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

Compact photothermal self-mixing interferometer for highly sensitive trace detection

Open Access Open Access

Abstract

A self-mixing interferometer combined with the photothermal spectroscopy is utilized as a remarkable sensor for highly sensitive trace detection, featuring the beneficial property of a He-Ne laser with back-mounted photodiode, to the best of our knowledge, acting as an excitation laser, also as a probe laser, and even more, as a detector. Utilizing the novel implementation of the photothermal self-mixing (PTSM) interferometer with an external cavity modulation, the concentration of the sample is directly measured by the PTSM parameter extracted from the PTSM signal. The metrological qualities of the PTSM interferometer were investigated by methylene blue trace detection. For a low excitation power of 5 mW, a 7.7 nM of the limit of detection was achieved with a relative standard deviation of ∼3%. The compact and simple structure with high sensitivity has guiding significance to a robust analytical tool for the analysis of photosensitive compounds and in the detection of aquatic product hazards in aquaculture.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Photothermal spectroscopy(PTS) is the thermal state change of the sample induced by the absorption of radiation, which has been widely applied in many fields, such as analytical chemistry [13], biology and medicine [4], environmental protection [5, 6], physics for the Soret coefficients determination [7, 8] and hyperspectral imaging [9]. Typically, PTS setups have a two-laser arrangement, one is the excitation laser, which is used to heat the sample and induce the photothermal effect, and the other is the probe laser, which is used to detect the resulting changes of refractive index. Usually, this kind of PTS setups combined with an interferometer, which is used to sensitively detect the temperature-induced phase shift caused by the photothermal effect in the sample, are wonderful apparatuses for trace detection. Various types of interferometers have been applied, such as the Fabry-Perot configuration, Mach-Zehnder, and Jamin type [1013]. Recently, H. Cabrera et al. firstly applied the self-mixing(SM) effect as a pump-probe in the photothermal system [14]. The SM effect, also known as self-mixing interference, generally happens when a laser is backscattered by an external object and re-entered back into the laser cavity, and mixes with the inside light field, resulting in a modulation in the spectra and the amplitude of the laser. Unlike traditional laser interference, the self-mixing interference does not need additional optical elements for inducing interference outside the laser cavity. Instead, the self-mixing effect occurs inside the laser cavity. Since the laser has a self-coherent nature, laser feedback systems are only sensitive to the laser reflected back to the laser cavity and only a small amount of light re-injected into the cavity can cause self-mixing effect. It is nowadays a well-established technique, because of the inherent simplicity, compactness and robustness as well as the self-aligning [15], and widely applied in different fields, such as displacement and vibration [1619], absolute distance [20], angle [21], velocity [22], thickness [23], refractive index [24], motor runout [25], laser parameters [26], tomographic imaging [27], integral strain [28], acoustic fields [29], medical diagnosis [30]. The use of SM interferometry as a transducer for PTS offers a compact spatial structure for phase shift detection, and the amplitude of the resulting SM signal is proportional to the induced temperature change. Therefore, they used this system for trace detection in water samples [31]. The drawbacks of the basic two-laser arrangement PTS setups are that the energy of the selected probe laser should not be absorbed by the sample and the spot size of the probe laser must be smaller than that of the excitation laser beam to maintain the effectiveness of the parabolic approximation of thermal lens(TL) which results in complex optical path adjustment [32].

The simplest single-laser PTS setups, which are easy to build, align and manipulate. However, the major drawback is that most of them are based on far-field detection, which requires a long deflection path length of the order of a few meters [33]. And extraneous/parasitic mechanical vibrations can disturb the measurement which limited outdoor use. Additionally, high sensitivity is required for pollutants trace detection by applying low power and low-cost excitation light.

Regarding the issue above, the work aims to develop a simple and compact photothermal self-mixing (PTSM) interferometer for highly sensitive trace detection. The self-mixing interferometry is combined with the single-beam thermal lens effect. A He-Ne laser subject to a weak optical feedback regime is used to both excite the sample and simultaneously probe the thermal lens effect. The photodiode mounted in the back of the laser is used to detect PTSM signals for trace detection. An important advantage of the proposed method is that the He-Ne laser with back-mounted photodiode acts as an excitation laser, also a probe laser, even more, a detector. Having fewer components, which are only a laser, a chopper, a cell, and a mirror, allows for a highly compact experimental setup, which has an added benefit of lower cost. Moreover, the resistance against extraneous/parasitic mechanical vibrations and the sensitivity of the method are improved by external phase modulation. The metrological qualities of this novel method were demonstrated by methylene blue (MB) trace detection.

2. Principle

Figure 1 shows the diagram of the proposed photothermal self-mixing interferometer where the maximum power of the He-Ne laser is 5 mW. A chopper blade with a low-speed DC motor is used as an optical chopper to create periodic excitation. A mirror fixed on a high-precision nano-displacement piezoelectric transducer (PZT) is used to modulate the external cavity length.

 figure: Fig. 1.

Fig. 1. Diagram of the photothermal self-mixing interferometer.

Download Full Size | PDF

As shown in Fig. 2(a), the propagation of the probe laser from its minimum beam waist ω0 at the plane of z = 0 to the mirror at z = L through the sample at z = d, then reflected back to the plane of z = 0, is calculated by the fundamental Gaussian-beam solution and the ABCD law for Gaussian beams. The size of the feedback beam is enlarged, and then the feedback-coupling coefficient is reduced. The fundamental Gaussian-beam solution q(z) is given by 1/q(z) = 1/R(z) – iλ/[πω2(z)], where R(z) is the radius of the laser wave front, ω(z) is the radius of the beam spot along the direction of laser propagation and λ is the wavelength of the laser. At the plane of z = 0, the radius of the minimum spot size is ω0, q0 = iπω20/λ = ib, where b is the confocal parameter of the laser beam. At the reflected plane of z = 0 with no TL effect, the Gaussian beam is described by

$${q_{20}} = \frac{{{A_{20}}{q_0} + {B_{20}}}}{{{C_{20}}{q_0} + {D_{20}}}}$$
where
$$\left( {\begin{array}{*{20}{c}} {{A_{20}}}&{{B_{20}}}\\ {{C_{20}}}&{{D_{20}}} \end{array}} \right) = \left( {\begin{array}{*{20}{c}} 1&{2L + 2({n_s} - 1)l}\\ 0&1 \end{array}} \right)$$
is the matrix of transformation for the six-step propagation of the laser from z = 0 to z = L, through the sample with no TL effect, and finally, reflected back to z = 0. ns is the refractive index of the sample and the refractive index of air is 1. l is the length of the sample. The area of the light spot at the reflected plane of z = 0 is given by
$$\pi \omega _{20}^2 = \frac{\lambda }{b}\{{{b^2} + 4{{[{L + ({n_s} - 1)l} ]}^2}} \}$$
Then the equivalent transmission coefficient of laser beam back into the laser cavity due to beam diffusion is
$${t_d} = \frac{{\omega _0^2}}{{\omega _{20}^2}}$$

 figure: Fig. 2.

