrates the experiment optical path of the FWGE-SHRS. During the Raman signal excitation process, the light beam output by the laser is finally focused via collimation lens L1 onto the sample surface after being filtered using a linear variable filter and a clean-up filter and then having its direction changed by the mirror and the dichroic beam splitter. The Raman signal light that is excited by the sample is then collected by collimation lens L1 and filtered using the long-pass and short-pass filters. The filtered signal light is finally collected by collimation lens L2 and launched into optical fibers for the spectral analysis. The core component of the experiment’s optical path is the principal prototype of the FWGE-SHRS. The signal light from the fiber output port is first collimated to a specific size by collimation lens L3 and is then divided into two beams by the beam splitter. One of the beams is diffracted by the first order of the normal grating G, and the other beam is diffracted by several orders of the echelle grating E; both the normal grating and the echelle grating are set in the Littrow condition [22], which can be expressed for the respective gratings as:

$$2{\sigma _{LG}}\sin {\alpha _{LG}} = \frac{1}{{{d_G}}}\textrm{ }$$
$$2{\sigma _{LEm}}\sin {\alpha _{LE}} = \frac{m}{{{d_E}}}\textrm{ }$$
where σLG and αLG are the Littrow wavenumber and the Littrow angle of normal grating G, respectively, and 1/dG is the groove density of the normal grating. Assuming that the echelle grating uses q continuous diffraction orders, beginning from the order p, m is the diffraction order used for the echelle grating, which can be expressed as m = p, p−1, , p-q + 1. σLEm is the Littrow wavenumber of the m-th order of echelle grating E, αLE is the Littrow angle of the echelle grating, and 1/dE is the echelle grating groove density.

 figure: Fig. 1.

Fig. 1. Optical layout of the experimental light path for the spatial heterodyne Raman spectrometer based on the field-widened grating-echelle structure.

Download Full Size | PDF

For the on-axis incident light signal, the diffraction equation for the two gratings with the off-Littrow wavenumbers can be expressed as follows:

$$\sigma ({\sin {\alpha_{LG}} + \sin ({{\alpha_{LG}} - {\beta_G}} )} )= \frac{1}{{{d_G}}}\textrm{ }$$
$$\sigma ({\sin {\alpha_{LE}} + \sin ({{\alpha_{LE}} - {\beta_E}} )} )= \frac{m}{{{d_E}}}\textrm{ }$$
where σ is the wavenumber of the incident light, and βG and βE are the off-Littrow angles of the diffracted light from the normal grating G and the echelle grating E, respectively. Substitution of Eq. (1) and Eq. (2) into Eq. (3) and Eq. (4), respectively, and expansion of these expressions to the first order as βG and βE gives [22]
$${\beta _G} = 2\tan {\alpha _{LG}}\left( {\frac{{\sigma - {\sigma_{LG}}}}{\sigma }} \right)$$
$${\beta _E} = 2\tan {\alpha _{LE}}\left( {\frac{{\sigma - {\sigma_{LEm}}}}{\sigma }} \right)$$

The two coherent light beams from the normal grating and the echelle grating exit the beam splitter and are then collected using the imaging optics. For the two coherent plane wavefronts are produced at the output whose wavevectors are inclined to the optical axis by small diffraction angle βG and βE, the spatial frequency fx of the interference fringes generated on the charge-coupled device (CCD) detector can be expressed as [22]

$${f_x} = \sigma ({\sin {\beta_G} + \sin {\beta_E}} )= 2({({\sigma - {\sigma_{LG}}} )\tan {\alpha_{LG}} + ({\sigma - {\sigma_{LEm}}} )\tan {\alpha_{LE}}} )$$

In the conventional SHS consisting of two normal gratings, the spatial frequency can be rewritten as fx= 4(σσLG)tanαLG, and when the wavenumber σ is equal to the Littrow wavenumber σLG, the spatial frequency fx is equal to zero. However, Eq. (7) clearly shows that for the grating-echelle SHRS (GESHRS) consisting of one normal grating and one echelle grating, if σLGσLEm and if the wavenumber σ is equal to σLG or σLEm, the spatial frequency is not equal to zero; this means that both σLG and σLEm cannot be the Littrow wavenumber of the GESHS. By setting fx in Eq. (7) to be equal to zero, the Littrow wavenumber of the GESHS can be calculated using

$$({{\sigma_{LGEm}} - {\sigma_{LG}}} )\tan {\alpha _{LG}} + ({{\sigma_{LGEm}} - {\sigma_{LEm}}} )\tan {\alpha _{LE}} = 0$$
where σLGEm is the Littrow wavenumber of the GESHS when the normal grating is coherent with the light beam diffracted by the m-th order of the echelle grating. By expanding the tangent function in Eq. (8) to the first order in σLG and σLEm and then solving the equation, the Littrow wavenumber σLGEm can be expressed as
$${\sigma _{LGEm}} = \frac{{{\sigma _{LG}}{\alpha _{LG}} + {\sigma _{LEm}}{\alpha _{LE}}}}{{{\alpha _{LG}} + {\alpha _{LE}}}}$$

To aid in the analysis of the spectral performance of the GESHS and to have the same calibration standard for both the normal grating and the echelle grating during construction of the instrument, it is helpful to set the Littrow wavenumber of the normal grating to be equal to the Littrow wavenumber of the p-th diffraction order of the echelle grating, and this Littrow wavenumber can be written as σLGEp = σLG = σLEp. Under this condition, the difference between the Littrow wavenumber of the GESHRS σLGEm and the Littrow wavenumber of the echelle grating σLEm can be expressed as:

$${\Delta _{GEm}} = {\sigma _{LGEm}} - {\sigma _{LEm}} = \frac{{({{\sigma_{LG}} - {\sigma_{LEm}}} ){\alpha _{LG}}}}{{{\alpha _{LG}} + {\alpha _{LE}}}} = \frac{{({{\sigma_{LEp}} - {\sigma_{LEm}}} ){\alpha _{LG}}}}{{{\alpha _{LG}} + {\alpha _{LE}}}}$$
where ΔGEm is the Littrow wavenumber difference in the m-th diffraction order between the GESHRS and the echelle grating. By substituting Eq. (2) into Eq. (10), the difference can finally be given by:
$${\Delta _{GEm}} = \frac{{({p - m} ){\alpha _{LG}}}}{{2({{\alpha_{LG}} + {\alpha_{LE}}} ){d_E}\sin {\alpha _{LE}}}}$$

