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

In situ detection of electrochemical reaction by weak measurement

Open Access Open Access

Abstract

In the field of electrochemical energy storage systems, the use of in situ detection technology helps to study the mechanism of electrochemical reaction. Our group has previously in situ detected the electrochemical reaction in vanadium flow batteries by total internal reflection (TIR) imaging. In order to further improve the detection resolution, in this study, the weak measurement (WM) method was introduced to in situ detect the electrochemical reaction during the linear sweep voltammetry or the cyclic voltammetry tests with quantitative measurement of the absolute current density, which lays a foundation for replacing the TIR for two-dimensional imaging of electrochemical reactions in vanadium flow batteries, oxygen/hydrogen evolution reaction, surface treatments, electrochemical corrosion and so on.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Due to energy shortages and environmental issues, the demand for renewable energy is becoming stronger. Electrochemical energy storage systems, as a method to solve the inherent intermittent problems of renewable energy including solar energy, wind energy, etc., have attracted considerable attention in recent years [1,2]. Among the electrochemical energy storage systems, vanadium redox flow batteries [3,4] show the promising prospect of industrialization with its advantages of long cycle life, environmentally friendly and independently tunable power and energy capacity.

Researchers pursue vanadium flow batteries with high electrochemical performance, however, the traditional methods for detecting electrochemical performance, such as linear sweep voltammetry (LSV) [5], cyclic voltammetry (CV) [6], electrochemical impedance spectroscopy (EIS) [7], etc., can only provide average information of the whole electrode. But the spatial distribution difference of the electrode affects the performance of energy efficiency, charge-discharge capacity, polarization and so on [8,9]. The electrochemistry distribution detection methods like potential probe [10], shunt resistor [11], printed circuit board [9] affect the battery. Therefore, it is necessary to find a method which can give spatial distribution of electrochemistry performance with little interference.

Optical refractive index detection methods—including surface plasmon resonance (SPR), total internal reflection (TIR), weak measurement (WM), and so on—has the advantages of small interference to analytes, high spatial resolution, high sensitivity and high throughput. The microscopic detection technology proposed by the Tao’s group combines the SPR method with electrochemical measurement [12,13], which achieved the same function as CV. And from each point on the result graph, a CV curve can be obtained, so the specific distribution can be obtained. Since the method was based on a microscopic light path, the light path needed to be updated to measure the macroscopic range of the electrode. Moreover, the gold film used in the SPR method not only had low durability in acid electrolyte and redox process, but also produced Faraday current interfering with the electrochemical process, thus limiting the application of the SPR method. Due to the various problems of the sensing film, our group used the TIR without sensing film to achieve a wide range of electrochemical measurements in the vanadium battery system [14,15]. But the resolution, sensitivity and signal-to-noise ratio of the TIR were worse than those of the SPR, so it is necessary to have a better performance method without sensing film.

The WM method was first proposed by Aharonov, Albert and Vaidman in 1988 [16]. This method not only provided a more in-depth explanation for quantum physics [17,18], but also was widely used in the detection of weak signals based on weak-value amplification effect, showing its potential in precise measurement [1922]. Our group has introduced the WM to the field of biomedical testing for the first time in 2015, measured the state of biomolecules and the biological reaction process, and obtained very impressive experimental results [2325]. We compared the performances of the WM and the SPR [26], and found that the resolution of the WM without coating has the same order of magnitude as the SPR. So it is very suitable for electrochemical measurement.

In this work, as far as we know, this is the first time that the WM method has been applied to electrochemical in situ detection, which has proved the feasibility of its application. The calibration experiment verified that its resolution was higher than that of the TIR, which laid the foundation for the subsequent replacement of the TIR in two-dimensional imaging measurement. Therefore, we believe that the WM method has great application prospects in electrochemical reaction measurement in vanadium flow batteries, oxygen/hydrogen evolution reaction [27,28], surface treatments [29], electrochemical corrosion [30] and so on.

2. Experimental setup and theory

2.1 Materials

Vanadyl sulfate (VOSO4·3.5H2O, 99% purity) was purchased from Shenyang Haizhongtian Fine Chemical Co., Ltd. (Shenyang, China). Sulfuric acid (H2SO4) was purchased from Dongguan Dongjiang Chemical Reagent Co., Ltd. (Dongguan, China). Prism (Chinese ZF6 glass, 12 × 12 × 12 mm) was ordered from Fuzhou Alpha Optics Co., Ltd. (Fuzhou, China). All other chemical reagents were purchased from Shenzhen Tianxiang Huabo Co., Ltd. (Shenzhen, China).

2.2 Experimental setup

The schematic diagram of the electrochemical WM system in this work was shown in Fig. 1(a). The incident light from a Superluminescent diode (SLD, A in Fig. 1(a), output with a fiber, 5 mW, IPSDS0803, Inphenix, Wuhan, China) centered at 830 nm, collimated by an achromatic lens (B in Fig. 1(a), f = 10 mm, GCL-010661, Daheng Optics Inc., Beijing, China), propagated through a linear polarizer (C in Fig. 1(a), extinction ratio of 100,000:1, LPVIS050-MP2, Thorlabs Inc., Shanghai, China). The light beam then entered a self-made electrochemical detection module (D in Fig. 1(a)) and was reflected by a prism in the module. The reflected light then passed through a Soleil-Babinet compensator (SBC, E in Fig. 1(a), SBC-IR, Thorlabs Inc., Shanghai, China), which was used for continuous phase adjustment and controlling the measuring range. The light beam then passed through another linear polarizer (F in Fig. 1(a), the same as C in Fig. 1(a)), which was almost perpendicular to the first polarizer to realize weak measurement, and was focused by another achromatic lens (G in Fig. 1(a), the same as B in Fig. 1(a)), finally collected by a spectrograph (H in Fig. 1(a), input with a fiber, HR2000, Ocean Optics, Shanghai, China). Figure 1(b) is a photograph of the electrochemical WM system in this work.

 figure: Fig. 1.