Fig. 2. Schematic for the Gaussian beam propagation of the photothermal self-mixing interferometer. (a) Gaussian beam propagation without TL effect and (b) Gaussian beam propagation with TL effect.

Download Full Size | PDF

Then the laser beam is reentered into the laser cavity and formed the SM effect. According to the three mirror Fabry-Perot(FP) model, which is usually used to descript the principle of the SM effect [34], the equivalent FP model with a sample is shown in Fig. 3.

 figure: Fig. 3.

Fig. 3. FP model with a wedge. (a) three mirror FP model and (b) equivalent FP model.

Download Full Size | PDF

As can be seen from Fig. 3(a), the optical wave is firstly exited from M1, and then transmitted and reflected by M2. The related electric fields can be written as E1 = E0exp[-i(ωt+φ)], E2 = E0exp[-i(ωt+φ+kwnDLlaser)], and E3 = r2E0exp[-i(ωt+φ+kwnDLlaser)], respectively, where Llaser represents the cavity length of He-Ne laser, E0, ω, and φ are the amplitude, angular frequency and initial phase of the electric field, respectively. M1 and M2 are two facets of the He-Ne laser. kw = 2π/λp is the wave vector while λp is the wavelength of the probe laser. r2 is the reflection coefficient of M2. nD is the refractive index of the laser cavity.

The electric field of optical wave transmitted through M2 again, which is transmitted through the sample twice and reflected by the mirror, can be written as

$${E_4} = (1 - r_2^2)t_a^2{t_d}{r_{2ext}}{E_0} \cdot exp \{{ - i[{\omega t + \varphi + {k_w}({{n_D}{L_{laser}} + 2({L + ({n_s} - 1)l} )} )} ]} \}$$
where, L = L0Lcos(2πft) is the distance from the probe laser to the mirror with an initial distance of L0, which is modulated by the PZT with a frequency of f and an amplitude of ΔL, r2ext is the reflection coefficient of the mirror, and ta is transmission coefficient of the sample due to the sample’s absorption.

As show in Fig. 3(b), the laser mirror M2, the sample and the mirror are combined into a single equivalent mirror. As reported in Ref [34]., the reflection coefficient of the equivalent mirror can be given by

$${r_{eff}} = {r_2} + (1 - r_2^2)t_a^2{t_d}{r_{2ext}}exp \{{ - i2{k_w}[{{n_D}{L_{laser}} + 2({L + ({n_s} - 1)l} )} ]} \}$$
Then the feedback coupling coefficient to the external cavity with no TL effect is expressed as
$${\kappa _{ext}} = \frac{{({1 - {r_2}^2} ){t_a}^2{t_d}{r_{2ext}}}}{{{r_2}}} = \frac{{\omega _0^2}}{{\omega _{20}^2}}\frac{{({1 - {r_2}^2} ){t_a}^2{r_{2ext}}}}{{{r_2}}}$$
Then, the power variation of the Gaussian-beam described as follows
$$\begin{array}{l} {P_{F0}} = {P_0}\left[ {1 + \frac{{{\kappa_{ext}}}}{{{L_{laser}}}}\cos \left( {\frac{{4\pi ({L + ({n_s} - 1)l} )}}{{{\lambda_p}}}} \right)} \right]\\ \;\;\;\;\; = {P_0}\left[ {1 + m\cos \left( {\frac{{4\pi ({L + ({n_s} - 1)l} )}}{{{\lambda_p}}}} \right)} \right] \end{array}$$
where P0 is the output power of the probe laser without optical feedback and m = κext/Llaser is the modulation index of the PTSM system for Gaussian beam. In the case of no TL effect,
$$m = \frac{{{\kappa _{ext}}}}{{{L_{laser}}}} = \frac{{({1 - {r_2}^2} ){t_a}^2{t_d}{r_{2ext}}}}{{{r_2}{L_{laser}}}} = \frac{{\omega _0^2}}{{\omega _{20}^2}}\frac{{({1 - {r_2}^2} ){t_a}^2{r_{2ext}}}}{{{r_2}{L_{laser}}}}$$
When the excitation laser with a Gaussian profile is periodically modulated and irradiates the sample, the sample absorbs the energy of the laser, resulting in the TL effect. The change in sample temperature can be expressed as [35]
$$\Delta T(t )\cong \frac{{\alpha {P_e}}}{{4\pi Jk}}\ln \left( {1 + \frac{{8{D_{TL}}t}}{{\omega_0^2}}} \right) = \frac{{\alpha {P_e}}}{{4\pi Jk}}\ln \left( {1 + \frac{{2t}}{{{t_c}}}} \right)$$
where Pe is the power of the excitation laser, k is the thermal conductivity of the sample, J = 4.18 is Joule’s coefficient, and α is the absorption coefficient of the sample. In our experiment, the cool time is the same as the heating time and the excitation period is 2t1. tc = ω20/(4DTL) is the thermal lens time constant, while DTL is the thermal diffusivity.

The temperature gradient change in the sample causes the refractive index gradient change, which in turn causes the phase shift during the propagation of the probe laser. The phase shift can be expressed as [35]

$$\Delta \phi (t) \cong \frac{{dn}}{{dT}}l\frac{{\alpha {P_e}}}{{2J{\lambda _p}k}}\ln \left( {1 + \frac{{2t}}{{{t_c}}}} \right)$$
where dn/dT is the refractive index of the sample changed with temperature coefficient.

