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

Sensing the profile of particulate beam attenuation coefficient through a single-photon oceanic Raman lidar

Open Access Open Access

Abstract

A lidar technique has been proposed and demonstrated for remotely sensing particulate beam attenuation coefficient (cp) profiles using the Raman backscattered signal from water. In Raman lidar, the backscatter coefficient at 180° can be considered constant, allowing for the determination of the lidar attenuation coefficient (Klidar) from the Raman backscattered signal. This scheme has these features. 1) The bandwidth of the filter that used to extract the Raman component from the backscattered signal of the lidar was optimized to ensure sufficient lidar signal strength while minimizing the influence of chlorophyll fluorescence on inversion. 2) A receiving telescope with narrow field of view (FOV) and small aperture was utilized to suppress multi-scattering components in the backscattered signal. 3) A relationship between the beam attenuation coefficient (c) and Klidar was established after simulations via a semi-analytic Monto Carlo. 4) The value of cp was obtained by subtracting the attenuation coefficient of pure seawater (cw) from c. According to the theoretical analysis, the maximum relative error of cp is less than 15% for chlorophyll concentrations up to 10 mg/m3. Due to the water Raman backscattered signal being several orders of magnitude lower than the elastic backscattered signal, a single-photon detector is required to significantly improve the detection sensitivity to the single-photon level. To validate this approach, a field experiment was conducted aboard the R/V Tan Kah Kee in the South China Sea from September 4th to September 5th, 2022, and continuous subsurface profiles of cp were obtained. These measurements confirm the robustness and reliability of the oceanic single-photon Raman lidar system and the inversion method.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The particulate beam attenuation coefficient (cp) is a crucial parameter in oceanography. It quantifies the rate at which light beam is absorbed and scattered by particles suspended in seawater, which affects the propagation of light and the amount of energy available for photosynthesis and other biological processes. Since cp is closely related to the concentration, size, shape, and composition of particles in the water column, it provides valuable information on the physical and biogeochemical properties of marine ecosystems.

One of the key applications of cp in ocean science is its use in estimating the amount of particulate organic carbon (POC) in the water column, which is a major component of the oceanic carbon cycle [1]. POC is an important constituent of carbon budgets in the ocean, being produced by primary producers such as phytoplankton and exported to the deep ocean through sinking and vertical mixing processes. The concentration of POC in seawater can be challenging to measure directly, but several indirect methods have been developed based on the relationship between cp and POC. Numerous studies have demonstrated a strong relationship between cp and POC in certain regions of the ocean [24]. It is now well established that cp is a useful proxy for estimating POC concentrations from remote sensing and in situ measurements.

Over the years, various technologies have been developed to measure c of water, from which cp can be obtained by subtracting the contribution of pure seawater (cw). These technologies include the Beer-Lambert law, the stimulated Brillouin scattering, transmissometer method, and spectrophotometric method, among others [57]. While in situ measurements offer high accuracy, they are constrained by their limited spatial coverage and shallow depth range, which can hinder their ability to capture the large-scale variability and dynamics of marine systems. On the other hand, remote sensing provides an efficient way to monitor changes in c of water over time and space, making it a valuable tool for environmental management and research.

However, accurately retrieving c through remote sensing techniques, both passive and active, is still a challenging task [810]. One approach is to use the chlorophyll concentration (Chl) that derived from passive ocean color remote sensing [11]. However, this approach can be limited by the influence of physiological processes, such as changes in intracellular pigmentation, as well as the absence of an accurate indicator of non-phytoplankton particle populations. Another approach for estimating c involves using lidar depolarization measurements and diffuse attenuation coefficients (Kd), which can be obtained from ocean color products or Brillouin lidar measurements [9]. However, simultaneous acquisitions of high-precision lidar depolarization and Kd using satellite sensors remain a challenge.

Fortunately, as an active technique, lidar has the capability to measure time-resolved backscattered and forward-scatter photon contributions arriving at the detector, making it an important technique for remote sensing of oceanic parameters such as particulate backscattering, size spectrum, and vertical distribution of ocean particles [12]. Despite its potential benefits, it remains a challenge using oceanic lidars in accurately retrieving c. The primary challenge lies in accurately inverting the lidar attenuation coefficient (Klidar), with the second challenge being the calculation of c or Kd from Klidar. Tremendous efforts have been devoted to developing advanced algorithms [13] and updating the hardware of lidar [14,15].

One major limitation of an elastic backscatter lidar is that it faces an ill-posed mathematic problem, as it needs to infer two unknowns, namely, attenuation coefficient (Klidar) and backscatter coefficient at 180° (β), from a single measurement. Numerous attempts have been made to resolve this inherent ill-posed problem in the lidar equation. Initially, various algorithms have been proposed without changing the mechanism of elastic backscatter lidar, including the slope method [16], Klett method [17], Fernald method [18] and perturbation method [19], among others. Nevertheless, each method is based on a set of assumptions that may not be perfect, leading to certain levels of inverse error. An important leap in retrieval accuracy has been achieved with the high spectral-resolution lidar (HSRL) technique, which can independently measure backscattering and attenuation by separating the particulate and molecular backscatters in wavelength distribution [20,21]. Recently, by combination the HSRL technique and a developed multiple scattering correction algorithm, Kd can be estimated accurately [15]. However, the complexity of the HSRL system and its high requirements for frequency stability limit its implementation and operation to some extent.

Compared to HSRL, Raman lidar systems are easier to implement and maintain. While a Raman lidar has been proposed and utilized to profile atmospheric attenuation coefficient [22], its application in remote sensing of oceanic optical parameters is still limited [23,24]. In this work, an oceanic Raman lidar is proposed and demonstrated for remotely sensing the cp profiles. In Raman lidar, β can be treated as a constant value [22]. Therefore, analyzing of the Raman lidar signal alone permits the determination of Klidar. On the other hand, to overcome the disadvantage that the intensity of Raman backscattered signals is 2-3 orders of magnitude lower than the elastic backscattered signals, a single-photon detector is required to enhance the detection sensitivity to a single-photon level. The single-photon detector technique has been widely applied in atmospheric and target imaging [2527], but its application in oceanic lidar is limited [2830], particularly in oceanic Raman lidar.

After acquiring Klidar, the next challenge is to invert the values of c from the Klidar data. Based on Monte Carlo (MC) simulation, researches have concluded that Klidar is determined by c when lidar backscattered signal is dominated by quasi-single scattering, while Klidar is given by Kd when backscattered signal is dominated by multi-scattering [31]. Based on these findings, in this work, an all-fiber connected Raman lidar with a receiver telescope with narrow FOV (2.1 mrad) and small aperture (22 mm) was developed to suppress the multi-scatter components in backscattered lidar signal. Finally, with a relationship between c and Klidar establish by MC simulations, c can be estimated from Klidar. The organization of this article is as follows: Firstly, an introduction to methodology is presented, including principles and an analysis of potential influencing factors. Next, the MC method is introduced to establish the relationship between c and Klidar. Finally, a field experiment is introduced, which validate the robustness and feasibility of this approach and the lidar system.