Equation (11) clearly shows that ΔGEm increases proportionally as the diffraction order m decreases. By combining Eq. (2) with Eq. (11), the Littrow wavenumber σLGEm can be expressed as

$${\sigma _{LGEm}} = {\sigma _{LEm}} + {\Delta _{GEm}} = {\sigma _{LEp}} - \frac{{(p - m)}}{{2{d_E}\sin {\alpha _{LE}}}} + \frac{{({p - m} ){\alpha _{LG}}}}{{2({{\alpha_{LG}} + {\alpha_{LE}}} ){d_E}\sin {\alpha _{LE}}}}$$

When σLG = σLEp, the Littrow wavenumber σLGEm can finally be expressed more concisely as

$${\sigma _{LGEm}} = {\sigma _{LG}} - (p - m)\frac{{{\alpha _{LE}}}}{{2({{\alpha_{LG}} + {\alpha_{LE}}} ){d_E}\sin {\alpha _{LE}}}} = {\sigma _{LG}} - (p - m){\Delta _{GESHS}}$$
where ΔGESHS has a definite value when the parameters of the normal grating and the echelle grating are confirmed, and the Littrow wavenumber of the GESHS for the different orders can be calculated simply at the same time.

To avoid ambiguity in the spectral information being caused by the one-dimensional interference fringes, the normal grating in GESHS is tilted at a small angle ε to introduce an extra spatial frequency fy and thus generate two-dimensional interference fringes. The introduced spatial frequency fy of the interference fringes can be expressed as

$${f_y} = 2\sigma \varepsilon$$

For the two-dimensional interference fringes generated on the CCD detector, the intensity distribution can finally be expressed as [23]

$$\begin{aligned} I({x,y} )&= \int\limits_0^\infty {B(\sigma )} [{1 + \cos ({2\pi ({{f_x}x + {f_y}y} )} )} ]d\sigma \\ & = \int\limits_0^\infty {B(\sigma )} [{1 + \cos ({2\pi ({({({{\sigma_{LGEm}} - {\sigma_{LG}}} )\tan {\alpha_{LG}} + ({{\sigma_{LGEm}} - {\sigma_{LEm}}} )\tan {\alpha_{LE}}} )x + 2\sigma \varepsilon y} )} )} ]d\sigma \end{aligned}$$
where B(σ) is the intensity of the incident polychromatic signal light at different wavenumbers, and x and y are the displacements on the grating along the row and column directions of the detector, respectively. The two-dimensional interference fringes received at the detector can be converted into the corresponding spectral information using a two-dimensional Fourier transform.

By assuming that the width of the signal light received by the CCD detector is WL, then the effective areas illuminated by the light of the normal grating and the echelle grating can be expressed respectively as

$${W_G}\textrm{ = }\frac{{{W_L}}}{{\cos {\alpha _{LG}}}}$$
$${W_E}\textrm{ = }\frac{{{W_L}}}{{\cos {\alpha _{LE}}}}$$

By combining the grating-echelle structure with the definition of the resolution limit [24] in the two-dimensional SHS, the spectral resolution δσ of the GESHRS can be calculated to be

$${\delta _\sigma } = \frac{1}{{2({{W_G}\sin {\alpha_{LG}} + {W_E}\sin {\alpha_{LE}}} )}} = \frac{1}{{2{W_L}({\tan {\alpha_{LG}} + \tan {\alpha_{LE}}} )}}$$

The corresponding resolution power R can then be given by

$$R = \frac{\sigma }{{{\delta _\sigma }}} = 2{W_L}\sigma ({\tan {\alpha_{LG}} + \tan {\alpha_{LE}}} )$$

In each row of the detector, when it is assumed that the number of the pixels is Np, the spectral range Δσ of the GESHRS can be calculated as follows:

$${\Delta _\sigma }\textrm{ = }{N_p}{\delta _\sigma }\textrm{ = }\frac{{{N_p}}}{{2{W_L}({\tan {\alpha_{LG}} + \tan {\alpha_{LE}}} )}}$$

The detectable maximum and minimum wavenumbers for each diffraction order of the FWGE-SHRS can be calculated as σLGEmmax = σLGEm + Δσ/2 and σLGEmmin = σLGEm - Δσ/2, respectively. To ensure the integrity of the spectrum obtained, it is important to set the detectable wavenumber ranges of pairs of adjacent diffraction orders to be partially overlapping during the design process of the FWGE-SHRS; this can be expressed as σLGEmmin < σLGEm−1max.

As an important performance index of the SHS, the FOV can be improved greatly by introducing field-widening prisms into the light path, as shown in Fig. 1, based on field-widening theory [22]. To field-widen the GESHS, the respective rotation angles of the field-widening prisms PG and PE can be calculated using

$$2({{n_P}^2 - 1} )\tan {\gamma _{PG}} = {n_P}^2\tan {\alpha _{LG}}$$
$$2({{n_P}^2 - 1} )\tan {\gamma _{PE}} = {n_P}^2\tan {\alpha _{LE}}$$
where nP is the refractive index of the field-widening prism, and γPG and γPE are the rotation angles of the field-widening prisms PG and PE, respectively. The rotation angle is also equal to the minimum deviation angle of the field-widening prism, and the corresponding apex angles of these prisms can then be calculated using
$${n_P}\sin \left( {\frac{{{\alpha_{PG}}}}{2}} \right) = \sin {\gamma _{PG}}$$
$${n_P}\sin \left( {\frac{{{\alpha_{PE}}}}{2}} \right) = \sin {\gamma _{PE}}$$
where αPG and αPE are the apex angles of the field-widening prisms PG and PE, respectively. Based on the calculations of Eqs. (21)–(24), setting the field-widening prisms PG and PE before the corresponding gratings in the GESHS with the prepared rotation angles and apex angles allows the FOV to be increased by two orders [25].

2.2 Calibration theory

During the calibration process for the conventional SHRS, it is only necessary to determine one Littrow wavenumber with two known emission lines from the standard light source; then, we can solve for the corresponding parameters of the spectrometer. However, Eq. (9) shows that there are four unknown parameters of the spectrometer that must be solved, which means we must determine at least four Littrow wavenumbers with eight known emission lines in four different diffraction orders. This not only makes the experimental process very cumbersome, but also makes solution of the equation groups a very complex task.