The refractive index gradient change also causes the change in focal length of thermal lens, which in turn causes the change of the coupling coefficient of the feedback light in the PTSM system, that is the variation in modulation index m. Taking methylene blue in ethanol solution as an example; its thermal lens effect can be equivalent to a concave lens. As shown in Fig. 2(b), the size of the feedback beam is further enlarged, and then the modulation index is reduced. The Gaussian beam of the probe laser from its minimum beam waist ω0 at the plane of z = 0 to the mirror at z = L, through the thermal lens of focal length F, then reflected back to the plane of z = 0, is described by

$${q_2} = \frac{{A{q_0} + B}}{{C{q_0} + D}}$$
where
$$\left( {\begin{array}{*{20}{c}} A&B\\ C&D \end{array}} \right) = \left( {\begin{array}{*{20}{c}} 1&d\\ 0&1 \end{array}} \right)\left( {\begin{array}{*{20}{c}} 1&0\\ { - \frac{1}{F}}&1 \end{array}} \right)\left( {\begin{array}{*{20}{c}} 1&{L - d}\\ 0&1 \end{array}} \right)\left( {\begin{array}{*{20}{c}} 1&{L - d}\\ 0&1 \end{array}} \right)\left( {\begin{array}{*{20}{c}} 1&0\\ { - \frac{1}{F}}&1 \end{array}} \right)\left( {\begin{array}{*{20}{c}} 1&d\\ 0&1 \end{array}} \right)$$
is the matrix of transformation for the six-step propagation of the probe laser from z = 0 to z = L, through the sample with TL effect, and finally, reflected back to z = 0. The inverse focal length for the heating portion of the Mth cycle for chopped excitation is
$$\begin{array}{l} \frac{1}{{{F_{heat}}}} = \left( {\frac{{dn}}{{dT}}} \right)\frac{{2{P_e}\alpha l}}{{\pi Jk\omega _0^2}}\sum\limits_{j = 1}^M {\frac{{{t_1} \cdot {t_c}}}{{[{{t_c} + 2j(2{t_1}) + 2t} ][{{t_c} + 2j(2{t_1}) + 2t - 2{t_1}} ]}}} \\ \;\;\;\;\;\;\; + \left( {\frac{{dn}}{{dT}}} \right)\frac{{2{P_e}\alpha l}}{{\pi Jk\omega _0^2}}\frac{t}{{{t_c} + t}} \end{array}$$
while for cooling
$$\begin{array}{l} \frac{1}{{{F_{cool}}}} = \left( {\frac{{dn}}{{dT}}} \right)\frac{{2{P_e}\alpha l}}{{\pi Jk\omega _0^2}}\sum\limits_{j = 1}^M {\frac{{{t_1} \cdot {t_c}}}{{[{{t_c} + 2j(2{t_1}) + 2t} ][{{t_c} + 2j(2{t_1}) + 2t - 2{t_1}} ]}}} \\ \;\;\;\;\;\;\; + \left( {\frac{{dn}}{{dT}}} \right)\frac{{2{P_e}\alpha l}}{{\pi Jk\omega _0^2}}\frac{{{t_1} \cdot {t_c}}}{{({{t_c} + 2t} )({{t_c} + 2t - 2{t_1}} )}} \end{array}$$
where t is the time the sample is being irradiated after the Mth chopper cycle and j is the count of excitation cycles.

The difference in inverse focal lengths between the maximum heating and minimum cooling times is

$$\frac{1}{{|{\Delta F} |}} = \left( {\frac{{dn}}{{dT}}} \right)\frac{{2{P_e}\alpha l}}{{\pi Jk\omega _0^2}}\frac{{{{({{t_1}/{t_c}} )}^2}}}{{({1 + 2{t_1}/{t_c}} )({1 + 4{t_1}/{t_c}} )}}$$
With Eqs. (12) and (13), the area of the Gaussian beam at the reflected plane of z = 0 with TL effect is described by
$$\pi \omega _{_{2t}}^2 = \frac{\lambda }{b}\cdot \frac{{{A^2}{b^2} + {B^2}}}{{AD - BC}} = \frac{\lambda }{b}({{A^2}{b^2} + {B^2}} )$$
Specifically, in our PTSM system, the excitation laser is also the probe laser λ = λp. The excitation laser beam is reflected back by the mirror (see Fig. 1) embedded with the information of the phase shift and change of the coupling coefficient of the feedback light induced by the TL effect, perturbs the intra-cavity electric field to form the photothermal self-mixing effect. The variation in laser power output can be described as follows
$${P_{Ft}} = {P_0}\left[ {1 + \frac{{\omega_0^2}}{{\omega_{2t}^2}}\frac{{({1 - {r_2}^2} ){t_a}^2{r_{2ext}}}}{{{r_2}}}\cos \left( {\frac{{4\pi ({L + ({n_s} - 1)l} )}}{\lambda } + 2\Delta \phi (t)} \right)} \right]$$
With Eqs. (8) and (18), we can get
$$\frac{{\Delta {P_{F0}}}}{{\Delta {P_{Ft}}}} = \frac{{{P_{F0}} - {P_0}}}{{{P_{Ft}} - {P_0}}} = \frac{{\omega _{2t}^2\cos \left[ {\frac{{4\pi ({L + ({n_s} - 1)l} )}}{\lambda }} \right]}}{{\omega _{20}^2\cos \left[ {\frac{{4\pi ({L + ({n_s} - 1)l} )}}{\lambda } + 2\Delta \phi (t)} \right]}}$$
As b2 >>L2, b2 >>d2, and L > l, for focal length sufficiently long, we can obtain
$$\frac{{|{\Delta {P_{F0}}} |}}{{|{\Delta {P_{Ft}}} |}} = \frac{{\omega _{2t}^2}}{{\omega _{20}^2}} \approx 1 - 4L \cdot \frac{1}{{|{\Delta F} |}}\;$$
Then the fractional change of the laser intensity at the end of each excitation cycle, which is expressed as |ΔPFt(at tt1)|, relative to |ΔPFt| at beginning of each excitation cycle, expressed as |ΔPF0|, is defined as the PTSM parameter S(M)
$$S(M = \infty ) = \frac{{|{\Delta {P_{F0}}} |- |{\Delta {P_{Ft(at\;tt1)}}} |}}{{|{\Delta {P_{Ft(at\;tt1)}}} |}} \approx {K_1}\left( {\frac{{dn}}{{dT}}} \right)\frac{{{P_e}\alpha l}}{{\pi Jk\omega _0^2}}\frac{{{{({{t_1}/{t_c}} )}^2}}}{{({1 + 2{t_1}/{t_c}} )({1 + 4{t_1}/{t_c}} )}}$$
where K1 = -4L. For an external cavity length modulation of micron displacement, L0 >> ΔL, K1 is a constant.