2. Methodology

Let the Raman lidar equation be written as follows:

$${P_w}({\lambda ,z} )= K \cdot O({z + n \cdot h} ){\beta _t}{\exp \left\{ { - \int_0^z {K_{lidar}^t({\lambda_L},\lambda ,\;\xi )\;} d\xi } \right\}} \left/ {{{(z + n \cdot h)}^2}}\right.$$
where Pw represents the water Raman backscattered signal at a depth of z; h represents the height at which the lidar is located above the water surface, which in this case is 15 m; λL denotes the wavelength of the transmitting laser, which is 532 nm; λ represents wavelength of Raman signal emitted by water; K is a constant that includes parameters that are independent of depth such as the output laser power, quantum efficiency of the detector, and transmittance of the optical system for both the transmitted and received signals; O(z+$n \cdot h$) is the overlap function [32]; n is the index of refraction of water; βt represents the total backscatter coefficient at 180° at 650 nm, which includes both the Raman backscatter coefficient at 180° of water (defined as βR) and the backscatter coefficient at 180° of chlorophyll fluorescence (defined as βF) at this wavelength; Ktlidar is the round-trip attenuation coefficient of lidar, including the attenuation coefficients of downwelling Klidar(532, z) and upwelling Klidar(650, z).

βt can be represented by the following equation:

$${\beta _t} = {\beta _R} + {\beta _F}$$
Assuming that βt is dominated by βR, while the influence of βF can be neglected (the error caused by this assumption will be analyzed in detail in section 4.1). Since the density of seawater varies slightly with depth within the subsurface water (≤15 m) [33], the influence of density on βR can be considered negligible in this work. Then, βR can be expressed as follows [34]:
$${\beta _R}(\lambda )= {b_R}({\lambda _L},\lambda ) \cdot {f_R}({\lambda _L},\lambda ) \cdot {\tilde{\beta }_R}(\pi )$$
where bR is the Raman scattering coefficient of the water molecules, which tells how much of the irradiance at the excitation wavelength λL scatters into all emission wavelengths λ; fR is the Raman wavelength redistribution function; ${\tilde{\beta }_R}(\pi )$ is the Raman scattering phase function. It should be noted that since βR is only related to water molecules and can be considered constant [34], the two unknowns in the lidar equation, βR and Klidar, can be reduced to only one unknown, Klidar. This is the reason why a Raman lidar can accurately invert Klidar.

When the Raman backscattered signal passes through a Raman filter, it can be expressed as the convolution of the Raman backscattered signal Pw and the transmission function of the filter, as shown in the following equation:

$${P_m}(z )= {P_w}(z )\otimes g(\lambda ,\sigma )$$
where $g(\lambda ,\sigma )$ is the transmittance function of the filter, which can be represented by a Gaussian function in this work, namely,
$$g(\lambda ,\sigma ) = {1 / {(\sigma \sqrt {2\pi } )\exp [{{{ - {\lambda^2}} / {(2{\sigma^2})}}} ]}}$$
where $\sigma$ is its standard deviation, which can be expressed as ${{FWHM} / {2\sqrt {\ln 4} }}$, and FWHM is the full width at half maximum of the Raman filter.

When the geometric overlap factor O(z) beneath the water surface is equal to 1, Ktlidar can be calculated using a simple slope method, as follows:

$$K_{lidar}^t(\;z) = \frac{d}{{dz}}\bigg\{{{\ln \bigg[{\frac {{1}}{P_m(z)\cdot {{(z + n \cdot h)}^2}} \bigg]}}}\bigg\}$$
Then, through a semi-analytic MC simulation, which will be described in section 5, the relationship between Ktlidar and the total beam attenuation coefficient can be established as follows:
$${c^t}(z) = \exp \{{m \cdot {{[{K_{lidar}^t(z)} ]}^2}\textrm{ + }n \cdot K_{lidar}^t(z) + t} \}$$
where m, n and t are fitted parameters. It should be noted that ct(z) is the sum of the beam attenuation coefficients c at the transmitted wavelength of 532 nm and the received wavelength of 650 nm, namely,
$${c^t}(z) = c(532,z) + c(650,z)$$
where c can be expressed as the sum of the absorption coefficient (a) and scattering coefficient (b). By separating the particle contribution and the pure water contribution in c, ct can also be expressed as:
$${c^t}(z) = c_p^t(z) + c_w^t$$
$c_w^t$ is the sum of the beam attenuation coefficients of pure water at 532 nm and 650 nm, which can be regarded as a constant. Similarly, $c_p^t$ represents the sum of the particulate beam attenuation coefficients at 532 nm and 650 nm. By substituting Eq. (9) into Eq. (7), $c_p^t$ can be obtained as follows:
$$c_p^t(z )= \exp [{m \cdot K_{lidar}^t{{(z)}^2}\textrm{ + }n \cdot K_{lidar}^t(z) + t} ]- c_{_w}^t$$
Subsequently, the particulate beam attenuation coefficient at 532 nm, namely cp (532, z), is obtained through the ratio of particulate beam attenuation coefficients (r650/532), which can be expressed as:
$${r_{650/532}} = {c_p}({650,z} )/{c_p}({532,z} )$$
Several models for the inherent optical properties (IOPs) of water bodies [3541] have revealed that, for both Case 1 and Case 2 waters, the ratio r650/532 only exhibits a small range for chlorophyll concentration (Chl) varying between 0.01 mg/m3 and 10 mg/m3 (further details regarding this will be provided in section 4.1). Thus, the profile cp (532, z) can be expressed as:
$${c_p}({532,\;z} )= \{{\exp [{m \cdot K_{lidar}^t{{(z)}^2}\textrm{ + }n \cdot K_{lidar}^t(z) + t} ]- c_{_w}^t} \}/(1 + {r_{650/532}})$$
To provide a clear overview of the inversion process, Fig. 1 provides a complete flowchart of the inversion process.

 figure: Fig. 1.

Fig. 1. Flowchart of the inversion process.

Download Full Size | PDF

3. Optimization of the bandwidth of the Raman filter

Based on Eq. (4), it is evident that the intensity of the backscattered signal is influenced by the shape and bandwidth of the filter after it passes through the filter. The primary impact of the filter on the backscattered signal is reflected in its effect on βt [25]. This effect can be analyzed using β′t, which is mathematically expressed as follows:

$${\beta ^{\prime}_t} = {\beta _t} \otimes g(\lambda ,\sigma ) = {\beta _R} \otimes g(\lambda ,\sigma ) + {\beta _F} \otimes g(\lambda ,\sigma )$$
To simplify the analysis of the impact of the filter transmission function on βR and βF, we define ${\beta ^{\prime}_R} = {\beta _R} \otimes g(\lambda ,\sigma )$ and ${\beta ^{\prime}_F} = {\beta _F} \otimes g(\lambda ,\sigma )$, where βR is obtained from Eq. (3). It should be noted that the center wavelengths of Raman scattering and fluorescence scattering are different, with respective values of 650 nm and 685 nm. Additionally, in order to simplify the analysis, a model for chlorophyll fluorescence is applied to calculate βF when only considering chlorophyll fluorescence and ignoring fluorescence from other substances in the water. This model is applicable in Case 1 water, where the inherent optical properties can be adequately described by phytoplankton (represented by chlorophyll concentration, or Chl). The effect of fluorescence from other substances on the inversion will be studied in our future work. Then, βF can be expressed as [42]:
$${\beta _F}(\lambda ,\textrm{Chl}) = {a_{ph}}(532,\textrm{Chl}) \cdot {\Phi _C}\frac{{532}}{\lambda }{h_C}(\lambda ) \cdot \frac{1}{{4\pi }}$$
where aph(532, Chl) is the chlorophyll absorption coefficient at the excitation wavelength of 532 nm, which can be obtained from Bricaud's model for a given Chl [43]; ФC is the quantum efficiency for chlorophyll fluorescence, with a value between 0.005 and 0.07 [44], and 0.06 being used in this study; hC denotes the chlorophyll fluorescence wavelength emission function from the model [45]. Further, aph (532, Chl) and hC(λ) can be mathematically expressed as:
$${a_{ph}}(532,\textrm{Chl}) = 0.0113 \cdot \textrm{Ch}{\textrm{l}^{0.871}}$$
$${h_C}(\lambda )\textrm{ = }W\sqrt {\frac{{4\ln 2}}{\pi }} \frac{1}{{25}}\exp [ - 4\ln 2{(\frac{{\lambda - 685}}{{25}})^2}] + (1 - W)\sqrt {\frac{{4\ln 2}}{\pi }} \frac{1}{{50}}\exp [ - 4\ln 2{(\frac{{\lambda - 730}}{{50}})^2}]$$
where W represents the weight of the Gaussians, which is set as 0.75 in this study.