Fig. 1. Electrochemical weak measurement (WM) system. (a) Schematic of the electrochemical WM system. A: Superluminescent diode (SLD); B: achromatic lens (L1); C: linear polarizer (P1); D: self-made electrochemical detection module (EDM); E: Soleil-Babinet compensator (SBC); F: linear polarizer (P2); G: achromatic lens (L2); H: spectrograph. (b) A photograph of the electrochemical WM system.

Download Full Size | PDF

2.3 Self-made electrochemical detection module

The expanded view of the self-made electrochemical detection module in Fig. 1(a) was shown in Fig. 2(a). A prism (D3 in Fig. 2(a)) was fixed with a prism holder (D2 in Fig. 2(a)), then it was pressed against an electrochemical reaction cell (D4 in Fig. 2(a)), which was surrounded by a water bath pool (D8 in Fig. 2(a)) to reduce the interference caused by outside temperature. Finally, the prism, the electrochemical reaction cell and the water bath pool were combined by fixators (D1 and D9 in Fig. 2(a)) with screws. In the electrochemical reaction cell, a graphite plate with insulating tape (D5 in Fig. 2(a)), a graphite rod (D6 in Fig. 2(a)) and a saturated calomel electrode (SCE, D7 in Fig. 2(a)) used as working electrode, counter electrode and reference electrode respectively. The three electrodes were partially immersed in the electrolyte (0.1 M V(IV) and 2 M H2SO4). In the electrochemical reaction cell, tetravalent vanadium ions (V(IV)) and pentavalent vanadium ions(V(V)) converted into each other. The change in the valence state led to the change of the concentrations of oxidized (V(V)) and reduced (V(IV)) species, then the refractive index changed as well [31]. The change of refractive index can be detected by a high-resolution [23,26,32] weak measurement system (the sensitivity is 2795 nm∕RIU and the resolution is $8.5 \times 10^{-7}$ RIU in this work) without coating. Figure 2(c) was a schematic of them combined together. Figure 2(c) was a photograph of the self-made electrochemical detection module.

 figure: Fig. 2.

Fig. 2. The self-made electrochemical detection module (EDM, D in Fig. 1(a)). (a) Expanded schematic view of the EDM, D1: fixator; D2: prism holder; D3: prism; D4: electrochemical reaction cell; D5: graphite plate with insulating tape (working electrode, WE); D6: graphite rod (counter electrode, CE); D7: saturated calomel electrode (reference electrode, RE); D8: water bath pool; D9: fixator. (b) Combined schematic view of the EDM. (c) Photograph o of the EDM. (d) Photograph of graphite plate with insulating tape (working electrode, WE).

Download Full Size | PDF

2.4 Theory

The WM systems were used to explore about quantum physics [17,18] or detected some physical quantities [1922]. Our group introduced the WM systems to measure the state of biomolecules and the biological reaction process [2325]. In this work, we used the WM system to detect electrochemical reaction. The theoretical basis is as follows:

In the electrochemical reaction cell of the electrochemical detection module (EDM), V(IV) and V(V) converted into each other. The detection signal collected by the WM system was contributed by the concentration variation of the oxidized (V(V)) and reduced (V(IV)) species on the surface between the graphite plate and the prism in the EDM. Considering this redox reaction V(IV) ↔ V(V), the scanning potential [33] is:

$$E({\rm t}) = E_i-vt$$

If the charge transfer rate of the reaction is fast enough, the electrode reaction can be described by the Nernst equation:

$$\displaystyle{{C_O(0,t)} \over {C_R(0,t)}} = \exp [\displaystyle{{nF} \over {RT}}(E_i-vt-E^\theta )]$$
where CO (0, t) and CR (0, t) are the concentrations of the oxidized (V(V)) and reduced (V(IV)) species respectively at the measurement point. n is the number of electrons transferred in the redox reaction, F is Faraday constant, R is Boltzmann's constant, T is temperature, Ei is the initial potential, v is the scanning rate of potential, t is the reaction time, and ${E^\theta }$ is the standard electrode potential.

The diffusion equation can be used to describe the real-time variation of electrolyte concentration:

$$\frac{{\partial {C_O}(z,t)}}{{\partial t}} = {D_o}\frac{{{\partial ^2}{C_O}(z,t)}}{{{\partial ^2}z}}$$
Then the real-time concentrations of the oxidized and reduced species are respectively:
$$C_O(0,t) = C_O^0 -[nF(\pi D_O)^{1/2}]^{-1}\int_0^t {i({t}^{\prime}){(t-{t}^{\prime})}^{1/2}} d{t}^{\prime}$$
$${C_R}(0,t) = C_R^0 + {[nF{(\pi {D_R})^{1/2}}]^{ - 1}}\int_0^t {i(t^{\prime}){{(t - t^{\prime})}^{1/2}}} dt^{\prime}$$
where $C_O^0$ and $C_R^0 $ are the initial concentrations of the oxidized (V(V)) and reduced (V(IV)) species respectively, DO and DR are the diffusion coefficients of the oxidized (V(V)) and reduced (V(IV)) species respectively, and $i(t^{\prime})$ is the current density.