According to the Beer law, the concentration of the sample is proportional to its absorption coefficient, α = K2C + α0, where K2 is a constant and α0 = 15.5×10−4 cm-1 is the absorption coefficient of ethanol [35]. dn/dT and k are the thermo-optic constants of the sample; and Pe, l, ω0, t1 and tc are constants of the system. We can get the linear relationship between the PTSM parameter and the concentration of the sample

$$S(M) \approx {K_1}{K_2}\left( {\frac{{dn}}{{dT}}} \right)\frac{{{P_e}l}}{{\pi Jk\omega _0^2}}\frac{{{{({{t_1}/{t_c}} )}^2}}}{{({1 + 2{t_1}/{t_c}} )({1 + 4{t_1}/{t_c}} )}}C + {K_3}$$
where K3 is a constant when the excitation time is the same,
$${K_3} = {K_1}\left( {\frac{{dn}}{{dT}}} \right)\frac{{{P_e}l}}{{\pi Jk\omega _0^2}}\frac{{{{({{t_1}/{t_c}} )}^2}}}{{({1 + 2{t_1}/{t_c}} )({1 + 4{t_1}/{t_c}} )}}{\alpha _0}$$
As expressed in Eq. (23), S(M) is linear with the concentration of the sample and the total excitation light power, and is inversely proportional to the square of the radius of the minimum spot size of the laser. Moreover, S(M) increases when the excitation frequency decreases and other parameters are constant.

The process is simulated and the simulation results are shown in Figs. 4 and 5. The corresponding parameters are shown in Table 1. The optical feedback is set in the weak optical feedback regime (the optical feedback level = 0.01).

 figure: Fig. 4.

Fig. 4. Simulation of signals for C = 0. (a) Inverse focal length, (b) PTSM signal ΔPFt, (c) and (d) Enlarged PTSM signal.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Simulation of signals for C = 8.0×10−6 mol/L. (a) Inverse focal length, (b) PTSM signal ΔPFt., (c) and (d) Enlarged PTSM signal.

Download Full Size | PDF

Tables Icon

Table 1. Simulation parameters

As shown in Fig. 4, the concentration of the sample is 0. In other words, the sample is ethanol (99.9%). The magnitude of the inverse focal length difference is almost zero (∼7.14×10−6 m-1). As the laser irradiation, the amplitude of ΔPFt shown in Fig. 4(b) remains almost constant. When the chopper blocks the laser, the amplitude of ΔPFt suddenly becomes zero until the next laser excitation cycle. At the beginning of each excitation cycle, there is no TL effect. Therefore, the amplitude of ΔPFt at this moment is |ΔPF0|. At the end moment of each excitation cycle, the amplitude of ΔPFt at this time is |ΔPFt(at tt1)|. The difference between |ΔPF0| and |ΔPFt(at tt1)| can hardly be observed and the envelope amplitude of the PTSM signal in every excitation cycle is almost a constant. The reason is that the laser power is only 5 mW and the absorption coefficient of ethanol for laser with a wavelength of 632.8 nm is quite small. The PTSM signal is mainly affected by the chopper and changes periodically.

As shown in Fig. 5, the concentration of the sample is 8.0×10−6 mol/L. K2 = 6.9662×10−6 mol/(L·cm) measured by ultraviolet-visible spectrophotometer (Duetta, fluorescence and absorbance spectrometer, HORIBA). As shown in Fig. 5(a), The variation of the inverse focal length in one excitation cycle is about 7.4×104 times that of C = 0. It can be clearly observed from Fig. 5(b) that the envelope of the PTSM signal is modulated not only by the chopper, but also by the TL effect. At the beginning of each excitation cycle, there is no TL effect. Therefore, the amplitude of ΔPFt at this moment is the maximum, as shown in Fig. 5(c). As the laser irradiation, the laser beam is gradually diffused due to the TL effect and the amplitude of ΔPFt gradually decreases. At the end moment of each excitation cycle, the amplitude of ΔPFt at this time is the minimum. When the chopper blocks the laser, the amplitude of ΔPFt suddenly becomes zero until the next laser excitation cycle.

It’s shown that the PTSM signal of each excitation cycle is almost the same when M is greater than 25. Figure 6 depicts the simulated PTSM parameters S(M = 99) for the sample concentration ranging from 0.2×10−6 mol/L to 10.0×10−6 mol/L. The heating times are 1.0 s, 0.8 s, and 0.6 s, respectively. The excitation power of the upper four curves is 5 mW, and that of the lower three curves is 1 mW. The length of the sample for the top curve is 2 cm, and that of others is 1 cm. Solid and dashed lines are the least squares linear fitting of the simulation data. It’s shown that the relationship between S(M) and the concentration of the sample is linear. In addition, with the same excitation power, longer heating time and longer length of sample lead to a larger slope of the linear fitting curve, which means higher sensitivity. Higher excitation power also leads to a larger slope of the linear fitting curve. Therefore, if we use a longer sample cell, longer heating time, and higher power laser, the sensitivity can be improved.

 figure: Fig. 6.

Fig. 6. Simulated PTSM parameters for the sample concentration ranging from 0.2×10−6 mol/L to 10.0×10−6 mol/L with different excitation frequencies(R2 is the correlation parameter).

Download Full Size | PDF

The frequency of the PTSM signal caused by the chopped excitation is

$${f_{tlF}} = \frac{1}{{2{t_1}}}$$
The self-mixing interference fringe frequency of the PTSM signal mainly caused by external cavity length modulation is
$${f_{mF}} = \frac{{4\Delta L}}{{\lambda /2}}f = \frac{{8\Delta L}}{\lambda }f$$
To avoid frequency overlapping problems, fmF must be greater than twice ftlF. Therefore, the following relation should be satisfied:
$${f_{mF}} \ge 2{f_{tlF}} \Rightarrow f \ge {f_{tlF}}\frac{\lambda }{{4\Delta L}}$$
From Eq. (26) we can see that the minimum modulation frequency of external cavity length needed is directly inversely proportional to the modulation amplitude of external cavity length and proportional to the frequency of the chopped excitation. Considering only detecting the envelope of the signal and the details of the interference fringes are not used in the system, the modulation amplitude and frequency of the external cavity only need to meet Eq. (26). To make the phase shift caused by external cavity length modulation more obvious than that caused by thermal lens effect, and make the experimental phenomenon more obvious, ΔL ≥ λ/2 is set to obtain at least one complete self-mixing interference fringe in one heating cycle. Of course, the parameter limitation of PZT, which is used for external cavity length modulation, needs to be considered in the experiment.