The effect of filter bandwidth on β′R is first analyzed, as shown in Fig. 2(a). it can be seen that the larger the bandwidth, the stronger the Raman backscattered signal, as indicated by the increasing value of β′R. Therefore, on one hand, using a larger filter bandwidth is advantageous for improving the signal-to-noise ratio (SNR) of the detection. On the other hand, a larger bandwidth also results in more chlorophyll fluorescence being detected in the lidar backscattered signal, as shown in Eq. (13). This increase in fluorescence causes the assumption of a constant βt to no longer hold, introducing errors in the inversion of cp.

 figure: Fig. 2.

Fig. 2. (a) Effect of the bandwidth of Raman filter on backscatter coefficient at 180°, (b) effect of filter bandwidth on β′t for the Chl ranging from 0.01 to 10 mg/m3.

Download Full Size | PDF

Then, the effect of filter bandwidth on β′t is analyzed as follows:

$$Effect = {{[{{{\beta}_t^{\prime}} (650,\sigma ,{Chl}) - {{\beta_t^{\prime}}}(650,\sigma ,0.01)} ]} / {{{\beta_t ^{\prime}}}(650 ,\sigma ,0.01)}} \times 100\%$$
where β′t (650, σ, 0.01) is a reference value when Chl is set to 0.01 mg/m3. As shown in Fig. 2(b), an increase in the filter bandwidth and an increase in Chl both lead to a larger relative error in β′t.

To minimize the influence of chlorophyll fluorescence while maintaining the strength of the Raman backscattered signal, the bandwidth of the Raman filter was optimized. Finally, a bandwidth of 6 nm was selected to guarantee that the effect of chlorophyll fluorescence on β′t remained below 5% for Chl ranging from 0.01 to 10 mg/m3.

4. Analysis of the effects of r650/532 and chlorophyll fluorescence on cp(532) inversion

4.1 r650/532 effect

As described in the methodology section, determining the r650/532 ratio is crucial for calculating cp(532) from ct data. In this study, we utilized the IOPs model for Case 1 waters and measured data from Case 2 waters to investigate the variation of r650/532 in different water bodies. Previous researches have extensively investigated the IOPs of Case 1 waters, which are primarily influenced by phytoplankton, as initially proposed by Morel and Prieur in 1977 [46] and is still widely adopted today. In this study, we employed two empirical relationships to calculate the particulate absorption coefficient (ap) [35,36,41], and four empirical relationships to calculate the particulate scattering coefficient (bp) [3740], as listed in Table 1. Figure 3(a) shows the spectrogram of cp from 400 nm to 700 nm, with Chl ranging from 0.01 to nearly 60.0 mg/m3. In most Case 1 waters, Chl ranges from 0.01 to 1.00 mg/m3 [4749]. However, to illustrate the distribution of r650/532 in Case 2 waters with high Chl, we extended Chl up to nearly 30.0 mg/m3.

 figure: Fig. 3.

Fig. 3. (a) Spectrogram of the cp under varying Chl, (b) distribution of r650/532 in case 1 water (lines) and case 2 water (circles). cp1 is the sum of ap1 and bp1; cp2 is the sum of ap1 and bp2; cp3 is the sum of ap1 and bp3; cp4 is the sum of ap1 and bp4; cp5 is the sum of ap2 and bp1; cp6 is the sum of ap2 and bp2; cp7 is the sum of ap2 and bp3; cp8 is the sum of ap2 and bp4 [Refs. 5055].

Download Full Size | PDF

Tables Icon

Table 1. Absorption and scattering coefficients of particles in case 1 water.

By combining two sets of empirical relationships for parameters ap [35,36,41] and bp [3740], we generated eight different models for parameter cp. These models were used to calculate the r650/532 ratio at different Chl levels, as shown by the dashed lines in Fig. 3(b). As shown in Fig. 3(b), the r650/532 ratio tends to stabilize as Chl increases, especially when Chl exceeds 3 mg/m3. However, for oceanic waters with relatively low Chl concentrations, r650/532 exhibits a significant relative variation with changes in Chl.

In contrast to Case 1 waters, cp in Case 2 waters is not solely determined by Chl. We analyzed the variations in r650/532 at ten different locations in Case 2 waters, and their corresponding values versus Chl are illustrated by circles in Fig. 3(b). Detailed information on the specific locations and relevant references is also provided in Fig. 3(b). The minimum and maximum values of r650/532 for Case 2 waters are 0.47 and 0.92, respectively, with corresponding Chl values of 19.8 mg/m3 and 1.2 mg/m3. The range of variation in r650/532 determined through this analysis is employed to investigate the impact of changes in the r650/532 value on the inversion of cp.

Based on the previous analysis, it is evident that the value of r650/532 actually varies with Chl. However, in the inversion process, we assume it to be a fixed value due to the lack of prior knowledge on Chl. Assuming that the ratio of particulate beam attenuation coefficients used in the inversion process is r′650/532, and the corresponding particulate beam attenuation coefficient obtained from the inversion is cp(r′650/532), the relative deviation between cp(r′650/532) and the true ground of cp(r650/532) can be expressed as follows:

$$Erro{r_r} = {{[{{c_p}({532,\;{{r^{\prime}}_{650/532}}} )- {c_p}({532,{r_{650/532}}} )} ]} / {{c_p}({532,{r_{650/532}}} )}}$$
Combined with Eq. (12) in the methodology section, Errorr can be simplified as:
$$Erro{r_r} = \{{[{1/(1 + {{r^{\prime}}_{650/532}})} ]- [{1/(1 + {r_{650/532}})} ]} \}/[{1/(1 + {r_{650/532}})} ]\; \times \;100\;\%$$
The distribution of Errorr, when the value of r′650/532 used in the inversion process, varies from 0.45 to 0.9, and the true ground value of r650/532 also varies within this range, is shown in Fig. 4(a). The results indicate that the Errorr is small when r′650/532 is close to r650/532, and the error increases as the deviation between r′650/532 and r650/532 becomes larger. However, the Errorr remains within the range of -23.7% to 31.0%. in this work, r′650/532 is set to 0.65, and Errorr ranges from -12.1% to 15.2%, as shown in Fig. 4(b).

 figure: Fig. 4.

Fig. 4. The estimated error caused by the change of the ratio r650/532 to the inversion of cp.