As mentioned earlier, it is helpful to set the Littrow wavenumbers σLGEp = σLG = σLEp; this means that the parameters for the normal grating and the echelle grating can both be calculated when σLGEp is determined. Therefore, the focus of the calibration process is to determine the Littrow wavenumber of the p-th diffraction order in GESHS, denoted by σLGEp.

During the calibration process, if the two known emission lines lie within the m-th diffraction order of the GESHS, then the Littrow wavenumber σLGEm can be calculated using

$${\sigma _{LGEm}} = \frac{{({{f_2}{\sigma_1}} )- ({{f_1}{\sigma_2}} )}}{{{f_2} - {f_1}}} = \frac{{({{{{f_2}} / {{\lambda_1}}}} )- ({{{{f_1}} / {{\lambda_2}}}} )}}{{{f_2} - {f_1}}} = \frac{1}{{{\lambda _{LGEm}}}}$$
where σ1 and σ2 are the wavenumbers of the known emission lines, f1 and f2 are the spatial frequencies that correspond to the two known emission lines in the interference fringes, and λ1 and λ2 are the corresponding wavelengths of these known emission lines, respectively; λLGEm is the Littrow wavelength of the m-th diffraction order in the GESHS. The spectral resolution of the GESHS can then be calculated using
$${\delta _\sigma } = \frac{{{\sigma _{LGEm}} - {\sigma _1}}}{{{f_1}}} = \frac{{{\sigma _{LGEm}} - {\sigma _2}}}{{{f_2}}}$$

Because the echelle grating has multiple diffraction orders, the calibration process of σLGEp is divided into two cases. In the first case, the two known emission lines lie within the p-th diffraction order of the GESHS, the Littrow wavenumber σLGEp can be calculated using Eq. (25), and the Littrow wavenumbers of the normal grating and the echelle grating of the p-th diffraction order can then be calculated to be σLG = σLEp = σLGEp.

In the second case, the two known emission lines lie within the n-th (n = p−1, , pq + 1) diffraction order of the GESHS, the Littrow wavenumber σLGEn can be calculated using Eq. (25), and the corresponding spectral resolution can be calculated using Eq. (26). In this case, we require another known emission line that lies within the p-th diffraction order of the GESHS to determine σLGEp, and this can be expressed as:

$${\sigma _{LG}} = {\sigma _{LEp}} = {\sigma _{LGEp}} = {\sigma _3} + {f_3}{\delta _\sigma }$$

In the two cases described above, after the Littrow wavenumber σLGEp has been determined, based on Eq. (1) and Eq. (2), and when groove densities of 1/dG and 1/dE are known parameter, the respective Littrow angles of the normal grating and the echelle grating can be calculated using

$${\alpha _{LG}} = \arcsin \left( {\frac{1}{{2{\sigma_{LG}}{d_G}}}} \right)$$
$${\alpha _{LE}} = \arcsin \left( {\frac{p}{{2{\sigma_{LEp}}{d_E}}}} \right)$$

Based on the analysis above and Eq. (13), after the Littrow angles have been determined, the value of ΔGESHS can then be determined, and the Littrow wavenumber can be obtained for each diffraction order σLGEm using Eq. (13). According to Eq. (18), after the spectral resolution and the Littrow angles have been determined, the width of the signal light received by the CCD detector can then be calculated using

$${W_L} = \frac{1}{{2{\delta _\sigma }({\tan {\alpha_{LG}} + \tan {\alpha_{LE}}} )}}$$

The effective areas of the normal grating and the echelle grating can then be calculated using Eq. (16) and Eq. (17), respectively.

3. Experimental

3.1 Breadboard

The optical light path of the experimental breadboard is shown in Fig. 2 and the optical parameters of the components used to construct the breadboard system are listed in Table 1. In the excitation and collection part of the experimental optical path, the linear variable filter can adjust the laser power output from the solid-state 532 nm laser in a linear manner from 0 to 400 mW; the 15-mm-focal-length collimation lens L1 collimates the laser beam on the sample and collects the Raman signal light beams; the 700 nm short-pass filter (84-714, Edmund Optics) and the 532 nm long-pass filter (532-LAB-80AC-2.5, CNI) were used to filter out any useless fluorescent light and reject the laser light; the 75-mm-focal-length collimation lens L2 then converged the filtered signal light beam into the input port of a 1.5-mm-diameter silica fiber set at the focal plane of the filter.

 figure: Fig. 2.

Fig. 2. Optical light path of the experimental breadboard for the spatial heterodyne Raman spectrometer based on the field-widened grating-echelle structure.

Download Full Size | PDF

Tables Icon

Table 1. Essential parameters for the optical elements used in the experimental breadboard

In the principal FWGE-SHRS prototype, the 150-mm-focal-length lens L3 collimated the signal light emitted from the fiber output port. The core section of the FWGE-SHRS was composed of a 50.8 × 50.8 × 50.8 mm3 cubic beam splitter, an echelle grating and a normal grating (Changchun Institute of Optics, Fine Mechanics and Physics) set in the Littrow condition, and two field-widening prisms (Changchun UP Optotech (Holding) Co., Ltd) that correspond to the echelle grating and the normal grating. The imaging optics comprised a camera lens (AF-S Micro Nikkor 105 mm 1:2.8G ED, Nikon) that was used for imaging of the two-dimensional interference fringes generated by the echelle grating and the normal grating. The image was received by a CCD detector (iKon-L 936, Andor) that was cooled to −80°C to reduce thermal noise and enhance the detector’s weak signal detection capability.

During the spectral analysis process, the two-dimensional interference fringes received by the detector were transformed into a corresponding two-dimensional spatial frequency using the two-dimensional fast Fourier transform (2DFFT) function. In the restoration of the spectroscopy results, the phase correction method [26] and the background removal function [27] were used to convert the spatial frequency accurately into the Raman spectrum and separate the signal from the background noise effectively.

3.2 Calibration

The interferogram for the mercury lamp produced by the FWGE-SHRS and the corresponding calibration results are shown in Fig. 3. Figure 3(b) clearly shows that the three spatial frequencies obtained from the mercury lamp’s interferogram, as shown in Fig. 3(a), are distributed over the m = 14, 15 diffraction orders of the FWGE-SHRS. In Fig. 3(c) and Fig. 3(d), in the m = 14 diffraction order, the spatial frequencies f1 = −424 and f2 = −308 correspond to the wavelengths of 579.067 nm and 576.961 nm, respectively; in the m = 15 diffraction order, the spatial frequency f3 = 53 corresponds to the wavelength of 546.075 nm. The distribution of the two-dimensional spatial frequencies clearly corresponds to the second case mentioned in the discussion of the calibration theory.

 figure: Fig. 3.