Similar to the previous work [12,14], considering the detection depth, the response of the WM system is:

$$\lambda (t) = K[{ {{\alpha_O}{C_O}(z,t)} |_{z = 0}} + { {{\alpha_R}{C_R}(z,t)} |_{z = 0}}]$$
where $\lambda (t)$ is the change value of the centroid wavelength shift. In the WM system, the two-peak spectrum are obtained by the spectrograph (H in Fig. 1(a)). When the refractive index of the electrolyte changes, the phase of the incident light will change, and the centroid wavelength of the two-peak spectrum will shift, as shown in Fig. 3(a). The relationship between the centroid wavelength shift of the two-peak spectrum and the phase is [26] (see Supplement 1, Section 1 for details):
$$\lambda \textrm{ = } - \frac{{2\pi k{{(\delta \lambda )}^2}}}{{{\lambda _0}}}{\mathop{\rm Im}\nolimits} {A_W} ={-} \frac{{4\pi k{{(\delta \lambda )}^2}\gamma \sin (a{\chi ^2} + \Delta )}}{{{\lambda _0}(1 + {\gamma ^2} - 2\gamma \cos (a{\chi ^2} + \Delta ))}}$$

 figure: Fig. 3.

Fig. 3. (a) The centroid wavelength shift of two-peak spectrum obtained by spectrograph. (b) Time-varying centroid wavelength shift based on the data in (a) and${\lambda _{pa}}$ is the point with the most drastic centroid wavelength shift change. (c) Denoising result by smoothing the centroid wavelength shift variations in (b) and$\Delta {\lambda _{\textrm{pa}}}$corresponds to${\lambda _{pa}}$in (b). (d) Curves by deconvolution calculation of the centroid wavelength shift variations in (c) and the peak current ipa is the result obtained by the deconvolution of$\Delta {\lambda _{\textrm{pa}}}$in (c).

Download Full Size | PDF

${\alpha _O}$ and ${\alpha _R}$ are the refractive index changes per unit concentration of the oxidized (V(V)) and reduced (V(IV)) species respectively. K is the change of the centroid wavelength shift caused by the per unit refractive index.

Then the centroid wavelength shift deduced from the above can be obtained as follow:

$$\lambda (t) = \lambda ({t_0}) + K({\alpha _R}D_R^{ - 1/2} - {\alpha _O}D_O^{ - 1/2}){(nF{\pi ^{1/2}})^{ - 1}}\int_0^t {i(t^{\prime})} {(t - t^{\prime})^{ - 1/2}}dt^{\prime}$$
where the centroid wavelength shift at the initial time is: $\lambda ({t_0}) = K({\alpha _O}C_O^0 - {\alpha _R}C_R^0)$ . Define a parameter: $b = K({\alpha _R}D_R^{ - 1/2} - {\alpha _O}D_O^{ - 1/2})$.

Then the centroid wavelength shift variation is:

$$\Delta \lambda (t) = \lambda (t) - \lambda ({t_0}) = \frac{b}{{nF}}\int_0^t {i(t^{\prime})} {[\pi (t - t^{\prime})]^{ - 1/2}}dt^{\prime}$$

Figure 3(b) displays the time-varying centroid wavelength shift $\lambda (t)$ and ${\lambda _{pa}}$ is the point with the most drastic centroid wavelength shift change. By subtracting the initial centroid wavelength shift at 0 V from all the subsequent centroid wavelength shift ($\Delta \lambda (t){\rm = }\lambda (t)-\lambda (t_0)$), the time-varying centroid wavelength shift variations $\Delta \lambda (t)$ is denoised by the “smooth” function in MATLAB R2017a. The denoised result is plotted in Fig. 3(c) and $\Delta {\lambda _{\textrm{pa}}}$ corresponds to ${\lambda _{pa}}$ in Fig. 3(b).

Rewrite Eq. (9) as:

$$\frac{{nF}}{b}\Delta \lambda (t) = \int_0^t {i(t^{\prime})} {[\pi (t - t^{\prime})]^{ - 1/2}}dt^{\prime} = convolution[i(t),{(\pi t)^{ - 1/2}}]$$

Therefore, the current density can be derived as:

$$i(t) = deconvolution[\frac{{nF}}{b}\Delta \lambda (t),{(\pi t)^{ - 1/2}}]$$

The quantitative relationship between the current density of electrochemical reaction and the centroid wavelength shift variation is established by Eq. (11). The derived curve in Fig. 3(d) shows the relative current densities along with the potential and the peak current ipa is the result obtained by the deconvolution of $\Delta {\lambda _{\textrm{pa}}}$ in Fig. 3(b).