Download Full Size | PDF

4.2 βF effect

In this section, we will analyze the error that arises from neglecting the vertical variation of the backscattered coefficient of chlorophyll fluorescence βF during the inversion of cp. Firstly, four typical Chl profiles were obtained, including open-ocean, mid-latitude case 1 water, lakes, and water surrounding Europe [5659]. These water bodies were chosen to represent a diverse range of non-uniform aquatic environments and their associated chlorophyll distributions. Then, cp(532, Chl) was calculated based on models using Chl profiles [3541]. According to Eq. (9) and Eq. (11), ct can be expressed as a function of cp as follows:

$${c^t} = ({1 + {r_{650/532}}} ){c_p}({532,\textrm{Chl}} )+ c_w^t({\lambda _{L,R}})$$
Then, Ktlidar was obtained by using the relationship between ct and Ktlidar established based on the inverse function of Eq. (10). βt was obtained by using Chl profile based on Eq. (2). Finally, the normalized signal of the Raman lidar was simulated based on Eq. (1). Then, using the approach provided in the methodology section, the particulate beam attenuation coefficient can be obtained, defines as c′p(532, Chl). The deviation between the inversion and true values of the particulate beam attenuation coefficient can be estimated as follows:
$$Erro{r_f} = {{[{{{c^{\prime}}_p}({532,\;\textrm{Chl}} )- {c_p}{{({532,\;\textrm{Chl}} )}_{gt}}} ]} / {{c_p}{{({532,\;\textrm{Chl}} )}_{gt}}}} \times 100\%$$
where cp(532, Chl)gt is the true value calculated by the models provided in Table 1, with the provided Chl profiles.

In the above evaluation process, the same value of r650/532 was used to generate the lidar backscattered signal from the known vertical distribution of Chl and to invert cp(532, Chl) from the lidar backscattered signal. Therefore, this relative deviation only considers the effect of Chl variation on βt, but does not consider the effect of r650/532 on the inversion.

To take the effect of r650/532 into account, r650/532 is no longer considered a constant but a variable related to the vertical profiles of Chl when using the vertical profile of Chl to generate the lidar backscattered signal. Based on this backscattered signal and using the same inversion method as described above, the particulate beam attenuation coefficient can be obtained, defined as c′p(532, r650/532, Chl). In this case, the relative deviation between the inverted value c′p(532, r650/532, Chl) and the true value cp(532, Chl)gt can be considered as a total error, defined as Errort, which takes into account both the influence of the vertical distribution of Chl on βt and r650/532. It can be expressed as follows:

$$Erro{r_t} = {{[{{{c^{\prime}}_p}({532,\;{r_{650/532}},\;Chl} )- {c_p}{{({532,\;Chl} )}_{gt}}} ]} / {{c_p}{{({532,\;Chl} )}_{gt}}}} \times 100\%$$
Figure 5 presents the distribution of three types of errors, namely Errorr, Errorf, and Errort under various Chl vertical distributions. As shown in Fig. 5(a) and Fig. 5(b), when there is a monotonic change in Chl, whether increasing or decreasing, the errors increase with an increase in Chl. In such cases, Errort is primarily influenced by Errorr, and the total error Errort remains below 10%. On the other hand, when there is a layered distribution of Chl, Errort is primarily attributed to Errorr when Chl is low or relatively stable. On the other hand, when there is a layered distribution of Chl, as shown in Fig. 5(c) and Fig. 5(d), Errort is primarily attributed to Errorr when Chl is low or relatively stable. However, if Chl is high, both Errorr and Errorf contribute approximately equally to Errort, and the magnitude of Errort increases with an increase in Chl variation rate, which is defined as ΔChl. Regarding Errorf, as shown in Fig. 5, especially in Fig. 5(c-3) and Fig. 5(d-3), even when the Chl varies significantly with depth, i.e., when βF varies vertically, Errorf, as calculated using Eq. (21), can be controlled to be less than 8% due to the utilization of a narrow Raman filter with a bandwidth of 6 nm. Overall, in these four water types, the maximum value of Errort does not exceed 15%.

 figure: Fig. 5.

Fig. 5. Theoretical analysis of errors. The sub-figures (a)-(d) show different chlorophyll vertical distribution, including (a) linearly decreasing [56], (b) linearly increasing [59], (c) bimodal with two Gaussian distribution [58], and (d) unimodal with a single Gaussian distribution [57]. Each sub-figure comprises three panels, with the top showing the corresponding Chl vertical distribution and relative Chl difference between adjacent depth intervals (ΔChl). The second panel provides the inversed c′p and its ground true cp, while the third panel displays the Errorf and Errorr distribution. The bottom panel shows the Errort distribution.

Download Full Size | PDF

5. Monte Carlo simulation

MC simulation is a widely used tool in lidar simulation and has been extensively applied to simulate the backscattered signal of oceanic lidars [24]. In this study, a brief introduction is provided to the MC-based simulation of Raman lidar backscattered signals, without going into specific details. For more information on the simulation process, recent articles can be referred to [24,60].

The MC is used to simulate the random trajectories of photon propagation in a medium. The step and direction of photon trajectories depend on the scattering and absorption properties of the medium. Meanwhile, the MC method treats the photon as a typical particle and ignores its wave properties. The propagation of laser in water is treated as the combination of many photon trajectories. Laser energy attenuation is determined by three factors, namely the absorption of the medium, the scattering probability, and the probability distribution of the steps. To enhance the utilization efficiency of individual photons, a semi-analytic MC model is applied. This model enables every photon in the telescope FOV to return its expected value of energy and record its position [61,62]. Due to the weak intensity of Raman signals, only the first-order Raman signals that arise from the initial laser wavelength are considered, while the second-order Raman that result from the first-order Raman signals are disregarded. The simulated lidar backscattered signal is dependent not only on the IOPs of the water, but also on the hardware and geometric parameters of the lidar. Table 2 presents the instrument parameters of the Raman lidar used as input in the MC simulation. It should be noted that the instrument parameters used in MC are the parameters of the actual shipborne Raman lidar.

Tables Icon

Table 2. Input parameters of lidar system

The absorption and scattering coefficients are modeled as follows:

$$a(\lambda ) = {a_w}(\lambda ) + 0.06A(\lambda ) \cdot \textrm{Ch}{\textrm{l}^{0.65}} + {a_y}(\lambda )$$
$$b(\lambda ) = {b_w}(\lambda ) + {b_p}(\lambda )$$
where aw is the absorption coefficient of pure seawater [65], A is the normalized spectral absorption values of phytoplankton pigments, ay is the absorption coefficient of yellow substance, bw is the scattering coefficient of pure water [66]. Table 3 provides the model used in MC simulation.

Tables Icon

Table 3. The models used in MC simulation

In the simulation, two commonly used scattering phase functions were utilized, namely Petzold phase function and a phase function proposed by Morel et al. [63,64]. The sampling length is 20 m with a sampling interval is 0.1 m, resulting in a total of 200 sampling points for each simulation. To reduce the effects of multiple scattering in the lidar backscattered signal, a small aperture telescope with a limited FOV was employed. As illustrated in Fig. 6(a), the simulated signal exhibits an exponential decay. Based on statistical analysis of the lidar signal, the percentage of multiple scattering, represented by the second scattering and higher-order scattering, is low when Chl is low, and the lidar signal is dominated by single scattering. However, as Chl increases, the percentage of multiple scattering increases. By using the inversion method introduced by the methodology section, Ktlidar values for different Chls can be obtained. The relationship between ct and Ktlidar is shown in Fig. 6(b). After exponential fitting, the relationship between ct and Ktlidar is determined as:

$${c^t} = \exp [{ - 0.334 \cdot {{(K_{lidar}^t)}^2} + 1.916 \cdot K_{lidar}^t - 1.540} ]$$
From Fig. 6(b), it can be observed that when Chl is low (i.e., ct is small and the lidar backscattered signal is dominated by single scattering), Ktlidar is close to ct. However, as Chl increases and the proportion of multiple scattering in the lidar backscattered signal increases, Ktlidar becomes smaller than ct and gradually deviates from the 1:1 line between Ktlidar and ct (as shown by the dashed line in Fig. 6(b)). However, there is still a strong correlation between Ktlidar and ct, as evidenced by a high coefficient of determination (R2) of 0.99.

 figure: Fig. 6.