Fig. 3. (a) Interferogram of the mercury lamp produced by the FWGE-SHRS. (b) Two-dimensional spatial frequency obtained from the 2D FFT of the interferogram. (c) Spatial frequency fx of the m = 14, 15 diffraction orders. (d) Corresponding detailed Raman spectrum of the mercury lamp after calibration.

Download Full Size | PDF

Using Eqs. (25)–(27), we are able to calculate the Littrow wavelengths of the m = 14, 15 diffraction orders to be λLGE15 = λLG = λLE15 = 546.935 nm and λLGE14 = 571.443 nm, along with the spectral resolution of δσ= 0.5434 cm−1. According to Eqs. (28) and (29), by combining the Littrow wavelength λLG = λLE15 with the groove densities of 1/dG and 1/dE, the Littrow angles for the normal grating and the echelle grating can be calculated to be αLG = 4.706° and αLE = 8.492°, respectively. The width of the signal light received by the CCD detector can be calculated using Eq. (30) to be WL = 39.724 mm, and the definite value ΔGESHS can also be determined using Eq. (13) to be ΔGESHS = 784.139 cm−1. By combining the ΔGESHS value with Eq. (13), the Littrow wavelengths can be calculated for all the diffraction orders, and the other spectral parameters are listed in Table 2.

Tables Icon

Table 2. Spectral parameters of the different diffraction orders

The results in Table 2 show that the Littrow wavenumbers satisfy the relationship σLGEmmin < σLGEm−1max, which means that the spectral ranges for each diffraction order are partially overlapped, and the overall detectable spectral range is thus continuous. By setting the wavenumber of 532 nm to be the zero wavenumber, the overall spectral range is from −43.158 cm−1 to 3422.156 cm−1 with four diffraction orders. For each pair of adjacent diffraction orders, the overlapping spectral range is 328.758 cm−1.

4. Results and Discussion

4.1 Raman spectrum detection and analysis of a mixture of organic solutions with different diffraction orders

Figure 4(a) shows the two-dimensional spatial frequency obtained from the interferogram of a mixture of organic solutions. The detailed spatial frequencies in the different diffraction orders are shown in Fig. 4(b), the parameters that describe the relationship between the diffraction order m and the spatial frequency fy are listed in Table 3, and the interval between pairs of adjacent diffraction orders is 58 times the spatial frequency fy. Each peak in the two-dimensional spatial frequency range extends in both the fx direction and the fy direction, which means the signal peak has a certain width in both directions. When compared with the MESHRS, where the interval between each pair of adjacent diffraction orders is only 3 times the spatial frequency fy [21], and the DESHRS, where the interval between each pair of adjacent diffraction orders is 15 times the spatial frequency fy [17], the larger interval for the diffraction orders in the GESHRS effectively prevents interference of the spatial frequencies fx in the different diffraction orders, and also resolves the ambiguities of the Raman peaks.

 figure: Fig. 4.

Fig. 4. (a) Two-dimensional spatial frequency obtained from the 2D FFT of the interferogram of the mixture of organic solutions. (b) Detailed spatial frequencies and transferred Raman peaks corresponding to each diffraction order m. (c) Measured Raman spectra of carbon tetrachloride, acetone, cyclohexane, and the mixture of these compounds acquired with a laser power of 180 mW and an integration time of 16 s.

Download Full Size | PDF

Tables Icon

Table 3. Relationship between the diffraction order m and the spatial frequency fy

From the conversion of the spatial frequency and the corresponding Raman spectra shown in Fig. 4(b), we clearly see that the spatial frequencies of the different diffraction orders correspond to the Raman peaks in different wavenumber ranges. Figure 4(c) shows the measured Raman spectra of carbon tetrachloride (CCl4), acetone, cyclohexane, and a mixture of these compounds as acquired using a laser power of 180 mW and an integration time of 16 s. For CCl4, the Raman peaks are mainly distributed within the detectable spectral range of the m = 15 diffraction order; the Raman peak at 218 cm−1 is assigned to mode E; the Raman peak at 315 cm−1 is assigned to mode T2; the Raman peak at 458 cm−1 is assigned to mode A1; the Raman peak at 762 cm−1 is assigned to the combination mode T2 + A1; and the Raman peak at 789 cm−1 is assigned to mode T1 [28]. For acetone and cyclohexane, the Raman peaks are mainly distributed within the detectable spectral ranges of the m = 14 and m = 12 diffraction orders. The Raman peaks at 532 cm−1, 788 cm−1, 1068 cm−1, 1222 cm−1, 1432 cm−1, 1706 cm−1, 2848 cm−1, 2923 cm−1, 2964 cm−1, and 3008 cm−1 correspond to the spectrum of acetone [29]. For cyclohexane, the Raman peaks at 803 cm−1, 2853 cm−1, and 2939 cm−1 are assigned to mode A1g; and the Raman peaks at 1030 cm−1, 1276 cm−1, 1446 cm−1, and 2924 cm−1 are assigned to mode Eg [30,31]. The high intensity Raman peaks of CCl4 were detected by the m = 15 diffraction order, and the high intensity Raman peaks of acetone and cyclohexane were mainly detected by the m = 12 diffraction order. The Raman peaks of CCl4 at 762 cm−1 and 789 cm−1 and the Raman peaks of cyclohexane at 788 cm−1 and 803 cm−1 are detected simultaneously by the m = 15 and m = 14 diffraction orders, and the Raman peak of acetone at 1706 cm−1 is detected simultaneously by the m = 14 and m = 13 diffraction orders.