3. Experimental results and discussion

The experimental setup and measurement process for the photothermal self-mixing interferometer is shown in Fig. 7. The wavelength of the He-Ne laser is 632.8 nm, the radius of the minimum spot size is 0.085 cm, and the cavity length is 29.50 cm. The initial distance from the laser to the mirror is 35.00 cm. The distance from the laser to the cell is 22.00 cm. The radius of the entrance to the laser is 1.15 mm. A chopper blade with a low-speed DC motor is used as an optical chopper to create periodic excitation. The mirror is mounted on a linear stage (P-603, PI) and vibrated in a sine wave with amplitude of 2.373 μm and a frequency of 60 Hz. The PTSM signals are detected and amplified by the PD(PDA100, Thorlabs) mounted in the back of the laser. The output signals of PD are collected after 2 minutes of chopped laser excitation to get stable PTSM waveforms and converted into digital signals by a data acquisition board (USB-6361, NI) with a sampling rate of 51.2 KHz. And then the data are processed on a PC. Firstly, the PTSM signals are normalized and processed to extract the heating segments. And then the maximum and minimum value of the envelope of each segment (each laser excitation segment) are detected to calculate S(M). Seven measurements of the S(M) for the excitation cycle were performed and the average value with the standard deviation was registered.

 figure: Fig. 7.

Fig. 7. Configuration of trace detection based on the photothermal self-mixing interferometer

Download Full Size | PDF

Methylene blue is of great interest because it is frequently used to cure fish diseases in aquaculture practices and for transport around the world, which can cause harmful effects on the human body. Thus, accurate detecting approaches are vitally important for monitoring residual hazardous chemicals in fish tissues [36]. And methylene blue is also widely used as a redox indicator in analytical chemistry. Therefore we use the MB as an example for trace detection. We prepared the MB ethanol solutions in the range of 0.2×10−6 to 10.0×10−6 mol/L. The solution contained in a 1 cm quartz cuvette is used as a test sample.

Figure 8 shows the signals acquired for a sample without MB with a chopping period of 2 s and a heating time of 1 s. It has shown that the amplitude of the PTSM signal is modulated by a square wave, which is produced by the chopper and changes periodically, and the change of envelope amplitude in every heating cycle caused by TL effect can hardly be observed.

 figure: Fig. 8.

Fig. 8. Signals for C = 0. (a) The blue line is the PTSM signal and the red line represents the modulated excitation, (b) and (c) Enlarged PTSM signal.

Download Full Size | PDF

Figure 9 shows the signals acquired for a sample of the concentration with 8.0×10−6 mol/L and other parameters are the same as those in Fig. 8. It’s shown that the measured PTSM signal is the result of the SM and TL effects. The feedback light passes through the sample with thermal lens effect twice and is reflected by the mirror. It is embedded with information of phase shift caused by the refractive index gradient change due to the TL effect and modulated by periodic motion of external cavity mirror. The feedback process brings the phase shift information back into the laser cavity. The contribution of this part is shown as self-mixing interference fringes in the PTSM signal. One fringe corresponds to the phase change of 2π. Furthermore, the beam is diffused by thermal lens. The light spot size increases gradually when the laser beam traverses the sample. Then the intensity of feedback light is modulated by the TL effect. The contribution of this part is shown as the envelope amplitude modulation of the self-mixing interference fringes in the PTSM signal. It’s shown that the envelope of the PTSM signal in every heating cycle decreases gradually during the laser excitation. In addition, during the next laser excitation cycle, it decreases gradually from the maximum value to the minimum value again. The shape of the envelope of the PTSM signal in every exciting period is characteristic of the thermal processes like the photothermal signals in the previously reported by single laser PTS, which indicates that the SM signal is modulated by the TL effect. This validates the proposed PTSM theoretical model.

 figure: Fig. 9.

Fig. 9. Signals for C = 8.0×10−6 mol/L. (a) The blue line is the PTSM signal and the red line represents the modulated excitation, and (b) Enlarged PTSM signal.

Download Full Size | PDF

Figure 10 shows the relationship between the concentrations of MB in ethanol solution and the PTSM parameter S(M). It’s shown that the PTSM parameter S(M) changes linearly with the concentrations of MB in ethanol solution with the correlation parameter R2 =0.991. The relative standard deviations (RSD) are less than 3% as shown in the inset of the figure. The LOD of the proposed method was calculated as LOD = 3×SD0/Ms = 3×0.0003/(0.1170/(10−6 mol/L)) 7.7 nM, where SD0 is the standard deviation of S(M) for the blank signal and Ms is the slope of the linear fitting curve shown in Fig. 10 in red. At higher concentrations, nonlinearities are observed due to the very low SNR and an exciting power of only 5 mW. When the concentration is high, due to the sample's stronger absorption and TL diffusion, the beam feedback to the inner laser cavity is reduced to a very low value, resulting in a very small modulation index, and then the PTSM signal is drowned by noise. This makes it hard to obtain the S(M) parameter correctly. At lower concentrations, due to the stronger intensity of the feedback light and the help of external cavity modulation, the PTSM signal presents lower noise, especially compared with no modulation. This fact produces a low standard deviation in the blank PTSM signals and obtains the smaller LOD. Moreover, the linear fitting curve is linear through the whole range of concentrations enabling more reliable measurements. Although the LOD of our method is higher than that of the method by the ultra-sensitive surface-enhanced Raman spectroscopy coated with gold nanorods [36], our method is direct and convenient. This value compares favorably with those previously reported with PTS studies of MB (60 ng/mL) [37]. Moreover, our method uses much lower excitation power, shorter optical path length, and only one laser. And it is about 4 times lower compared to that determined based on the direct photothermal detection scheme in a flowing stream [38].

 figure: Fig. 10.

Fig. 10. Relationship between the concentrations of MB in ethanol solution and the PTSM parameter S(M) and the inset shows the RSD for each measurement point (n = 7).

Download Full Size | PDF

Figure 11 shows the relationship between the concentrations of MB in ethanol solution and the PTSM parameter S(M) with a different heating period. The heating times are 1.0 s and 0.6 s, respectively. Solid lines are the corresponding linear fitting curves. It has shown that the longer heating time leads to a larger slope of the linear fitting curve, which means higher sensitivity in trace detection. Therefore, if we use a longer sample cell, longer heating time, and higher power laser, the sensitivity can be improved.

 figure: Fig. 11.