Fig. 6. (a) Simulation lidar backscattered signals (lines) and the corresponding proportion of multiple scattering (PMS) in the signals (scatters) for Chl ranging from 0.01 to 10 mg/m3, using the Morel phase function, (b) relationships between ct and cp(532) with Ktlidar, where scatter represents the results of MC simulations, and the solid line represents the results fitted by the formula.

Download Full Size | PDF

6. Field experiment

6.1 Lidar system

Figure 7 illustrates a schematic diagram of the single-photon Raman lidar system setup, which consists four subsystems: a 532 nm pulsed laser, a transceiver, an optical receiver, and a data acquisition system. The system uses a compact fiber-based laser that follows a master oscillator power amplifier (MOPA) architecture, which utilizes a single-mode pulsed seed laser operating at 1064 nm. This seed laser is amplified through a single-mode ytterbium-doped fiber amplifier (SM-YDFA) and a high-power ytterbium-doped fiber amplifier (HP-YDFA) before passing through a lithium borate crystal (LBO) for second harmonic generation, resulting in up to an average power output of 1.0 W with a beam divergence of 0.5 mrad. The output pulse width is 3 ns, and the repetition frequency is 340 KHz.

 figure: Fig. 7.

Fig. 7. Optical layout of the single-photon Raman lidar. SM-YDFA: Single-Mode Ytterbium-Doped Fiber Amplifier; HP-YDFA: High-Power Ytterbium-Doped Fiber Amplifier; L: lens; LBO: lithium borate; DM: dichroic mirror; MMF: Multimode fiber; SPAD: single-photon avalanche diode; TDC: time-to-digital converter; FG: function generator; PC: personal computer.

Download Full Size | PDF

To achieve a miniaturized and robust structure, a fiber-connected configuration is specifically designed for the Raman lidar system. In this configuration, the backscattered signal from water is coupled into a 105-µm multimode fiber (MMF) through a collimator with a 50.8 mm focal length, resulting in a narrow FOV of approximately 2.1 mrad. This narrow FOV not only provides significant suppression of backscattered noise but also suppresses multi-scattering components in the backscattered signal. The distance between the transmitted laser and the received collimator is approximately 15 mm. The backscattered Raman photons are extracted using a 45° dichroic mirror (DM) that reflects the elastic signal at 532 nm and transmits the signal from 550-670 nm, along with a bandpass Raman filter centered at 650 nm with a bandwidth (full width of half maximum) of 6 nm. By combining the DM with the Raman filter, the isolation degree of the 532 nm elastic backscattered signal can be achieved up to 55 dB, while the transmittance of the 650 nm Raman signal is > 80%. For photon detection, a compact silicon single-photon avalanche diode (SPAD) with an efficiency of approximately 52% and 100 dark counts per second (cps) at 650 nm is employed. The electronic module utilizes a self-built function generator (FG) using a Field Programmable Gate Array (FPGA) to generate precise control signals for the laser, the SPAD, and a self-designed time-to-digital converter (TDC). The TDC, with a resolution of 500 ps, is utilized for recording the timing information of pulse emission and photon detection. A summary of the system parameters is presented in Table 4.

Tables Icon

Table 4. Key parameters of the Raman lidar system

6.2 Field experiment

To validate the effectiveness of the single-photon Raman lidar for retrieving cp, a field experiment was conducted aboard the R/V Tan Kah Kee in the South China Sea from 8:38:04 PM on September 4th to 1:10:54 AM on September 5th, 2022. The lidar was installed on the deck of the research vessel, positioned approximately 15 m above the water surface, and the laser beam penetrated the water at a near-zenith angle of 0° after being reflected by a mirror located in front of the lidar. In this setup, the overlap function beneath the surface can be assumed to be 1.

Initially, the Raman lidar collected data with a time resolution of 1 second and a depth resolution of 0.5 ns, as shown in Fig. 8(a). To reduce inversion errors, the time resolution was adjusted to 15 seconds and the depth resolution was set to 1 m. The lidar signals was then multiplied by the square of the corresponding sampling depth, followed by a logarithmic operation. Ktlidar profiles were calculated using the slope method, as provided in section 2. Subsequently, ct profiles were derived based on the established relationship between ct and Ktlidar, as shown in Fig. 8(b). Finally, cp(532) profiles were obtained using the ratio r650/532 (which was selected as 0.65), as illustrated in Fig. 8(c). While the filter combination can isolate elastic scattering by around 55 dB, it should be noted that the specular reflection on the water surface can still interfere with the Raman signal in the first 3 m. Therefore, the inversion results in Fig. 8(b) and 8(c) start from 3 m. In our future work, we will further improve the isolation of elastic scattering. The raw signal presented in Fig. 8(a) indicates the presence of a scattering layer beneath the sea surface, whose characteristics changed slowly during the R/V Tan Kah Kee cruise. By inverting cp(532), as shown in Fig. 8(c), variations can be clearly observed.

 figure: Fig. 8.

Fig. 8. Inversion results of the field experiment: (a) original Pm(z); (b) profiles of ct; (c) profiles of cp (532).

Download Full Size | PDF

7. Conclusion

In this work, we proposed a Raman lidar inversion method for measuring the profile of particulate beam attenuation coefficient, which remains a challenge for current remote sensing technologies. In Raman lidar, β can be treated as a constant value. Therefore, analysis of the Raman lidar signal alone permits the determination of Ktlidar by using a simple slope method. Furthermore, using the relationship between ct and Ktlidar, which was established by a semi-analytic MC mode, ct can be obtained. Finally, cp(532) can be obtained from ct after assuming a ratio (r650/532) of the particulate beam attenuation coefficient at two wavelengths.

In this work, systematic theoretical calculations were conducted, focusing on two main aspects. Firstly, the bandwidth of the Raman filter used to extract the Raman backscattered signal from water was optimized to 6 nm, which reduced the impact of chlorophyll fluorescence on cp(532) inversion while maintaining the intensity of the Raman backscattered signal. Secondly, an analysis was conducted to investigate the effects of r650/532 and chlorophyll fluorescence on cp(532) inversion. The calculations showed that Errort is less than 15% when Chl falls within the range of 0.01-10 mg/m3.

In terms of hardware design, a low-noise and high-efficiency single photon detector was adopted to enhance the performance of Raman lidar, enabling the detection of Raman backscattered profiles from water using a low-power laser. Additionally, a telescope with a small diameter and narrow FOV was designed to reduce the multiple scattering component in the lidar backscattered signal, resulting in Klidar being closer to c. Ultimately, conducting field experiments served to validate the proposed method and demonstrate the feasibility and effectiveness of the Raman lidar in practical applications.

In future work, we plan to conduct a large number of comparative experiments to further validate our proposed method, comparing Raman lidar measurements with measurements obtained using in-situ methods. Additionally, more scattering phase functions will be used in the MC simulations to validate the established relationship between ct and Ktlidar. Moreover, to reduce the influence of chlorophyll fluorescence on the Raman backscattered signal, a laser with shorter wavelength, such as blue laser, will be considered as a transmitter. Once the effect of chlorophyll fluorescence on the Raman backscattered signal is reduced, the bandwidth of the Raman filter can be further increased, thus improving the SNR of the Raman backscattered signal. Furthermore, to enable the single-photon oceanic lidar to operate during daylight, this lidar will be integrated into an AUV platform in our future work. In summary, we believe that this work has significant potential, and the development of this technology will enable scientists to better study ocean carbon stocks and cycles, promoting a greater understanding of the role of the ocean in the global carbon cycle and the marine ecosystem as a whole.