2.5 Calibration experiment

In order to obtain the current density, these parameters (n, F, b, π) and variables ($\Delta \lambda (t)$, t) should be known according to Eq. (11). Among them, n, F and π were known, $\Delta \lambda (t)$ and t were obtained by the WM. The parameter b was the only unknown quantity, which was about the diffusion coefficients of the oxidized (V(V)) and reduced (V(IV)) species (DO, DR) and the refractive index changes per unit concentration of the oxidized (V(V)) and reduced (V(IV)) species (${\alpha _O}$, ${\alpha _R}$). As b was unknown, we conducted the calibration experiment to obtain the current density with a physical unit.

A working electrode with known reaction area was needed to obtain the current density because the electrochemical workstation can only provide total current. A graphite plate with smooth surface was used as the working electrode for calibrating. Only one interface of the graphite plate (area: 1×1 cm2) contacted with the prism and the others were wrapped by insulating tapes to avoid contact with the electrolyte, as shown in Fig. 2(d). Since the electrochemical reaction only occurred on the exposed interface, the current density can be obtained by dividing the current by the area assuming that the reaction on the interface was uniform [14]. Because the two contact surfaces between the graphite plate and the prism were not ideally smooth, the gap between them will be immersed by the electrolyte. But the electrolyte in the gap cannot be refreshed effectively, so we adopted the linear sweep voltammetry (LSV) to reduce the electrolyte renewal difference by shortening time instead of the CV [14].

A series of scan rate was set in the experiment in order to tune the current density. The current densities obtained by dividing the electrochemical workstation (EW) currents by the area (1×1 cm2) should be equal to the relative current densities obtained by the WM. Thus, the different peak oxidation current densities of the LSV at different scan rates were obtained to achieve calibration, that is, the relative current density obtained by the WM can be converted to the current density with a physical unit (mA cm-2).

3. Results and discussion

3.1 Linear detection range

According to Eq. (8), the phase change of the light can be measured by detecting the centroid wavelength shift of the emission spectrum [26]. The phase change and the centroid wavelength shift was taken as the abscissa and the ordinate respectively. The relationship between the two was shown in Fig. 4. It can be seen from Fig. 4 that the centroid wavelength shift of the emission spectrum was not a uniform change process with the continuous increase of the phase change. When the phase increased, the centroid wavelength shift changed slowly until it reached a minimum, increases suddenly and rapidly, and then slowly decreased after reaching a maximum value. It means that the response sensitivity of the centroid wavelength shift of the emission spectrum to the phase was not the same in different phase change regions. For the rapidly changing region in the middle, the response was sharp and relatively linear, which was very suitable for high-sensitivity phase detection. Therefore, the phase difference caused by the concentration of the oxidized (V(V)) and reduced (V(IV)) species in this experiment can be detected with this linear range.

 figure: Fig. 4.

Fig. 4. The centroid wavelength shifts with respect to the phase. The solid line is the theoretical expectation. The dots are the experimental data

Download Full Size | PDF

3.2 Performance comparison of the TIR method and the WM method

In order to compare the resolution of the TIR and the WM, the TIR and the WM systems were built by the same experiment components (see Supplement 1, Section 1 for details) for reducing the impact of equipment as much as possible and worked within their linear range. For convenience, both systems were tested by LSV at scan rate of 1 mVs-1 and the potential window of 0.5 V-1.2 V. The results were shown in the Fig. 5. Figure 5(a) showed the light intensity variation measured by the TIR system, and Fig. 5(b) showed the centroid wavelength shift measured by the WM system. In the initial stage, the vanadium ions in the vanadium electrolyte were in the form of V(IV) and almost unchanged, so the curve was nearly flat. As the potential increased, V(IV) were oxidized to V(V), so the curve dropped sharply caused by the change of refractive index in the different concentration of the oxidized (V(V)) and reduced (V(IV)) species.

 figure: Fig. 5.

Fig. 5. Performance comparison of total internal reflection (TIR) method and weak measurement (WM) method. (a) Light intensity variation measured by the TIR. (b) Centroid wavelength shift measured by the WM.

Download Full Size | PDF

Generally, the resolution of a system can be evaluated by the ratio of the noise level $\sigma$ to the response of the reaction $\Delta $. In this experiment, the standard deviation of the first 200 s without chemical reaction was used to calculate noise level and the relative resolution $\delta n = {{({{{\sigma _{TIR}}} / {{\Delta _{TIR}}}})} / {({{{\sigma _{WM}}} / {{\Delta _{WM}}}})}}\textrm{ = }4.21$ was used to judge the method performance. It can be seen from the result that the resolution of the WM was four times better than that of the TIR.

3.3 Calibration of weak measurement (WM) system

As we know, the gold film used in the SPR method has low durability in acid electrolyte and produces Faraday current interfering with the electrochemical process, so our group introduced the TIR method without sensing film to detect the electrochemical process [14]. But the resolution of the TIR method is worse than that of the SPR, so it is necessary to have a better performance method without sensing film. From above experiments, we know that the resolution of the WM without sensing film was better than that of the TIR, so it was very suitable for electrochemical measurement. Now we verified that the WM method can be applied to electrochemical measurement, and the calibration experiment was conducted to obtain the current density with a physical unit (details can be seen in Section 2.5).