Fig. 11. Relationship between the concentrations of MB in ethanol solution and the PTSM parameter S(M) with a different heating period.

Download Full Size | PDF

4. Conclusions

The paper illustrates high sensitive trace sensing by the newly developed photothermal self-mixing interferometer with an external cavity modulation. We deduced and demonstrated the linear relationship between the PTSM parameter S(M) and the concentration of sample; the capabilities of the method were demonstrated on the example of MB trace detection. With a low excitation power of 5 mW, a 7.7 nM LOD was achieved with an RSD of ∼3%. The novelty of this work is that the He-Ne laser with back-mounted photodiode acts as an excitation laser, also a probe laser, even more, a detector. This makes the sample with TL effect become a main component of the laser cavity, thus needing a very small amount of sample required for sensing. The proposed PTSM interferometer is simple, compact, and easy to operate, which can provide an application in the detection of aquatic product hazards in aquaculture, a possible application for the analysis of photosensitive compounds, or a technique as a detector for liquid chromatography.

Funding

Key Scientific Research Platforms and Projects in Guangdong Universities (2020ZDZX2055); National Natural Science Foundation of China (61705095); Natural Science Foundation of Guangdong Province (2019A1515011461); Special Projects in Key Areas of Rural Rejuvenation of Guangdong Higher Education Institutes (2019KZDZX2008); Science and Technology Program of Zhanjiang (2020A03003, 2020B01211, 2021A05042); Open Project Program of Guangdong Provincial Key Laboratory of Development and Education for Special Needs Children (TJ202001); Lingnan Normal University Natural Science Foundation (ZL1815); Overseas Scholarship Program for Elite Young and Middle-aged Teachers of Lingnan Normal University; Yanling Outstanding Yong Teacher Training Program Funded Project of Lingnan Normal University (YL20200102).

Acknowledgements

The authors thank Dr. E. E. Ramírez-Miquet and Dr. H. Cabrera for helpful discussions about the values of the thermo-optic constants of the samples.

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. M. Franko and C. D. Tran, “Analytical thermal lens instrumentation,” Rev. Sci. Instrum. 67(1), 1–18 (1996). [CrossRef]  

2. N. G. C. Astrath, F. B. G. Astrath, J. Shen, J. Zhou, K. H. Michaelian, C. Fairbridge, L. C. Malacarne, P. R. B. Pedreira, A. N. Medina, and M. L. Baesso, “Thermal-lens study of photochemical reaction kinetics,” Opt. Lett. 34(22), 3460–3462 (2009). [CrossRef]  

3. M. Franko, “Thermal lens spectrometric detection in flow injection analysis and separation techniques,” Appl. Spectrosc. Rev. 43(4), 358–388 (2008). [CrossRef]  

4. X. Guo, D. Zhang, K. Shojaei-Asanjan, K. Sivagurunathan, A. Melnikov, P. Song, and A. Mandelis, “Noninvasive in vivo glucose detection in human finger interstitial fluid using wavelength-modulated differential photothermal radiometry,” J. Biophotonics 12(7), e201800441 (2019). [CrossRef]  

5. L. Hernández Carabalí, E. Cedeño, A. Mantilla, S. Alvarado Ramírez, H. Cabrera, A. Mansanares, A. Calderón, and E. Marín, “Application of thermal lens microscopy (TLM) for measurement of Cr(VI) traces in wastewater,” J. Environ. Manage. 232, 305–309 (2019). [CrossRef]  

6. W. Jin, Y. Cao, F. Yang, and H. L. Ho, “Ultra-sensitive all-fibre photothermal spectroscopy with large dynamic range,” Nat. Commun. 6(1), 6767 (2015). [CrossRef]  

7. H. Cabrera, F. Cordido, A. Velásquez, P. Moreno, E. Sira, and S. A. López-Rivera, “Measurement of the Soret coefficients in organic/water mixtures by thermal lens spectrometry,” C. R. Mec. 341(4-5), 372–377 (2013). [CrossRef]  

8. L. C. Malacarne, N. G. C. Astrath, A. N. Medina, L. S. Herculano, M. L. Baesso, P. R. B. Pedreira, J. Shen, Q. Wen, K. H. Michaelian, and C. Fairbridge, “Soret effect and photochemical reaction in liquids with laser-induced local heating,” Opt. Express 19(5), 4047–4058 (2011). [CrossRef]  

9. R. Farahi, A. Passian, L. Tetard, and T. Thundat, “Pump–probe photothermal spectroscopy using quantum cascade lasers,” J. Phys. D: Appl. Phys. 45(12), 125101 (2012). [CrossRef]  

10. J. P. Waclawek, H. Moser, and B. Lendl, “Balanced-detection interferometric cavity-assisted photothermal spectroscopy employing an all-fiber-coupled probe laser configuration,” Opt. Express 29(5), 7794–7808 (2021). [CrossRef]  

11. J. P. Waclawek, V. C. Bauer, H. Moser, and B. Lendl, “2f-wavelength modulation Fabry-Perot photothermal interferometry,” Opt. Express 24(25), 28958–28967 (2016). [CrossRef]  

12. M. Isoda, M. Fukuma, and A. Harata, “Ultraviolet-excitation photothermal heterodyne interferometer as a micro-HPLC detector,” Anal. Sci. 35(12), 1311–1315 (2019). [CrossRef]  

13. H. Cabrera, E. Cedeño, P. Grima, E. Marín, A. Calderón, and O. Delgado, “Thermal lens microscope sensitivity enhancement using a passive Fabry–Perot-type optical cavity,” Laser Phys. Lett. 13(5), 055702 (2016). [CrossRef]  

14. H. Cabrera, I. Ashraf, F. Matroodi, E. E. Ramírez-Miquet, J. Akbar, J. J. Suárez-Vargas, J. F. Barrera Ramírez, D. Korte, H. Budasheva, and J. Niemela, “Photothermal lens technique: a comparison between conventional and self-mixing schemes,” Laser Phys. 29(5), 055703 (2019). [CrossRef]  

15. S. Donati and M. Norgia, “Overview of self-mixing interferometer applications to mechanical engineering,” in (SPIE, 2018), 13.