Funding

National Key Research and Development Program of China (2022YFB3901704); Joint Funds of the National Natural Science Foundation of China (U2106210); Natural Science Foundation of Fujian Province (No. 2020J01026); MEL-RLAB Joint Fund for Marine Science & Technology Innovation.

Acknowledgment

We acknowledge the captain and crew of the R/V Tan Kah Kee for the help during the cruises.

Disclosures

The authors declare no conflicts of interest.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

1. M. J. Behrenfeld and E. Boss, “Beam attenuation and chlorophyll concentration as alternative optical indices of phytoplankton biomass,” J. Mar. Res. 64(3), 431–451 (2006). [CrossRef]  

2. W. Gardner, A. Mishonov, and M. Richardson, “Global POC concentrations from in-situ and satellite data,” Deep Sea Res., Part II 53(5-7), 718–740 (2006). [CrossRef]  

3. D. Stramski, R. A. Reynolds, M. Babin, S. Kaczmarek, M. R. Lewis, R. Röttgers, A. Sciandra, M. Stramska, M. Twardowski, and B. Franz, “Relationships between the surface concentration of particulate organic carbon and optical properties in the eastern South Pacific and eastern Atlantic Oceans,” Biogeosciences 5(1), 171–201 (2008). [CrossRef]  

4. M. Galí, M. Falls, H. Claustre, O. Aumont, and R. Bernardello, “Bridging the gaps between particulate backscattering measurements and modeled particulate organic carbon in the ocean,” Biogeosciences 19(4), 1245–1275 (2022). [CrossRef]  

5. M. Ouyang, J. Shi, L. Zhao, X. Chen, H. Jing, and D. Liu, “Real time measurement of the attenuation coefficient of water in open ocean based on stimulated Brillouin scattering,” Appl. Phys. B 91(2), 381–385 (2008). [CrossRef]  

6. M. S. Twardowski, J. M. Sullivan, P. L. Donaghay, and J. R. V. Zaneveld, “Microscale quantification of the absorption by dissolved and particulate material in coastal waters with an ac-9,” J. Atmos. Oceanic Technol. 16(6), 691–707 (1999). [CrossRef]  

7. N. Kozlova, A. Kozlova, and Z. A. Goreeva, “Spectrophotometric methods and their capabilities to study material optical parameters,” in 2017 2nd International Ural Conference on Measurements (UralCon), (IEEE, 2017), 281–288.

8. E. Boss, W. H. Slade, M. Behrenfeld, and G. Dall’Olmo, “Acceptance angle effects on the beam attenuation in the ocean,” Opt. Express 17(3), 1535–1550 (2009). [CrossRef]  

9. Y. Hu, M. Behrenfeld, C. Hostetler, J. Pelon, C. Trepte, J. Hair, W. Slade, I. Cetinic, M. Vaughan, and X. Lu, “Ocean lidar measurements of beam attenuation and a roadmap to accurate phytoplankton biomass estimates,” in EPJ Web of Conferences, (EDP Sciences, 2016), 22003.

10. G. Yang, Z. Tian, Z. Bi, Z. Cui, F. Sun, and Q. Liu, “Measurement of the Attenuation Coefficient in Fresh Water Using the Adjacent Frame Difference Method,” in Photonics, (MDPI, 2022), 713.

11. K. J. Voss, “A spectral model of the beam attenuation coefficient in the ocean and coastal areas,” Limnol. Oceanogr. 37(3), 501–509 (1992). [CrossRef]  

12. J. H. Churnside and J. A. Shaw, “Lidar remote sensing of the aquatic environment,” Appl. Opt. 59(10), C92–C99 (2020). [CrossRef]  

13. Z. Zhang, P. Chen, C. Jamet, D. Dionisi, Y. Hu, X. Lu, and D. Pan, “Retrieving bbp and POC from CALIOP: A deep neural network approach,” Remote Sens. Environ. 287, 113482 (2023). [CrossRef]  

14. J. H. Churnside, V. V. Tatarskii, and J. J. Wilson, “Oceanographic lidar attenuation coefficients and signal fluctuations measured from a ship in the Southern California Bight,” Appl. Opt. 37(15), 3105–3112 (1998). [CrossRef]  

15. Y. Zhou, Y. Chen, H. Zhao, C. Jamet, D. Dionisi, M. Chami, P. Di Girolamo, J. H. Churnside, A. Malinka, and H. Zhao, “Shipborne oceanic high-spectral-resolution lidar for accurate estimation of seawater depth-resolved optical properties,” Light: Sci. Appl. 11(1), 261 (2022). [CrossRef]  

16. J. H. Churnside, J. W. Hair, C. A. Hostetler, and A. J. Scarino, “Ocean backscatter profiling using high-spectral-resolution lidar and a perturbation retrieval,” Remote Sens. 10(12), 2003 (2018). [CrossRef]  

17. J. D. Klett, “Stable analytical inversion solution for processing lidar returns,” Appl. Opt. 20(2), 211–220 (1981). [CrossRef]  

18. F. G. Fernald, “Analysis of atmospheric lidar observations: some comments,” Appl. Opt. 23(5), 652–653 (1984). [CrossRef]  

19. J. H. Churnside and R. D. Marchbanks, “Inversion of oceanographic profiling lidars by a perturbation to a linear regression,” Appl. Opt. 56(18), 5228–5233 (2017). [CrossRef]  

20. J. A. Schulien, M. J. Behrenfeld, J. W. Hair, C. A. Hostetler, and M. S. Twardowski, “Vertically-resolved phytoplankton carbon and net primary production from a high spectral resolution lidar,” Opt. Express 25(12), 13577–13587 (2017). [CrossRef]  

21. J. Hair, C. Hostetler, Y. Hu, M. Behrenfeld, C. Butler, D. Harper, R. Hare, T. Berkoff, A. Cook, and J. Collins, “Combined atmospheric and ocean profiling from an airborne high spectral resolution lidar,” in EPJ Web of Conferences, (EDP Sciences, 2016), 22001.

22. A. Ansmann, M. Riebesell, and C. Weitkamp, “Measurement of atmospheric aerosol extinction profiles with a Raman lidar,” Opt. Lett. 15(13), 746–748 (1990). [CrossRef]  

23. Q. Liu, S. Wu, B. Liu, J. Liu, K. Zhang, G. Dai, J. Tang, and G. Chen, “Shipborne variable-FOV, dual-wavelength, polarized ocean lidar: design and measurements in the Western Pacific,” Opt. Express 30(6), 8927–8948 (2022). [CrossRef]  

24. D. J. Spence, B. R. Neimann, and H. M. Pask, “Monte Carlo modelling for elastic and Raman signals in oceanic LiDAR,” Opt. Express 31(8), 12339–12348 (2023). [CrossRef]  

25. M. Shangguan, H. Xia, C. Wang, J. Qiu, G. Shentu, Q. Zhang, X. Dou, and J.-W. Pan, “All-fiber upconversion high spectral resolution wind lidar using a Fabry-Perot interferometer,” Opt. Express 24(17), 19322–19336 (2016). [CrossRef]  

26. M. Shangguan, H. Xia, C. Wang, J. Qiu, S. Lin, X. Dou, Q. Zhang, and J.-W. Pan, “Dual-frequency Doppler lidar for wind detection with a superconducting nanowire single-photon detector,” Opt. Lett. 42(18), 3541–3544 (2017). [CrossRef]  

27. C. Yu, M. Shangguan, H. Xia, J. Zhang, X. Dou, and J.-W. Pan, “Fully integrated free-running InGaAs/InP single-photon detector for accurate lidar applications,” Opt. Express 25(13), 14611–14620 (2017). [CrossRef]  