In order to verify the feasibility of the WM method, LSV test on the positive electrolyte (0.1 M V(IV) and 2 M H2SO4) at different scan rates (0.5 mVs-1, 1 mVs-1, 1.5 mVs-1, 2 mVs-1) and the potential window of 0.5 V-1.2 V was conducted, the results were shown in Fig. 6. The LSV curves of the graphite plate with tape recorded by the EW were obtained, as shown in Fig. 6(a) and the LSV curves obtained by deconvolution calculation of the centroid wavelength shift variations recorded by the WM were shown in Fig. 6(b).

 figure: Fig. 6.

Fig. 6. (a) Linear sweep voltammetry (LSV) curves of the graphite plate with tape in the positive electrolyte of 0.1 M V(IV) and 2 M H2SO4 at different scan rates (0.5 mVs-1, 1 mVs-1, 1.5 mVs-1, 2 mVs-1) recorded by the EW. (b) LSV curves by deconvolution calculation of the centroid wavelength shift variations recorded by the WM. (c) The relationship between the current density obtained by the EW and the relative current density obtained by the WM with the square root of the scan rate. (d) The relationship between the current density obtained by the EW and the relative current density obtained by the WM.

Download Full Size | PDF

According to the comparison of Fig. 6(a) and Fig. 6(b), the results obtained by the WM were similar to the results obtained by the EW, the peak oxidation currents were both increase along with the increasing scan rates. The peak oxidation currents (Ipa (EW)) from the LSV curves in Fig. 6(a) were picked out and calculated their densities (ipa (EW)). The relationship between the current density obtained by the EW (ipa (EW)) with the square root of the scan rate (v1/2) and the relationship between the relative current density obtained by the WM (ipa (WM)) with the v1/2 were both plotted in Fig. 6(c). The linear fitting relations are ${i_{pa}}(EW) = 5.911 \times {v^{{1 / 2}}} - 0.2185$ and ${i_{pa}}(WM) = 0.0547 \times {v^{{1 / 2}}} + 0.0029$ respectively. Both ipa (EW) and ipa (WM) were linear with the v1/2, proving the Randles-Sevcik equation [33]. This verified that the WM method can be applied to electrochemical measurement.

The current density of the graphite plate with tape obtained by dividing the EW currents by the cross-sectional area (1×1 cm2) should be equal to the relative current densities obtained by the WM. In order to obtain the quantitative relationship between the current density obtained by the EW and the relative current density obtained by the WM, the peak oxidation values at different scan rates from the LSV curves in Fig. 6(a), b was picked out and plotted. After linear fitting of the peak oxidation values, the quantitative equation between the current density obtained by the EW and the relative current density obtained by the WM is

$$i_{pa}(EW) = 0.1078 \times i_{pa}(WM)-0.0005$$
as shown in Fig. 6(d). Therefore, the relative current density obtained by the WM was converted to the current density with a physical unit (mA·cm-2).

3.4 CV test by the weak measurement system

As mentioned above, the LSV test was conducted for calibration. But in general, CV test is a more common experiment to evaluate the electrochemical performance in vanadium flow batteries. So the CV test was conducted to further demonstrate the capability of our WM system. A weak measurement system with the electrochemical detection module containing graphite felt was rebuilt for CV testing. The only difference between the CV test and the LSV test was the working electrodes. A platinum electrode (D6, Pt017, ϕ 1mm × 37mm, purity: 99.95% Tianjin Aida Hengsheng Technology Development Co., Ltd, Tianjin, China) was inserted through the graphite felt (10mm × 10mm × 5.4mm, Gansu Haoshi Cabon Fiber Co., Ltd, Gansu, China) as the working electrode in CV test (see Supplement 1, Section 3 for details), while the working electrode in LSV test was a graphite plate with tape. The scanning rate was set as 1 mVs-1, and the potential window was set as 0.55V-1.15V. Figure 7(a) showed the curve of centroid wavelength shift caused by refractive index change collected by the spectrograph during one CV cycle.

 figure: Fig. 7.

Fig. 7. (a) Time-varying centroid wavelength shift of a graphite felt in the positive electrolyte of 0.1 M V(IV) and 2 M H2SO4 at a scan rate of 1 mVs-1 during the CV. (b) CV curves obtained by deconvolution calculation based on the data in (a). (c) CV curve recorded by the EW.

Download Full Size | PDF

According to the changes in the electrochemical reaction, the curve of centroid wavelength shift can be divided into four stages. The stage I (t1→t2) was the initial stage of the reaction, when the vanadium ions in the vanadium electrolyte were at V(IV). In the stage II (t2→t4), as the applied potential increased, vanadium ions gradually lost electrons and were oxidized to V(V). The different concentration of the oxidized (V(V)) and reduced (V(IV)) species caused a change in refractive index, which led to a sharp drop in the centroid wavelength shift. In the stage III (t4→t5), the reverse potential was provided by the EW, the vanadium ions from the previous oxidation were still in the form of V(V). At the stage IV(t5→t7), the vanadium ions V(V) were gradually reduced to V(IV), so the centroid wavelength shift rose again.

Figure 7(b) showed the relationship between the potential and the relative current density obtained by the deconvolution of the centroid wavelength shift variations in Fig. 7(a). Figure 7(c) showed the relationship between the potential and the absolute current recorded by the EW. According to the comparison of Fig. 7(b) and Fig. 7(c), the results obtained by the WM were similar to the results obtained by the EW, so the WM method can be used to conduct the CV test.