16. D. Guo, L. Shi, Y. Yu, W. Xia, and M. Wang, “Micro-displacement reconstruction using a laser self-mixing grating interferometer with multiple-diffraction,” Opt. Express 25(25), 31394–31406 (2017). [CrossRef]  

17. M. Usman, U. Zabit, O. D. Bernal, G. Raja, and T. Bosch, “Detection of multimodal fringes for self-mixing-based vibration measurement,” IEEE Trans. Instrum. Meas. 69(1), 258–267 (2020). [CrossRef]  

18. D. Li, Z. Zhang, Z. Huang, X. Wang, Z. Zhang, and Z. Huang, “Self-mixing interference vibration measurement based on even equivalent wavelength Fourier transform algorithm under weak feedback regime,” Opt. Eng. 59, 074101 (2020). [CrossRef]  

19. K. Zhu, B. Guo, Y. Lu, S. Zhang, and Y. Tan, “Single-spot two-dimensional displacement measurement based on self-mixing interferometry,” Optica 4(7), 729–735 (2017). [CrossRef]  

20. M. Veng, J. Perchoux, and F. Bony, “Fringe disappearance in self-mixing interferometry laser sensors: model and application to the absolute distance measurement scheme,” IEEE Sens. J. 19(14), 5521–5528 (2019). [CrossRef]  

21. C. Wang, X. Fan, Y. Guo, H. Gui, H. Wang, J. Liu, B. Yu, and L. Lu, “Full-circle range and microradian resolution angle measurement using the orthogonal mirror self-mixing interferometry,” Opt. Express 26(8), 10371–10381 (2018). [CrossRef]  

22. R. Atashkhooei, E. E. Ramírez-Miquet, R. d, C. Moreira, A. Quotb, S. Royo, and J. Perchoux, “Optical feedback flowmetry: impact of particle concentration on the signal processing method,” IEEE Sens. J. 18(4), 1457–1463 (2018). [CrossRef]  

23. L. Xu, S. Zhang, Y. Tan, and L. Sun, “Simultaneous measurement of refractive-index and thickness for optical materials by laser feedback interferometry,” Rev. Sci. Instrum. 85(8), 083111 (2014). [CrossRef]  

24. T. Hieta, M. Merimaa, M. Vainio, J. Seppä, and A. Lassila, “High-precision diode-laser-based temperature measurement for air refractive index compensation,” Appl. Opt. 50(31), 5990–5998 (2011). [CrossRef]  

25. R. Atashkhooei, J. C. Urresty, S. Royo, J. R. Riba, and L. Romeral, “Runout tracking in electric motors using self-mixing interferometry,” IEEE/ASME Trans. Mechatron. 19(1), 184–190 (2014). [CrossRef]  

26. Y. Gao, Y. Yu, J. Xi, and Q. Guo, “Simultaneous measurement of vibration and parameters of a semiconductor laser using self-mixing interferometry,” Appl. Opt. 53(19), 4256–4263 (2014). [CrossRef]  

27. T. Mohr, S. Breuer, G. Giuliani, and W. Elsäßer, “Two-dimensional tomographic terahertz imaging by homodyne self-mixing,” Opt. Express 23(21), 27221–27229 (2015). [CrossRef]  

28. M. Dabbicco, A. Intermite, and G. Scamarcio, “Laser-self-mixing fiber sensor for integral strain measurement,” J. Lightwave Technol. 29(3), 335–340 (2011). [CrossRef]  

29. K. Bertling, J. Perchoux, T. Taimre, R. Malkin, D. Robert, A. D. Rakić, and T. Bosch, “Imaging of acoustic fields using optical feedback interferometry,” Opt. Express 22(24), 30346–30356 (2014). [CrossRef]  

30. J. Perchoux, A. Quotb, R. Atashkhooei, F. J. Azcona, E. E. Ramírez-Miquet, O. Bernal, A. Jha, A. Luna-Arriaga, C. Yanez, and J. Caum, “Current developments on optical feedback interferometry as an all-optical sensor for biomedical applications,” Sensors 16(5), 694 (2016). [CrossRef]  

31. H. Cabrera, E. E. Ramírez-Miquet, J. J. Suarez-Vargas, J. F. B. Ramírez, D. Korte, and J. J. Niemela, “Pump-probe photothermal self-mixing system for highly sensitive trace detection,” IEEE Sens. J. 19(7), 2547–2552 (2019). [CrossRef]  

32. M. Franko and C. D. Tran, “Thermal lens spectroscopy,” Encyclopedia of analytical chemistry, 1249–1279 (2010).

33. A. K. Malik and W. Faubel, “Photothermal and light emitting diodes as detectors for trace detection in capillary electrophoresis,” Chem. Soc. Rev. 29(4), 275–282 (2000). [CrossRef]  

34. M. Koelink, M. Slot, F. De Mul, J. Greve, R. Graaff, A. Dassel, and J. Aarnoudse, “Laser Doppler velocimeter based on the self-mixing effect in a fiber-coupled semiconductor laser: theory,” Appl. Opt. 31(18), 3401–3408 (1992). [CrossRef]  

35. J. Stone, “Measurements of the absorption of light in low-loss liquids,” J. Opt. Soc. Am. 62(3), 327–333 (1972). [CrossRef]  

36. T. Xu, X. Wang, Y. Huang, K. Lai, and Y. Fan, “Rapid detection of trace methylene blue and malachite green in four fish tissues by ultra-sensitive surface-enhanced Raman spectroscopy coated with gold nanorods,” Food Control 106, 106720 (2019). [CrossRef]  

37. M. A. Proskurnin, E. V. Ageeva, V. V. Senyuta, N. V. Orlova, A. V. Fokin, O. B. Ovchinnikov, and S. V. Egerev, “Comparison of optoacoustic and photothermal-lens determination of lipopolysaccharides,” Talanta 81(1-2), 377–384 (2010). [CrossRef]  

38. B. Chouhan and P. K. Dasgupta, “Direct photothermal measurement of optical absorption in a flow system,” Anal. Chem. 91(4), 2923–2931 (2019). [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 (11)