28. A. Maccarone, A. Mccarthy, X. Ren, R. E. Warburton, A. M. Wallace, J. Moffat, Y. Petillot, and G. S. J. O. E. Buller, “Underwater depth imaging using time-correlated single-photon counting,” Opt. Express 23(26), 33911–33926 (2015). [CrossRef]  

29. A. Maccarone, F. Rocca, A. Mccarthy, R. Henderson, and G. S. J. O. E. Buller, “Three-dimensional imaging of stationary and moving targets in turbid underwater environments using a single-photon detector array,” Opt. Express 27(20), 28437–28456 (2019). [CrossRef]  

30. X. Shen, W. Kong, P. Chen, T. Chen, G. Huang, and R. Shu, “A Shipborne Photon-Counting Lidar for Depth-Resolved Ocean Observation,” Remote Sens. 14(14), 3351 (2022). [CrossRef]  

31. H. R. Gordon, “Interpretation of airborne oceanic lidar: effects of multiple scattering,” Appl. Opt. 21(16), 2996–3001 (1982). [CrossRef]  

32. U. Wandinger and A. Ansmann, “Experimental determination of the lidar overlap profile with Raman lidar,” Appl. Opt. 41(3), 511–514 (2002). [CrossRef]  

33. V. Gladkikh and R. Tenzer, “A mathematical model of the global ocean saltwater density distribution,” Pure Appl. Geophys. 169(1-2), 249–257 (2012). [CrossRef]  

34. J. S. Bartlett, K. J. Voss, S. Sathyendranath, and A. Vodacek, “Raman scattering by pure water and seawater,” Appl. Opt. 37(15), 3324–3332 (1998). [CrossRef]  

35. A. Bricaud, A. Morel, M. Babin, K. Allali, and H. Claustre, “Variations of light absorption by suspended particles with chlorophyllaconcentration in oceanic (case 1) waters: Analysis and implications for bio-optical models,” J. Geophys. Res.: Oceans 103(C13), 31033–31044 (1998). [CrossRef]  

36. A. Bricaud, A. Morel, and L. Prieur, “Absorption by dissolved organic matter of the sea (yellow substance) in the UV and visible domains1,” Limnol. Oceanogr. 26(1), 43–53 (1981). [CrossRef]  

37. A. Morel and S. Maritorena, “Bio-optical properties of oceanic waters: A reappraisal,” J. Geophys. Res.: Oceans 106(C4), 7163–7180 (2001). [CrossRef]  

38. Y. Huot, A. Morel, M. Twardowski, D. Stramski, and R. Reynolds, “Particle optical backscattering along a chlorophyll gradient in the upper layer of the eastern South Pacific Ocean,” Biogeosciences 5(2), 495–507 (2008). [CrossRef]  

39. H. Loisel and A. Morel, “Light scattering and chlorophyll concentration in case 1 waters: A reexamination,” Limnol. Oceanogr. 43(5), 847–858 (1998). [CrossRef]  

40. A. Morel, “Optical modeling of the upper ocean in relation to its biogenous matter content (case I waters),” J. Geophys. Res. 93(C9), 10749–10768 (1988). [CrossRef]  

41. A. Morel, “Light and marine photosynthesis: a spectral model with geochemical and climatological implications,” Prog. Oceanogr. 26(3), 263–306 (1991). [CrossRef]  

42. L. Zotta, S. Matteoli, M. Diani, and G. Corsini, “AFRODiTE: A fluorescence LiDAR simulator for underwater object detection applications,” IEEE Trans. Geosci. Remote Sensing 53(6), 3022–3041 (2015). [CrossRef]  

43. A. Bricaud, M. Babin, A. Morel, and H. Claustre, “Variability in the chlorophyll-specific absorption coefficients of natural phytoplankton: Analysis and parameterization,” J. Geophys. Res. 100(C7), 13321–13332 (1995). [CrossRef]  

44. S. Maritorena, A. Morel, and B. Gentili, “Determination of the fluorescence quantum yield by oceanic phytoplankton in their natural habitat,” Appl. Opt. 39(36), 6725–6737 (2000). [CrossRef]  

45. C. D. Mobley, Light and water: radiative transfer in natural waters (Academic press, 1994).

46. A. Morel and L. Prieur, “Analysis of variations in ocean color 1,” Limnol. Oceanogr. 22(4), 709–722 (1977). [CrossRef]  

47. B. Bharathiraja, J. Jayamuthunagai, M. Chakravarthy, R. R. Kumar, D. Yogendran, and R. Praveenkumar, “Algae: Promising Future Feedstock for Biofuels,” in Algae and Environmental Sustainability (Springer2015), pp. 1–8.

48. J. O’Reilly, “Status and trends in primary productivity and chlorophyll from 1996 to 2014 in large marine ecosystems and the Western Pacific warm pool, based on data from satellite ocean colour sensors,” Intergovernmental Oceanographic Commission Technical Series 120, 130 (2017).

49. T. Kostadinov, D. Siegel, and S. Maritorena, “Retrieval of the particle size distribution from satellite ocean color observations,” J. Geophys. Res.: Oceans 114(C9), C09015 (2009). [CrossRef]  

50. G. Campbell, S. R. Phinn, and P. Daniel, “The specific inherent optical properties of three sub-tropical and tropical water reservoirs in Queensland, Australia,” Hydrobiologia 658(1), 233–252 (2011). [CrossRef]  

51. A. Reinart, B. Paavel, D. Pierson, and N. Strömbeck, “Inherent and apparent optical properties of Lake Peipsi, Estonia,” Boreal Environ. Res. 9, 429 (2004).

52. S. Kratzer and G. Moore, “Inherent optical properties of the Baltic Sea in comparison to other seas and oceans,” Remote Sens. 10(3), 418 (2018). [CrossRef]  

53. J. Huang, X. Chen, T. Jiang, F. Yang, L. Chen, and L. Yan, “Variability of particle size distribution with respect to inherent optical properties in Poyang Lake, China,” Appl. Opt. 55(22), 5821–5829 (2016). [CrossRef]  

54. D. Ficek, J. Meler, T. Zapadka, B. Woźniak, and J. Dera, “Inherent optical properties and remote sensing reflectance of Pomeranian lakes (Poland),” Oceanologia 54(4), 611–630 (2012). [CrossRef]  

55. Ø Frette, S. Rune Erga, B. Hamre, J. Aure, and J. J. Stamnes, “Seasonal variability in inherent optical properties in a western Norwegian fjord,” Sarsia 89(4), 276–291 (2004). [CrossRef]  

56. R. P. Souto, P. L. Silva Dias, H. F. Campos Velho, S. Stephany, and M. Kampel, “New developments on reconstruction of high resolution chlorophyll-a vertical profiles,” Comp. Appl. Math. 36(3), 1195–1204 (2017). [CrossRef]  

57. H. Loisel, D. Stramski, B. G. Mitchell, F. Fell, V. Fournier-Sicre, B. Lemasle, and M. Babin, “Comparison of the ocean inherent optical properties obtained from measurements and inverse modeling,” Appl. Opt. 40(15), 2384–2397 (2001). [CrossRef]  

58. J. Sánchez-España, C. Falagán, D. Ayala, and K. Wendt-Potthoff, “Adaptation of Coccomyxa sp. to extremely low light conditions causes deep chlorophyll and oxygen maxima in acidic pit lakes,” Microorganisms 8(8), 1218 (2020). [CrossRef]  

59. C. D. Mobley, “Ecosystem Predictions with Approximate vs. Exact Light Fields,” (SEQUOIA SCIENTIFIC INC BELLEVUE WA, 2009).