As detailed in the analysis above, for the three types of measured organic solution, the intervals of some of the Raman peaks are close together and overlap with each other. For example, in the measured Raman spectrum of the mixture of organic solutions, the Raman peak of CCl4 at 789 cm−1, the Raman peak of acetone at 788 cm−1, and the Raman peak of cyclohexane at 803 cm−1 clearly overlap with each other, and the relative intensity of the mixed Raman peak is higher than the individual peaks for each organic solution. However, the overlap of some of the Raman peaks does not affect identification of the characteristics of the organic solution, and each type of organic solution can still be identified clearly. As shown in the analysis of the interferogram of the mixture and the single types of organic solutions, the FWGE-SHRS demonstrates its ability to detect and analyze the components of the mixed organic solutions over a broad spectral range and with a high spectral resolution. When compared with the small interval between the adjacent diffraction orders in the two-dimensional spatial frequency results generated by the conventional DESHRS, the large interval between adjacent diffraction orders in the two-dimensional spatial frequency results generated by the FWGE-SHRS are better for prevention of spectral interference between the different spatial frequencies; this is highly meaningful when used in measurements with multiple diffraction orders.

4.2 Raman spectral and SNR analyses of sulfur with various integration times

Figure 5(a) shows the Raman spectra of solid sulfur when measured using different integration times at the same laser power of 20 mW. Because of the flammability of solid sulfur, we used a laser power of 20 mW to study the vibrational Raman spectrum of sulfur at various integration times, which also represents a test of the measurement capability of the designed FWGE-SHRS under relatively low Raman excitation conditions at the same time. The results of the experiment show that all three characteristic Raman peaks at 153 cm−1, 218 cm−1, and 472 cm−1 were detected clearly. The Raman peaks at 153 cm−1 and 218 cm1 were assigned to the antisymmetric bond-bending mode and the symmetric bond-bending mode of the S8 molecule, respectively. The Raman peak at 472 cm−1 is caused by the symmetric bond-stretching mode of the S8 molecule [32].

 figure: Fig. 5.

Fig. 5. (a) Raman spectra of solid sulfur as measured using the FWGE-SHRS under conditions of different integration times and the same laser power of 20 mW. (b) SNR plot of the measured Raman spectra of solid sulfur under conditions of different integration times and the same laser power of 20 mW.

Download Full Size | PDF

The performance of the measured Raman spectrum is usually quantified using the signal-to-noise ratio (SNR), which can be written as

$$SNR\textrm{ = }\frac{{{I_P}}}{{RM{S_{noise}}}}$$
where Ip is the peak value of the Raman characteristic, and RMSnoise is the root-mean-square (RMS) value of the noise signal in the detailed Raman spectrum. To quantify the performance of the FWGE-SHRS in measuring the sulfur Raman spectrum, we consider the highest Raman peak at 218 cm−1 to be Ip and substitute this value into Eq. (31); we can then plot the SNR of the measured sulfur Raman spectrum versus the integration time, as shown in Fig. 5(b). As the results in Fig. 5(b) show, as the integration time increases gradually from 1 s to 18 s, the SNR of the measured sulfur Raman spectrum also increases rapidly at first, before then growing more slowly; when the integration time increases gradually from 19 s to 28 s, the SNR of the measured sulfur Raman spectrum begins to decrease and fluctuations occur. The intensity of the interferogram on the CCD detector approaches saturation when the integration time increases from 24 s to 28 s, and saturation finally occurs at 28 s, which causes the rapid decrease in the SNR. Combining Fig. 5(a) with Fig. 5(b) shows that as the integration time grows, although the peak value increases, the SNR of the spectrum initially increases and then decreases.

Furthermore, as shown in Fig. 6, we compared the Raman spectrum of sulfur as measured using a commercial instrument, the B&W Tek i-Raman Pro [33], with that measured using the designed FWGE-SHRS. Under the measurement conditions of a 0.5 s integration time and 8 mW of laser power, the spectrum measured by the i-Raman Pro has high background noise and the characteristic peaks are at least two orders of magnitude lower than those measured by the designed FWGE-SHRS under the same conditions. According to the analysis above, the FWGE-SHRS has great potential for use in Raman spectrum measurement.

 figure: Fig. 6.

Fig. 6. Comparison of the measured sulfur Raman spectrum from an i-Raman Pro instrument from B&W Tek with that from the designed FWGE-SHRS under the same conditions of an integration time of 0.5 s and 8 mW laser power.

Download Full Size | PDF

4.3 Raman spectral and SNR analyses of organic solutions with various laser powers and various concentrations

Figure 7(a) shows the Raman spectra of acetone as measured under the conditions of the same integration time of 25 s and different laser powers, and the vibrational Raman peaks are all identified well when the laser power decreases from 200 mW to 20 mW. The Raman peak at 532 cm−1 corresponds to the in-plane deformation mode of C-(=O)-C. The Raman peak at 788 cm−1 corresponds to the symmetric stretching mode of C-C(=O)-C. The Raman peak at 1068 cm−1 is caused by the rocking vibration of CH3-C(=O-). The Raman peak at 1222 cm−1 is caused by C-C stretching in the methyl group. The Raman peak at 1358 cm−1 is assigned to the CH3 umbrella mode. The Raman peak at 1432 cm−1 is assigned to the CH3 deformation vibration. The Raman peak observed at 1706 cm−1 is caused by C = O stretching. The Raman peaks at 2848 cm−1 and 2923 cm−1 are both caused by CH3 symmetric stretching, and the Raman peaks at 2964 cm−1 and 3008 cm−1 are both caused by CH3 asymmetric stretching [29].

 figure: Fig. 7.

Fig. 7. (a) Raman spectra of acetone as measured by FWGE-SHRS under conditions of the same integration time of 25 s and different laser powers. (b) Raman spectra of methanol versus different concentrations as measured by FWGE-SHRS under the conditions of the same integration time of 20 s and a laser power of 180 mW.

Download Full Size | PDF

Figure 7(b) shows the Raman spectra of methanol versus different concentrations as measured under conditions of the same integration time of 20 s and the same laser power of 180 mW. The Raman peak at 1038 cm−1 is assigned to the stretching vibration mode of -CO and the Raman peak at 1454 cm−1 is assigned to the deformation vibration mode of -CH3. The two main Raman peaks at 2835 cm−1 and 2945 cm−1 are assigned to the symmetric and antisymmetric stretching vibration modes of the -CH3 group, respectively [34]. As Fig. 7(b) shows, the vibrational Raman peaks are all identified clearly when the concentration is reduced from 99% to 30%. The detection capabilities of the FWGE-SHRS shown in this experiment are highly promising for application to the measurement of chemicals using different laser powers and chemical concentrations.