Fig. 1.
Fig. 1. Diagram of the photothermal self-mixing interferometer.
Fig. 2.
Fig. 2. Schematic for the Gaussian beam propagation of the photothermal self-mixing interferometer. (a) Gaussian beam propagation without TL effect and (b) Gaussian beam propagation with TL effect.
Fig. 3.
Fig. 3. FP model with a wedge. (a) three mirror FP model and (b) equivalent FP model.
Fig. 4.
Fig. 4. Simulation of signals for C = 0. (a) Inverse focal length, (b) PTSM signal ΔPFt, (c) and (d) Enlarged PTSM signal.
Fig. 5.
Fig. 5. Simulation of signals for C = 8.0×10−6 mol/L. (a) Inverse focal length, (b) PTSM signal ΔPFt., (c) and (d) Enlarged PTSM signal.
Fig. 6.
Fig. 6. Simulated PTSM parameters for the sample concentration ranging from 0.2×10−6 mol/L to 10.0×10−6 mol/L with different excitation frequencies(R2 is the correlation parameter).
Fig. 7.
Fig. 7. Configuration of trace detection based on the photothermal self-mixing interferometer
Fig. 8.
Fig. 8. Signals for C = 0. (a) The blue line is the PTSM signal and the red line represents the modulated excitation, (b) and (c) Enlarged PTSM signal.
Fig. 9.
Fig. 9. Signals for C = 8.0×10−6 mol/L. (a) The blue line is the PTSM signal and the red line represents the modulated excitation, and (b) Enlarged PTSM signal.
Fig. 10.
Fig. 10. Relationship between the concentrations of MB in ethanol solution and the PTSM parameter S(M) and the inset shows the RSD for each measurement point (n = 7).
Fig. 11.
Fig. 11. Relationship between the concentrations of MB in ethanol solution and the PTSM parameter S(M) with a different heating period.

Tables (1)

Tables Icon

Table 1. Simulation parameters

Equations (26)

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

q 20 = A 20 q 0 + B 20 C 20 q 0 + D 20
( A 20 B 20 C 20 D 20 ) = ( 1 2 L + 2 ( n s 1 ) l 0 1 )
π ω 20 2 = λ b { b 2 + 4 [ L + ( n s 1 ) l ] 2 }
t d = ω 0 2 ω 20 2
E 4 = ( 1 r 2 2 ) t a 2 t d r 2 e x t E 0 e x p { i [ ω t + φ + k w ( n D L l a s e r + 2 ( L + ( n s 1 ) l ) ) ] }
r e f f = r 2 + ( 1 r 2 2 ) t a 2 t d r 2 e x t e x p { i 2 k w [ n D L l a s e r + 2 ( L + ( n s 1 ) l ) ] }
κ e x t = ( 1 r 2 2 ) t a 2 t d r 2 e x t r 2 = ω 0 2 ω 20 2 ( 1 r 2 2 ) t a 2 r 2 e x t r 2
P F 0 = P 0 [ 1 + κ e x t L l a s e r cos ( 4 π ( L + ( n s 1 ) l ) λ p ) ] = P 0 [ 1 + m cos ( 4 π ( L + ( n s 1 ) l ) λ p ) ]
m = κ e x t L l a s e r = ( 1 r 2 2 ) t a 2 t d r 2 e x t r 2 L l a s e r = ω 0 2 ω 20 2 ( 1 r 2 2 ) t a 2 r 2 e x t r 2 L l a s e r
Δ T ( t ) α P e 4 π J k ln ( 1 + 8 D T L t ω 0 2 ) = α P e 4 π J k ln ( 1 + 2 t t c )
Δ ϕ ( t ) d n d T l α P e 2 J λ p k ln ( 1 + 2 t t c )
q 2 = A q 0 + B C q 0 + D
( A B C D ) = ( 1 d 0 1 ) ( 1 0 1 F 1 ) ( 1 L d 0 1 ) ( 1 L d 0 1 ) ( 1 0 1 F 1 ) ( 1 d 0 1 )
1 F h e a t = ( d n d T ) 2 P e α l π J k ω 0 2 j = 1 M t 1 t c [ t c + 2 j ( 2 t 1 ) + 2 t ] [ t c + 2 j ( 2 t 1 ) + 2 t 2 t 1 ] + ( d n d T ) 2 P e α l π J k ω 0 2 t t c + t
1 F c o o l = ( d n d T ) 2 P e α l π J k ω 0 2 j = 1 M t 1 t c [ t c + 2 j ( 2 t 1 ) + 2 t ] [ t c + 2 j ( 2 t 1 ) + 2 t 2 t 1 ] + ( d n d T ) 2 P e α l π J k ω 0 2 t 1 t c ( t c + 2 t ) ( t c + 2 t 2 t 1 )
1 | Δ F | = ( d n d T ) 2 P e α l π J k ω 0 2 ( t 1 / t c ) 2 ( 1 + 2 t 1 / t c ) ( 1 + 4 t 1 / t c )
π ω 2 t 2 = λ b A 2 b 2 + B 2 A D B C = λ b ( A 2 b 2 + B 2 )
P F t = P 0 [ 1 + ω 0 2 ω 2 t 2 ( 1 r 2 2 ) t a 2 r 2 e x t r 2 cos ( 4 π ( L + ( n s 1 ) l ) λ + 2 Δ ϕ ( t ) ) ]
Δ P F 0 Δ P F t = P F 0 P 0 P F t P 0 = ω 2 t 2 cos [ 4 π ( L + ( n s 1 ) l ) λ ] ω 20 2 cos [ 4 π ( L + ( n s 1 ) l ) λ + 2 Δ ϕ ( t ) ]
| Δ P F 0 | | Δ P F t | = ω 2 t 2 ω 20 2 1 4 L 1 | Δ F |
S ( M = ) = | Δ P F 0 | | Δ P F t ( a t t t 1 ) | | Δ P F t ( a t t t 1 ) | K 1 ( d n d T ) P e α l π J k ω 0 2 ( t 1 / t c ) 2 ( 1 + 2 t 1 / t c ) ( 1 + 4 t 1 / t c )
S ( M ) K 1 K 2 ( d n d T ) P e l π J k ω 0 2 ( t 1 / t c ) 2 ( 1 + 2 t 1 / t c ) ( 1 + 4 t 1 / t c ) C + K 3
K 3 = K 1 ( d n d T ) P e l π J k ω 0 2 ( t 1 / t c ) 2 ( 1 + 2 t 1 / t c ) ( 1 + 4 t 1 / t c ) α 0
f t l F = 1 2 t 1
f m F = 4 Δ L λ / 2 f = 8 Δ L λ f
f m F 2 f t l F f f t l F λ 4 Δ L
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.