We noticed that the oxidation peak (ipa) obtained by the WM was basically consistent with that obtained by the EW, but the reduction peak (ipc) appeared later than that obtained by the EW. The WM system detected a point on the interface where the graphite felt closed to the prism. However, a graphite felt was a cube with a thickness of 5.4 mm and a length and a width of 10 mm. When an electrochemical reaction occurred, it took a certain time to diffuse from the reaction center to the detection point. But the EW detected the graphite felt electrode as a whole, so there was no diffusion problem. In addition, the internal distribution of the graphite felt was not strictly uniform, which caused the point we detected to be different from the average value. Therefore, the result of the WM lag behind that of the EW. This is the reason why we used only one exposed surface graphite plate to do the calibration experiment and why we needed to study two-dimensional imaging.

4. Conclusion

In this paper, we reported that electrochemical reaction can be in situ detected by the weak measurement (WM) method. The feasibility of the WM method was verified by conducting the LSV and the CV tests. Compared with the traditional optical methods such as SPR [12], the WM method without coating can be used in the acid electrolyte of vanadium flow battery, and the higher resolution than TIR in our previous work [14] helped to more accurately detect tiny electrochemical reactions. This extended the methods of electrochemical in situ detection and laid the foundation for two-dimensional (2D) imaging. Based on the existing SPR imaging work [34], a swept frequency light source and a high-speed area array CCD camera are needed for 2D imaging in the next work. In the 2D imaging, a centroid wavelength shift curve can be obtained from each pixel on the area array CCD camera, from which the current density of the point can be obtained, so as to realize the detection of the current density distribution. Thus, we believe that the WM method will be helpful to study the mechanism of electrochemical reaction and to evaluate the performance of electrochemical devices.

Funding

National Natural Science Foundation of China (61875102, 61975089); Oversea cooperation foundation, Graduate School at Shenzhen, Tsinghua University (HW2018007); science and technology research program of Shenzhen City (JCYJ20180508152528735, JCYJ20200109110606054); Tsinghua University Spring Breeze Fund (2020Z99CFZ023).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

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.

Supplemental document

See Supplement 1 for supporting content.

References

1. B. Dunn, H. Kamath, and J. M. Tarascon, “Electrical energy storage for the grid: a battery of choices,” Science. 334(6058), 928–935 (2011). [CrossRef]  

2. Turgut M Gür, “Review of electrical energy storage technologies, materials and systems: challenges and prospects for large-scale grid storage,” Energy Environ. Sci. 11(10), 2696–2767 (2018). [CrossRef]  

3. X. F. Li, H. M. Zhang, Z. S. Mai, H. Z. Zhang, and I. Vankelecom, “Ion exchange membranes for vanadium redox flow battery (VRB) applications,” Energy Environ. Sci. 4(4), 1147–1160 (2011). [CrossRef]  

4. H. M. Zhang, “Vanadium batteries will be cost-effective,” Nature 508(7496), 319 (2014). [CrossRef]  

5. L. Nadjo and J. Savéant, “Linear sweep voltammetry: Kinetic control by charge transfer and/or secondary chemical reactions: I. Formal kinetics,” Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 48(1), 113–145 (1973). [CrossRef]  

6. P. T. Kissinger and W. R. Heineman, “Cyclic voltammetry,” J. Chem. Educ. 60(9), 702 (1983). [CrossRef]  

7. H. Schichlein, A. C. Müller, M. Voigts, A. Krügel, and E. Ivers-Tiffée, “Deconvolution of electrochemical impedance spectra for the identification of electrode reaction mechanisms in solid oxide fuel cells,” J. Appl. Electrochem. 32(8), 875–882 (2002). [CrossRef]  

8. Y. A. Gandomi, D. S. Aaron, T. A. Zawodzinski, and M. M. Mench, “In situ potential distribution measurement and validated model for all-vanadium redox flow battery,” J. Electrochem. Soc. 163(1), A5188–A5201 (2016). [CrossRef]  

9. J. T. Clement, D. S. Aaron, and M. M. Mench, “In situ localized current distribution measurements in all-vanadium redox flow batteries,” J. Electrochem. Soc. 163(1), A5220–A5228 (2016). [CrossRef]  

10. M. Becker, N. Bredemeyer, N. Tenhumberg, and T. Turek, “Polarization curve measurements combined with potential probe sensing for determining current density distribution in vanadium redox-flow batteries,” Journal of Power Sources 307, 826–833 (2016). [CrossRef]  

11. W. Y. Hsieh, C.H. Leu, C.H. Wu, and Y. S. Chen, “Measurement of local current density of all-vanadium redox flow batteries,” J. Power Sources. 271, 245–251 (2014). [CrossRef]  

12. X. N. Shan, U. Patel, S. P. Wang, R. Iglesias, and N. J. Tao, “Imaging local electrochemical current via surface plasmon resonance,” Science 327(5971), 1363–1366 (2010). [CrossRef]  

13. Y. C. Chao, X. N. Shan, and N. J. Tao, “Appling plasmonics based electrochemical microscopy to thin-layer electrochemistry,” J. Electroanal. Chem. 781, 161–165 (2016). [CrossRef]  

14. K. J. Ma, Y. N. Zhang, L. Liu, J. Y. Xi, X. P. Qiu, T. Guan, and Y. H. He, “In situ mapping of activity distribution and oxygen evolution reaction in vanadium flow batteries,” Nat. Commun. 10(1), 5286 (2019). [CrossRef]  