4.4 Raman spectrum detection and analysis of a mixture of inorganic powders in different containers

Figure 8 shows the measured Raman spectra of sodium sulfate (Na2SO4), calcium carbonate (CaCO3), and sodium nitrite (NaNO2), along with the measured Raman spectra of a mixture of these inorganic powders in different containers. All spectra were measured under conditions of laser power of 120 mW and an integration time of 20 s. From Fig. 8, we see that the Raman peaks for these inorganic powders are distributed within the detectable spectral ranges of the m = 15 and m = 14 diffraction orders in the designed FWGE-SHRS. In this part of experiment, the thickness of the plastic bag was 0.04 mm, and the thickness of both the glass bottle and the plastic bottle was 1 mm. All three containers were transparent.

 figure: Fig. 8.

Fig. 8. Measured Raman spectra of Na2SO4, CaCO3, NaNO2, and of a mixture of these compounds in different containers under conditions of a laser power of 120 mW and an integration time of 20 s.

Download Full Size | PDF

As shown in Fig. 8, for the Raman spectra of Na2SO4, the Raman peaks at 449 cm−1 and 467 cm−1 are assigned to the bending vibration mode v2(E); the Raman peaks at 620 cm−1, 632 cm−1, and 647 cm−1 are all assigned to the bending vibration mode v4(F2); the highest Raman peak at 993 cm−1 is assigned to the symmetric stretching mode v1(A1); the Raman peaks at 1101 cm−1, 1131 cm−1, and 1152 cm−1 are all assigned to the antisymmetric stretching mode v3(F2) [35]. For the Raman spectra of CaCO3, the Raman peaks at 154 cm−1 and 281 cm−1 are assigned to the lattice mode; the Raman peak at 711 cm−1 is assigned to the in-plane bending mode v4(E); and the Raman peak at 1085 cm−1 is assigned to the symmetric stretching mode v1(A1) [36,37]. For the Raman spectra of NaNO2, the Raman peaks at 120 cm−1 and 154 cm−1 are assigned to the lattice mode; the Raman peak at 828 cm−1 is assigned to the bending vibration mode v4(A1); and the Raman peak at 1328 cm−1 is assigned to the symmetric stretching mode v1(A1) [38].

In the Raman spectra of the mixture when measured in the three different containers, the vibration bands of Na2SO4 all remain clearly visible. Although the Raman peaks of CaCO3 and NaNO2 are relatively weaker than those of Na2SO4, they can still be identified clearly in the Raman spectra of the mixed inorganic powders. Both CaCO3 and NaNO2 have a Raman peak at 154 cm−1 caused by the lattice mode, and the overall Raman peak at 154 cm−1 in the Raman spectra of the mixture is clearly the result of superposition of the intensities of both Raman peaks of these two inorganic powders. However, we can still identify the compositions of the inorganic powders easily from their characteristic Raman peaks, e.g., the highest Raman peak for each type of inorganic powder caused by the symmetric stretching mode v1(A1). Comparison with the Raman spectra of the mixture of the organic solutions shows that all Raman peaks for these three inorganic powder types are separated well within the Raman spectrum of the mixture of inorganic powders.

To quantify the measurement performance for the Raman spectrum of the mixture when it is placed in the different containers, we take the Raman peak at 933 cm−1 as Ip and substitute it into Eq. (31) along with the measured SNRs of the mixtures in the plastic bag, the glass bottle, and the plastic bottle of 462, 415, and 291, respectively. There are no differences among the spectral resolutions and the spectral ranges obtained for the mixtures in the three types of transparent container under the same integration time and laser power conditions. However, the intensity and the SNR of the Raman spectra in the plastic bottle are clearly lower than the corresponding characteristics in the plastic bag and the glass bottle, which may be caused by the higher reflection characteristic of the thick plastic. The analysis above demonstrates the ability of the FWGE-SHRS to identify the components of mixed inorganic powders and to measure the samples when sealed in transparent containers.

5. Summary and conclusions

In this paper, we proposed the concept of a grating-echelle spatial heterodyne spectrometer structure. By combining the GESHS with field-widening prisms, a high-resolution, broadband spatial heterodyne Raman spectrometer based on a field-widened grating-echelle structure was designed and experiments were performed to demonstrate its capabilities. The mathematical expressions and the corresponding derivation process for the Littrow wavenumber and the other spectral parameters in FWGE-SHRS are presented here. The combination of the normal grating with the echelle grating provides a broadband spectral range and a high spectral resolution with measurement of multiple diffraction orders. Additionally, the GESHS not only eliminated ghost images without using the mask required in the conventional DESHS structure, but also can be field-widened to enlarge its FOV.

With the measured Raman spectra and the corresponding analyses performed for mixtures of organic solutions with four diffraction orders and inorganic powders, the designed FWGE-SHRS demonstrated its high-resolution, broadband measurement capabilities in Raman detection, which can aid in the analysis of the components of mixed target materials by measuring the Raman spectrum of the mixture. The interferogram generated by FWGE-SHRS can also avoid interference among the spatial frequencies in the different diffraction orders, and also resolved the ambiguous nature of the Raman peaks. The measured Raman spectrum of the mixed inorganic powders also demonstrates an ability to detect samples sealed in transparent containers. From the experimental results and analysis performed for solid sulfur, acetone, and methanol, we can see that the designed FWGE-SHRS can measure chemicals under conditions involving different integration times, laser powers, and concentrations. Comparison between the FWGE-SHRS and a commercial spectrometer under the same conditions also demonstrates that the FWGE-SHRS has a considerable advantage in the Raman measurement field. All the analyses presented above show that the designed FWGE-SHRS is suitable for use in high-resolution, broadband Raman spectrum measurements with a wide range of potential applications.

