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

Calibration experiments based on a CO2 absorption cell for the 1.57-µm spaceborne IPDA LIDAR

Open Access Open Access

Abstract

The spaceborne IPDA LIDAR has the potential to measure the global atmosphere CO2 column concentrations with high accuracy. For this kind of LIDAR, system calibration experiments in the laboratory are of high importance. In this study, a specially-customized CO2 absorption cell is employed to simulate the CO2 column absorption of the spaceborne platform. Then calibration experiments are constructed for the receiving system and the entire LIDAR system. The absorption of several different XCO2 concentrations from 400 to 415 ppm in the atmosphere is equivalent to that of the absorption cell charged with different pressures of pure CO2. Under the zero pressure of the absorption cell, the calculated equivalent column average concentration (XCO2) is 12.53 ppm, which acts as system bias. In the calibration experiments, the absolute errors are all less than 1 ppm. And the standard deviations (STDs) are less than 1.1 ppm (148-shot averaging) and 0.8 ppm (296-shot averaging) for receiving system and less than 1.2 ppm and 0.9 ppm for the IPDA LIDAR system. All the results of different average times are close to each other and less than 1 ppm, which proves the high accuracy of the IPDA LIDAR system. In addition, the XCO2 concentrations Allan deviation of 0.25 ppm and 0.35 ppm at 100 s shows that the receiving system and IPDA LIDAR system function with long-term stability. Using a CO2 absorption cell as a standard calibration device in the laboratory validates the measurement accuracy and stability of the spaceborne IPDA LIDAR prototype. Furthermore, the proposed absorption cell may serve as a standard calibration device for related atmosphere trace gases sounding research.

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

1. Introduction

Global surface temperature in the first two decades of the 21st century (2001-2020) was 0.99 [0.84-1.10]°C higher than that in 1850-1900, and in 2011-2020, especially, was 1.09 [0.95-1.20]°C higher [1]. The increased concentrations of Greenhouse gases (GHGs) in the atmosphere are the primary cause of global warming. Regarding the radiative forcing, CO2 contributes 63.5% of all GHGs [2,3]. To develop appropriate emission reduction programs, it is necessary to accurately obtain the sources and sinks of GHGs (such as CO2) globally. Compared to ground-based and airborne observations, spaceborne observations can cover a global scale. At present, the primary observation method for global CO2 measurements is passive spaceborne remote sensing, such as the Greenhouse gases Observing SATellite (GOSAT) and the Carbon Observatory-2 (OCO-2) [4,5]. However, passive spaceborne spectrometers can only work under sunlight and in cloudless conditions, which means massive useless measurement data and large measurement errors [68]. In comparison, the spaceborne integrated path differential absorption (IPDA) light detection and ranging (LIDAR) is a widely recognized instrument currently with high potential in the global and high-precision GHGs column concentrations. Due to its global measurement, all-day operation, high signal-to-noise ratio (SNR), and ability to identify clouds and grounds, CO2 measurement IPDA LIDAR relevant studies have been widely studied, including airborne experiments and feasibility analysis. Several groups had made contributions [911].

Since the Active Sensing of CO2 Emissions over Nights, Days, and Seasons (ASCENDS) project was proposed in 2007 [12], NASA Goddard Space Flight Center (GSFC) has proposed several IPDA LIDAR concepts and related validation experiments. As one of the ASCENDS missions, Abshire et al. developed a multiple-wavelength pulsed IPDA LIDAR system based on a multi-wavelength locked laser. The R16 absorption line of CO2 at 1572.23 nm was selected and the first airborne campaign was implemented in 2011 [9]. In the airborne campaigns in 2014 and 2016, the system was improved with a more sensitive HgCdTe APD detector. The XCO2 concentration was consistent with that from in situ sensor. The measurement accuracy averaging on the desert surfaces within 1s was ∼ 0.7 ppm [13]. NASA Langley Research Center (LaRC) and ITT Exelis developed an intensity-modulated continuous-wave (IM-CW) IPDA LIDAR called multifunctional fiber laser lidar (MFLL) [14]. By conducting multiple airborne campaigns under different environmental conditions based on MFLL [1517], Joel et al. demonstrated that CO2 measurements remained consistent across seasons and had an absolute calibration error (standard deviation) of 0.80 ppm [17]. In addition, NASA LaRC has also developed an airborne double-pulse 2-µm IPDA LIDAR system, and the measured CO2 weighted-average column dry-air volume mixing ratio (XCO2) with an uncertainty of 1.02% and an additional bias of 0.28% was validated [18]. Based on the Advanced Space Carbon and Climate Observation of Planet Earth (A-SCOPE) mission of ESA, Amediek et al. developed a new IPDA LIDAR CHARM-F (CO2 and CH4 Remote Monitoring-Flugzeug) for simultaneous observation of GHGs CO2 and CH4. The results of the first airborne campaign showed an average measurement accuracy of below 0.5% over 20 km [19]. In addition, Fix et al. have studied a high-precision energy monitoring method to reduce the speckle effect of laser, which was important for IPDA LIDAR measurements [20]. The Aerosol and Carbon Detection Lidar (ACDL) developed by Shanghai Institute of Optics and Fine Mechanics (SIOM), Chinese Academy of Sciences was launched in 2022 as the primary payload of the Atmospheric Environment Monitoring Satellite (AEMS). It is the first GHGs spaceborne lidar in space and mainly aims at CO2 column concentrations detection on a global scale with an online wavelength of 1572.024 nm and an offline wavelength of 1572.085 nm [21].