15. K. J. Ma, X. R. Kuang, L. Liu, Y. N. Zhang, Y. X. Sun, J. Y. Xi, Y. H. He, and D. Z. Li, “An optimized angular total internal reflection sensor with high resolution in vanadium flow batteries,” IEEE T. Instrum. Meas. 69(6), 3170–3178 (2020). [CrossRef]  

16. Y. Aharonov, D. Z. Albert, and L. Vaidman, “How the result of a measurement of a component of the spin of a spin-1/2 particle can turn out to be 100,” Phys. Rev. Lett. 60(14), 1351–1354 (1988). [CrossRef]  

17. O. Hosten and P. Kwiat, “Observation of the spin Hall effect of light via weak measurements,” Science 319(5864), 787–790 (2008). [CrossRef]  

18. Z. H. Liu, W. W. Pan, X. Y. Xu, M. Yang, J. Zhou, Z. Y. Luo, K. Sun, J. L. Chen, J. S. Xu, Ch. F. Li, and G. C. Guo, “Experimental exchange of grins between quantum Cheshire cats,” Nat. Commun. 11(1), 3006 (2020). [CrossRef]  

19. C. Ferrie and J. Combes, “Weak value amplification is suboptimal for estimation and detection,” Phys. Rev. Lett. 112(4), 040406 (2014). [CrossRef]  

20. A. N. Jordan, J. Martínez-Rincón, and J. C. Howell, “Technical advantages for weak-value amplification: when less is more,” Phys. Rev. X 4(1), 011031 (2014). [CrossRef]  

21. J. Harris, R. W. Boyd, and J. S. Lundeen, “Weak value amplification can outperform conventional measurement in the presence of detector saturation,” Phys. Rev. Lett. 118(7), 070802 (2017). [CrossRef]  

22. M. Hallaji, A. Feizpour, G. Dmochowski, J. Sinclair, and A. M. Steinberg, “Weak-value amplification of the nonlinear effect of a single photon,” Nat. Phys. 13(6), 540–544 (2017). [CrossRef]  

23. Y. L. Zhang, D. M. Li, Y. H. He, Z. Y. Shen, and Q. H. He, “Optical weak measurement system with common path implementation for label -free biomolecule sensing,” Opt. Lett. 41(22), 5409–5412 (2016). [CrossRef]  

24. D. M. Li, G. Tian, Y. H. He, F. Liu, A. P. Yang, Q. H. He, Z. Y. Shen, and M. G. Xin, “A chiral sensor based on weak measurement for the determination of Proline enantiomers in diverse measuring circumstances,” Biosens. Bioelectron. 110, 103–109 (2018). [CrossRef]  

25. Y. Xu, L. X. Shi, S. X. Li, T. Guan, S. Y. Zhong, X. S. Zhou, D. M. Li, C. X. Guo, Y. X. Yang, X. N. Wang, Z. Y. Li, Y. H. He, L. Y. Xie, and Z. H. Gan, “Detection of macromolecular content in a mixed solution of protein macromolecules and small molecules using a weak measurement linear differential system,” Anal. Chem. 91(18), 11576–11581 (2019). [CrossRef]  

26. T. Guan, X. N. Wang, D. M. Li, Y. L. Zhang, Y. H. He, L. X. Shi, Y. Q. Liu, Y. X. Yang, Y. Xu, and R. Cui, “Determination of tumor marker carcinoembryonic antigen with biosensor based on optical quantum weak measurements,” Sensors 18(5), 1550 (2018). [CrossRef]  

27. S. Schlicht, K. Percin, S. Kriescher, A. Hofer, C. Weidlich, M. Wessling, and J. Bachmann, “Atomic layer deposition for efficient oxygen evolution reaction at Pt/Ir catalyst layers,” Beilstein. J. Nanotech. 11(1), 952–959 (2020). [CrossRef]  

28. J. Dundálek, I. Šnajdr, O. Libánský, J. Vrána, J. Pocedič, P. Mazúr, and J. Kosek, “Zinc electrodeposition from flowing alkaline zincate solutions: Role of hydrogen evolution reaction,” J. Power Sources. 372, 221–226 (2017). [CrossRef]  

29. J. Lee, D. W.M. Arrigan, and D. S. Silvester, “Mechanical polishing as an improved surface treatment for platinum screen-printed electrodes,” Sensing and Bio-Sensing Research 9, 38–44 (2016). [CrossRef]  

30. R. M. Darling, H. S. Shiau, A. Z. Weber, and M. L. Perry, “The Relationship between Shunt Currents and Edge Corrosion in Flow Batteries,” J. Electrochem. Soc. 164(11), E3081–E3091 (2017). [CrossRef]  

31. Y. N. Zhang, K. J. Ma, X. R. Kuang, L. Liu, Y. X. Sun, and J.Y. Xi, “Real time study of the disequilibrium transfer in vanadium redox flow battery at different state of charge via refractive index detection,” J. Phys. Chem. C 122(50), 28550–28555 (2018). [CrossRef]  

32. Y. Xu, L. X. Shi, T. Guan, C. X. Guo, D. M. Li, Y. X. Yang, X. N. Wang, L. Y. Xie, Y. H. He, and W. Y. Xie, “Optimization of a quantum weak measurement system with its working areas,” Opt. Express 26(16), 21119–21131 (2018). [CrossRef]  

33. A. J. Bard and L. R. Faulkner, Electrochemical methods: fundamentals and applications, 2nd edn. Wiley, New York (2001).