Funding

Jilin Province Research Projects in China (20220204091YY); National Natural Science Foundation of China (52227810, 61975255, 62075216, 62205333, U2006209); Natural Science Foundation of Shaanxi Province (2024JC-YBQN-0051).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. Z. Lang, S. Qiao, and Y. Ma, “Fabry–Perot-based phase demodulation of heterodyne light-induced thermoelastic spectroscopy,” Light Adv. Manuf. 4(3), 233–242 (2023). [CrossRef]  

2. Y. Liu, S. Qiao, C. Fang, et al., “A highly sensitive LITES sensor based on a multi-pass cell with dense spot pattern and a novel quartz tuning fork with low frequency,” Opto-Electron. Adv. 7(3), 230230 (2024). [CrossRef]  

3. G. W. Auner, S. K. Koya, C. Huang, et al., “Applications of Raman spectroscopy in cancer diagnosis,” Cancer Metastasis Rev. 37(4), 691–717 (2018). [CrossRef]  

4. J. Bonefacino, H.-Y. Tam, T. S. Glen, et al., “Ultra-fast polymer optical fibre Bragg grating inscription for medical devices,” Light: Sci. Appl. 7(3), 17161 (2018). [CrossRef]  

5. H. G. M. Edwards, J. Jehlička, and A. Culka, “Portable Raman Spectroscopy in Field Geology and Astrobiology Applications,” in Portable Spectroscopy and Spectrometry (2021), pp. 377–400. [CrossRef]  

6. D. Bersani and P. P. Lottici, “Raman spectroscopy of minerals and mineral pigments in archaeometry,” J. Raman Spectrosc. 47(5), 499–530 (2016). [CrossRef]  

7. A. M. Zheltikov, “The Raman effect in femto- and attosecond physics,” Phys.-Usp. 54(1), 29–51 (2011). [CrossRef]  

8. G. Pezzotti, “Raman spectroscopy of piezoelectrics,” J. Appl. Phys. 113(21), 211301 (2013). [CrossRef]  

9. S. Mattana, M. Mattarelli, L. Urbanelli, et al., “Non-contact mechanical and chemical analysis of single living cells by microspectroscopic techniques,” Light: Sci. Appl. 7(2), 17139 (2018). [CrossRef]  

10. G. Pezzotti, “Raman spectroscopy in cell biology and microbiology,” J. Raman Spectrosc. 52(12), 2348–2443 (2021). [CrossRef]  

11. E. C. Cull, M. E. Gehm, B. D. Guenther, et al., “Standoff Raman spectroscopy system for remote chemical detection,” Proc. SPIE 5994, 59940H (2005). [CrossRef]  

12. L.-S. Fan, L. Constantin, D.-w. Li, et al., “Ultraviolet laser photolysis of hydrocarbons for nondiamond carbon suppression in chemical vapor deposition of diamond films,” Light: Sci. Appl. 7(4), 17177 (2018). [CrossRef]  

13. W.-l. Zhang, Z.-y. Liu, H. Wang, et al., “Research status of spatial Heterodyne spectroscopy–A review,” Microchem. J. 166, 106228 (2021). [CrossRef]  

14. J. Liu, Bayanheshig, X. Qi, et al., “Backscattering Raman spectroscopy using multi-grating spatial heterodyne Raman spectrometer,” Appl. Opt. 57(33), 9735–9745 (2018). [CrossRef]  

15. Q. Chu, Y. Sun, C. Sun, et al., “Broadband, high-resolution spatial heterodyne Raman spectroscopy measurement based on a multi-Littrow-angle multi-grating,” Opt. Express 31(19), 31284–31299 (2023). [CrossRef]  

16. J. Qiu, X. Qi, X. Li, et al., “Broadband transmission Raman measurements using a field-widened spatial heterodyne Raman spectrometer with mosaic grating structure,” Opt. Express 26(20), 26106–26119 (2018). [CrossRef]  

17. J. Qiu, X. Qi, X. Li, et al., “Broadband, high-resolution Raman observations from a double-echelle spatial heterodyne Raman spectrometer,” Appl. Opt. 57(30), 8936–8941 (2018). [CrossRef]  

18. C. R. Englert, C. M. Brown, B. Bach, et al., “High-efficiency echelle gratings for MIGHTI, the spatial heterodyne interferometers for the ICON mission,” Appl. Opt. 56(8), 2090–2098 (2017). [CrossRef]  

19. K. D. Marr, C. R. Englert, J. M. Harlander, et al., “Thermal sensitivity of DASH interferometers: the role of thermal effects during the calibration of an Echelle DASH interferometer,” Appl. Opt. 52(33), 8082–8088 (2013). [CrossRef]  

20. D. Wei, Y. Zhu, J. Liu, et al., “Thermally stable monolithic Doppler asymmetric spatial heterodyne interferometer: optical design and laboratory performance,” Opt. Express 28(14), 19887–19900 (2020). [CrossRef]  

21. J. Qiu, X. Qi, X. Li, et al., “Development of a spatial heterodyne Raman spectrometer with echelle-mirror structure,” Opt. Express 26(9), 11994–12006 (2018). [CrossRef]  

22. J. M. Harlander, “Spatial heterodyne spectroscopy: interferometric performance at any wavelength without scanning,” (The University of Wisconsin-Madison, 1991), Chap2, Chap 3.

23. R. J. Bell, Introductory Fourier Transform Spectroscopy (Academic Press, 1972), Chap 3.

24. V. Saptari, Fourier Transform Spectroscopy Instrumentation Engineering, SPIE Press monograph (SPIE Optical Engineering Press, 2004), Chap 3.

25. J. M. Harlander, F. L. Roesler, C. R. Englert, et al., “Robust monolithic ultraviolet interferometer for the SHIMMER instrument on STPSat-1,” Appl. Opt. 42(15), 2829–2834 (2003). [CrossRef]  

26. C. R. Englert, J. M. Harlander, J. G. Cardon, et al., “Correction of phase distortion in spatial heterodyne spectroscopy,” Appl. Opt. 43(36), 6680–6687 (2004). [CrossRef]  

27. M. Misiti, Y. Misiti, G. Oppenheim, et al., “Wavelet toolbox user’s guide,” The Math Works Ins, 2–36 (1996).

28. D. Tuschel, “Practical group theory and Raman spectroscopy, part II: Application of polarization,” Spectroscopy (2014).