60. S. Chen, P. Chen, L. Ding, and D. Pan, “A New Semi-Analytical MC Model for Oceanic LIDAR Inelastic Signals,” Remote Sens. 15(3), 684 (2023). [CrossRef]  

61. D. Liu, P. Xu, Y. Zhou, W. Chen, B. Han, X. Zhu, Y. He, Z. Mao, C. Le, and P. Chen, “Lidar remote sensing of seawater optical properties: experiment and Monte Carlo simulation,” IEEE Trans. Geosci. Remote Sensing 57(11), 9489–9498 (2019). [CrossRef]  

62. P. Chen, C. Jamet, Z. Mao, and D. Pan, “OLE: A novel oceanic Lidar emulator,” IEEE Trans. Geosci. Remote Sensing 59(11), 9730–9744 (2021). [CrossRef]  

63. T. J. Petzold, “Volume scattering functions for selected ocean waters,” (Scripps Institution of Oceanography La Jolla Ca Visibility Lab, 1972).

64. A. Morel, D. Antoine, and B. Gentili, “Accounting for Raman emission and varying particle scattering phase function,” Appl. Opt. 41(30), 6289–6306 (2002). [CrossRef]  

65. L. Prieur and S. Sathyendranath, “An optical classification of coastal and oceanic waters based on the specific spectral absorption curves of phytoplankton pigments, dissolved organic matter, and other particulate materials 1,” Limnol. Oceanogr. 26(4), 671–689 (1981). [CrossRef]  

66. A. Morel, “Optical properties of pure water and pure seawater,” Optical aspects of oceanography (1974).

67. Z. Lee and J. Tang, “The Two Faces of “Case-1” Water,” J. Remote Sens. 2022, 9767452 (2022). [CrossRef]  

Data availability

The data that support the findings of this study are available from the corresponding author 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. Flowchart of the inversion process.
Fig. 2.
Fig. 2. (a) Effect of the bandwidth of Raman filter on backscatter coefficient at 180°, (b) effect of filter bandwidth on β′t for the Chl ranging from 0.01 to 10 mg/m3.
Fig. 3.
Fig. 3. (a) Spectrogram of the cp under varying Chl, (b) distribution of r650/532 in case 1 water (lines) and case 2 water (circles). cp1 is the sum of ap1 and bp1; cp2 is the sum of ap1 and bp2; cp3 is the sum of ap1 and bp3; cp4 is the sum of ap1 and bp4; cp5 is the sum of ap2 and bp1; cp6 is the sum of ap2 and bp2; cp7 is the sum of ap2 and bp3; cp8 is the sum of ap2 and bp4 [Refs. 5055].
Fig. 4.
Fig. 4. The estimated error caused by the change of the ratio r650/532 to the inversion of cp.
Fig. 5.
Fig. 5. Theoretical analysis of errors. The sub-figures (a)-(d) show different chlorophyll vertical distribution, including (a) linearly decreasing [56], (b) linearly increasing [59], (c) bimodal with two Gaussian distribution [58], and (d) unimodal with a single Gaussian distribution [57]. Each sub-figure comprises three panels, with the top showing the corresponding Chl vertical distribution and relative Chl difference between adjacent depth intervals (ΔChl). The second panel provides the inversed c′p and its ground true cp, while the third panel displays the Errorf and Errorr distribution. The bottom panel shows the Errort distribution.
Fig. 6.
Fig. 6. (a) Simulation lidar backscattered signals (lines) and the corresponding proportion of multiple scattering (PMS) in the signals (scatters) for Chl ranging from 0.01 to 10 mg/m3, using the Morel phase function, (b) relationships between ct and cp(532) with Ktlidar, where scatter represents the results of MC simulations, and the solid line represents the results fitted by the formula.
Fig. 7.
Fig. 7. Optical layout of the single-photon Raman lidar. SM-YDFA: Single-Mode Ytterbium-Doped Fiber Amplifier; HP-YDFA: High-Power Ytterbium-Doped Fiber Amplifier; L: lens; LBO: lithium borate; DM: dichroic mirror; MMF: Multimode fiber; SPAD: single-photon avalanche diode; TDC: time-to-digital converter; FG: function generator; PC: personal computer.
Fig. 8.
Fig. 8. Inversion results of the field experiment: (a) original Pm(z); (b) profiles of ct; (c) profiles of cp (532).

Tables (4)

Tables Icon

Table 1. Absorption and scattering coefficients of particles in case 1 water.

Tables Icon

Table 2. Input parameters of lidar system

Tables Icon

Table 3. The models used in MC simulation

Tables Icon

Table 4. Key parameters of the Raman lidar system

Equations (25)

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

P w ( λ , z ) = K O ( z + n h ) β t exp { 0 z K l i d a r t ( λ L , λ , ξ ) d ξ } / ( z + n h ) 2
β t = β R + β F
β R ( λ ) = b R ( λ L , λ ) f R ( λ L , λ ) β ~ R ( π )
P m ( z ) = P w ( z ) g ( λ , σ )
g ( λ , σ ) = 1 / ( σ 2 π ) exp [ λ 2 / ( 2 σ 2 ) ]
K l i d a r t ( z ) = d d z { ln [ 1 P m ( z ) ( z + n h ) 2 ] }
c t ( z ) = exp { m [ K l i d a r t ( z ) ] 2  +  n K l i d a r t ( z ) + t }
c t ( z ) = c ( 532 , z ) + c ( 650 , z )
c t ( z ) = c p t ( z ) + c w t
c p t ( z ) = exp [ m K l i d a r t ( z ) 2  +  n K l i d a r t ( z ) + t ] c w t
r 650 / 532 = c p ( 650 , z ) / c p ( 532 , z )
c p ( 532 , z ) = { exp [ m K l i d a r t ( z ) 2  +  n K l i d a r t ( z ) + t ] c w t } / ( 1 + r 650 / 532 )
β t = β t g ( λ , σ ) = β R g ( λ , σ ) + β F g ( λ , σ )
β F ( λ , Chl ) = a p h ( 532 , Chl ) Φ C 532 λ h C ( λ ) 1 4 π
a p h ( 532 , Chl ) = 0.0113 Ch l 0.871
h C ( λ )  =  W 4 ln 2 π 1 25 exp [ 4 ln 2 ( λ 685 25 ) 2 ] + ( 1 W ) 4 ln 2 π 1 50 exp [ 4 ln 2 ( λ 730 50 ) 2 ]
E f f e c t = [ β t ( 650 , σ , C h l ) β t ( 650 , σ , 0.01 ) ] / β t ( 650 , σ , 0.01 ) × 100 %
E r r o r r = [ c p ( 532 , r 650 / 532 ) c p ( 532 , r 650 / 532 ) ] / c p ( 532 , r 650 / 532 )
E r r o r r = { [ 1 / ( 1 + r 650 / 532 ) ] [ 1 / ( 1 + r 650 / 532 ) ] } / [ 1 / ( 1 + r 650 / 532 ) ] × 100 %
c t = ( 1 + r 650 / 532 ) c p ( 532 , Chl ) + c w t ( λ L , R )
E r r o r f = [ c p ( 532 , Chl ) c p ( 532 , Chl ) g t ] / c p ( 532 , Chl ) g t × 100 %
E r r o r t = [ c p ( 532 , r 650 / 532 , C h l ) c p ( 532 , C h l ) g t ] / c p ( 532 , C h l ) g t × 100 %
a ( λ ) = a w ( λ ) + 0.06 A ( λ ) Ch l 0.65 + a y ( λ )
b ( λ ) = b w ( λ ) + b p ( λ )
c t = exp [ 0.334 ( K l i d a r t ) 2 + 1.916 K l i d a r t 1.540 ]
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.