34. Y. J. Zeng, R. Hua, L. Wang, D. Y. Gu, J. N. He, S. Y. Wu, H. P. Ho, X. J. Li, J. L. Qu, B. Z. Gao, and Y. H. Shao, “Recent advances in surface plasmon resonance imaging: detection speed, sensitivity, and portability,” Nanophotonics-Berlin 6(5), 1017–1030 (2017). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental document

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 (7)

Fig. 1.
Fig. 1. Electrochemical weak measurement (WM) system. (a) Schematic of the electrochemical WM system. A: Superluminescent diode (SLD); B: achromatic lens (L1); C: linear polarizer (P1); D: self-made electrochemical detection module (EDM); E: Soleil-Babinet compensator (SBC); F: linear polarizer (P2); G: achromatic lens (L2); H: spectrograph. (b) A photograph of the electrochemical WM system.
Fig. 2.
Fig. 2. The self-made electrochemical detection module (EDM, D in Fig. 1(a)). (a) Expanded schematic view of the EDM, D1: fixator; D2: prism holder; D3: prism; D4: electrochemical reaction cell; D5: graphite plate with insulating tape (working electrode, WE); D6: graphite rod (counter electrode, CE); D7: saturated calomel electrode (reference electrode, RE); D8: water bath pool; D9: fixator. (b) Combined schematic view of the EDM. (c) Photograph o of the EDM. (d) Photograph of graphite plate with insulating tape (working electrode, WE).
Fig. 3.
Fig. 3. (a) The centroid wavelength shift of two-peak spectrum obtained by spectrograph. (b) Time-varying centroid wavelength shift based on the data in (a) and${\lambda _{pa}}$ is the point with the most drastic centroid wavelength shift change. (c) Denoising result by smoothing the centroid wavelength shift variations in (b) and$\Delta {\lambda _{\textrm{pa}}}$corresponds to${\lambda _{pa}}$in (b). (d) Curves by deconvolution calculation of the centroid wavelength shift variations in (c) and the peak current ipa is the result obtained by the deconvolution of$\Delta {\lambda _{\textrm{pa}}}$in (c).
Fig. 4.
Fig. 4. The centroid wavelength shifts with respect to the phase. The solid line is the theoretical expectation. The dots are the experimental data
Fig. 5.
Fig. 5. Performance comparison of total internal reflection (TIR) method and weak measurement (WM) method. (a) Light intensity variation measured by the TIR. (b) Centroid wavelength shift measured by the WM.
Fig. 6.
Fig. 6. (a) Linear sweep voltammetry (LSV) curves of the graphite plate with tape in the positive electrolyte of 0.1 M V(IV) and 2 M H2SO4 at different scan rates (0.5 mVs-1, 1 mVs-1, 1.5 mVs-1, 2 mVs-1) recorded by the EW. (b) LSV curves by deconvolution calculation of the centroid wavelength shift variations recorded by the WM. (c) The relationship between the current density obtained by the EW and the relative current density obtained by the WM with the square root of the scan rate. (d) The relationship between the current density obtained by the EW and the relative current density obtained by the WM.
Fig. 7.
Fig. 7. (a) Time-varying centroid wavelength shift of a graphite felt in the positive electrolyte of 0.1 M V(IV) and 2 M H2SO4 at a scan rate of 1 mVs-1 during the CV. (b) CV curves obtained by deconvolution calculation based on the data in (a). (c) CV curve recorded by the EW.

Equations (12)

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

E ( t ) = E i v t
C O ( 0 , t ) C R ( 0 , t ) = exp [ n F R T ( E i v t E θ ) ]
C O ( z , t ) t = D o 2 C O ( z , t ) 2 z
C O ( 0 , t ) = C O 0 [ n F ( π D O ) 1 / 2 ] 1 0 t i ( t ) ( t t ) 1 / 2 d t
C R ( 0 , t ) = C R 0 + [ n F ( π D R ) 1 / 2 ] 1 0 t i ( t ) ( t t ) 1 / 2 d t
λ ( t ) = K [ α O C O ( z , t ) | z = 0 + α R C R ( z , t ) | z = 0 ]
λ  =  2 π k ( δ λ ) 2 λ 0 Im A W = 4 π k ( δ λ ) 2 γ sin ( a χ 2 + Δ ) λ 0 ( 1 + γ 2 2 γ cos ( a χ 2 + Δ ) )
λ ( t ) = λ ( t 0 ) + K ( α R D R 1 / 2 α O D O 1 / 2 ) ( n F π 1 / 2 ) 1 0 t i ( t ) ( t t ) 1 / 2 d t
Δ λ ( t ) = λ ( t ) λ ( t 0 ) = b n F 0 t i ( t ) [ π ( t t ) ] 1 / 2 d t
n F b Δ λ ( t ) = 0 t i ( t ) [ π ( t t ) ] 1 / 2 d t = c o n v o l u t i o n [ i ( t ) , ( π t ) 1 / 2 ]
i ( t ) = d e c o n v o l u t i o n [ n F b Δ λ ( t ) , ( π t ) 1 / 2 ]
i p a ( E W ) = 0.1078 × i p a ( W M ) 0.0005
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.