29. V. Calvino-Casilda, K. Stawicka, M. Trejda, et al., “Real-time Raman monitoring and control of the catalytic acetalization of glycerol with acetone over modified mesoporous cellular foams,” J. Phys. Chem. C 118(20), 10780–10791 (2014). [CrossRef]  

30. M. J. Pelletier, “Effects of Temperature on Cyclohexane Raman Bands,” Appl. Spectrosc. 53(9), 1087–1096 (1999). [CrossRef]  

31. K. B. Wiberg and A. Shrake, “A vibrational study of cyclohexane and some of its isotopic derivatives—I: Raman and infrared spectra and assignments of cyclohexaae and cyclohexane-d12,” Spectrochim. Acta. A. 27(7), 1139–1151 (1971). [CrossRef]  

32. J. Lin and Y.-q. Li, “Ultralow frequency Stokes and anti-Stokes Raman spectroscopy of single living cells and microparticles using a hot rubidium vapor filter,” Opt. Lett. 39(1), 108–110 (2014). [CrossRef]  

33. J. Qiu, X. Qi, X. Li, et al., “Raman measurements using a field-widened spatial heterodyne Raman spectrometer,” J. Raman Spectrosc. 50(10), 1602–1613 (2019). [CrossRef]  

34. J. P. Perchard, F. Romain, and Y. Bouteiller, “Determination of vibrational parameters of methanol from matrix-isolation infrared spectroscopy and ab initio calculations. Part 1–Spectral analysis in the domain 11 000–200 cm− 1,” Chem. Phys. 343(1), 35–46 (2008). [CrossRef]  

35. K. Ben Mabrouk, T. H. Kauffmann, H. Aroui, et al., “Raman study of cation effect on sulfate vibration modes in solid state and in aqueous solutions,” J. Raman Spectrosc. 44(11), 1603–1608 (2013). [CrossRef]  

36. H. N. Rutt and J. H. Nicola, “Raman spectra of carbonates of calcite structure,” J. Phys. C 7(24), 4522–4528 (1974). [CrossRef]  

37. G. Behrens, L. T. Kuhn, R. Ubic, et al., “Raman spectra of vateritic calcium carbonate,” Spectrosc. Lett. 28(6), 983–995 (1995). [CrossRef]  

38. C. K. Asawa and M. K. Barnoski, “Raman Spectra and Mode Frequency Shifts of Ferroelectric Sodium Nitrite at 77 and 294 K,” Phys. Rev. B 2(1), 205–213 (1970). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1.
Fig. 1. Optical layout of the experimental light path for the spatial heterodyne Raman spectrometer based on the field-widened grating-echelle structure.
Fig. 2.
Fig. 2. Optical light path of the experimental breadboard for the spatial heterodyne Raman spectrometer based on the field-widened grating-echelle structure.
Fig. 3.
Fig. 3. (a) Interferogram of the mercury lamp produced by the FWGE-SHRS. (b) Two-dimensional spatial frequency obtained from the 2D FFT of the interferogram. (c) Spatial frequency fx of the m = 14, 15 diffraction orders. (d) Corresponding detailed Raman spectrum of the mercury lamp after calibration.
Fig. 4.
Fig. 4. (a) Two-dimensional spatial frequency obtained from the 2D FFT of the interferogram of the mixture of organic solutions. (b) Detailed spatial frequencies and transferred Raman peaks corresponding to each diffraction order m. (c) Measured Raman spectra of carbon tetrachloride, acetone, cyclohexane, and the mixture of these compounds acquired with a laser power of 180 mW and an integration time of 16 s.
Fig. 5.
Fig. 5. (a) Raman spectra of solid sulfur as measured using the FWGE-SHRS under conditions of different integration times and the same laser power of 20 mW. (b) SNR plot of the measured Raman spectra of solid sulfur under conditions of different integration times and the same laser power of 20 mW.
Fig. 6.
Fig. 6. Comparison of the measured sulfur Raman spectrum from an i-Raman Pro instrument from B&W Tek with that from the designed FWGE-SHRS under the same conditions of an integration time of 0.5 s and 8 mW laser power.
Fig. 7.
Fig. 7. (a) Raman spectra of acetone as measured by FWGE-SHRS under conditions of the same integration time of 25 s and different laser powers. (b) Raman spectra of methanol versus different concentrations as measured by FWGE-SHRS under the conditions of the same integration time of 20 s and a laser power of 180 mW.
Fig. 8.
Fig. 8. Measured Raman spectra of Na2SO4, CaCO3, NaNO2, and of a mixture of these compounds in different containers under conditions of a laser power of 120 mW and an integration time of 20 s.

Tables (3)

Tables Icon

Table 1. Essential parameters for the optical elements used in the experimental breadboard

Tables Icon

Table 2. Spectral parameters of the different diffraction orders

Tables Icon

Table 3. Relationship between the diffraction order m and the spatial frequency fy

Equations (31)

Equations on this page are rendered with MathJax. Learn more.

2σLGsinαLG=1dG 
2σLEmsinαLE=mdE 
σ(sinαLG+sin(αLGβG))=1dG 
σ(sinαLE+sin(αLEβE))=mdE 
βG=2tanαLG(σσLGσ)
βE=2tanαLE(σσLEmσ)
fx=σ(sinβG+sinβE)=2((σσLG)tanαLG+(σσLEm)tanαLE)
(σLGEmσLG)tanαLG+(σLGEmσLEm)tanαLE=0
σLGEm=σLGαLG+σLEmαLEαLG+αLE
ΔGEm=σLGEmσLEm=(σLGσLEm)αLGαLG+αLE=(σLEpσLEm)αLGαLG+αLE
ΔGEm=(pm)αLG2(αLG+αLE)dEsinαLE
σLGEm=σLEm+ΔGEm=σLEp(pm)2dEsinαLE+(pm)αLG2(αLG+αLE)dEsinαLE
σLGEm=σLG(pm)αLE2(αLG+αLE)dEsinαLE=σLG(pm)ΔGESHS
fy=2σε
I(x,y)=0B(σ)[1+cos(2π(fxx+fyy))]dσ=0B(σ)[1+cos(2π(((σLGEmσLG)tanαLG+(σLGEmσLEm)tanαLE)x+2σεy))]dσ
WG = WLcosαLG
WE = WLcosαLE
δσ=12(WGsinαLG+WEsinαLE)=12WL(tanαLG+tanαLE)
R=σδσ=2WLσ(tanαLG+tanαLE)
Δσ = Npδσ = Np2WL(tanαLG+tanαLE)
2(nP21)tanγPG=nP2tanαLG
2(nP21)tanγPE=nP2tanαLE
nPsin(αPG2)=sinγPG
nPsin(αPE2)=sinγPE
σLGEm=(f2σ1)(f1σ2)f2f1=(f2/λ1)(f1/λ2)f2f1=1λLGEm
δσ=σLGEmσ1f1=σLGEmσ2f2
σLG=σLEp=σLGEp=σ3+f3δσ
αLG=arcsin(12σLGdG)
αLE=arcsin(p2σLEpdE)
WL=12δσ(tanαLG+tanαLE)
SNR = IPRMSnoise
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.