Before the launch of AEMS, the performance of the spaceborne IPDA LIDAR was evaluated and the random errors were less than 0.3% [21]. In addition, Wang et al. demonstrated that the total systematic error was 0.589 ppm and that 61.25% of the global random error was less than 1ppm, which showed its future performance [22]. Zhu et al. accomplished the performance analysis and processing algorithm correction on the airborne flight prototype of the spaceborne IPDA LIDAR system [23]. Then an airborne campaign was carried out in 2019 and 1.02 ppm STD in the open area was validated under atmospheric conditions [24]. Furthermore, the nonlinear correction of the detector module was studied and the detector amplifier circuit board (ACB) was optimized. The simulation results showed that the random error was less than 0.728 ppm in more than 90% of the worldwide regions [25].

However, calibration experiments are significant to verify the measurement accuracy for the spaceborne IPDA LIDAR system. The calibration equipment with an in-door CO2 absorption cell has the potential to simulate the laser absorption of the spaceborne IPDA LIDAR and verify its performance. Sakaizawa et al. tested the performance of the 1.57 µm CO2 measurement LIDAR with a 3-m-long CO2 absorption cell. The result showed good agreement between the CO2 concentration and the measured optical depth (OD), and the correlation between theoretical and experimental OD reached 0.998. These results verified the feasibility of the calibration experiment [26]. However, in the study, the absorption cell cannot simulate the CO2 absorption of the entire atmosphere. Hu et al. designed an absorption cell device that can effectively simulate the CO2 absorption of the atmosphere [27]. Relevant calibration experiments on the receiving system and the IPDA LIDAR system were carried out based on this customized 15.213-m long CO2 absorption cell, respectively. Experimental results confirmed that IPDA LIDAR could achieve 1 ppm measurement accuracy, and the CO2 absorption cell could serve as standard equipment for calibration experiments. As a laboratory calibration test for ACDL, the results of this experiment are highly significant. Meanwhile, the absorption cell calibration equipment can provide a new method for other related researchers to simulate atmosphere GHGs absorption.

In this paper, the basic principle of IPDA LIDAR and the system parameters are introduced briefly in section 2. The experimental setup in detail and measurement results for CO2 absorption cell calibration of the receiving system are shown in section 3. Testing and analysis of the measurement accuracy of the LIDAR system are presented in section 4. In section 5, a summary and outlook of this work are given.

2. Spaceborne IPDA LIDAR system and CO2 absorption cell

2.1 Principle of IPDA LIDAR in the atmosphere

Spaceborne IPDA LIDAR is based on the CO2 differential absorption of the integrated path and selected two wavelengths in the absorption line of CO2 [19]. The online wavelength chooses the wing of the absorption line with strong CO2 absorption. And the offline wavelength locates in the weak absorption area serves as a reference. The IPDA detector receives the reflected echo signals from hard targets such as the ground surface, which is given by the lidar equation as [28]

$${P_e}({{\lambda_{on,off}},{R_G}} )= \eta \cdot O \cdot \frac{A}{{{{({{R_A} - {R_G}} )}^2}}} \cdot \frac{{E({{\lambda_{on,off}}} )}}{{\Delta t({{\lambda_{on,off}}} )}} \cdot \frac{\rho }{\pi } \cdot \exp [{ - ({O{D_m} + O{D_{C{O_2}}}({{\lambda_{on,off}}} )} )} ]\textrm{ ,}$$
where Pe is the echo signal power received by the LIDAR, and E is the energy of the emitted laser, both of which are related to the wavelengths. $\eta $ is the optical efficiency of the receiver and transmitter, O is the overlap function that represents the degree of overlap between the transmitted and received fields of view, and A is the effective area of the receiving telescope. ρ is hard target surface reflectivity, which needs to be divided by π to represent the ratio of the reflected power per steradian and incident power. RA is the altitude of the satellite platform and RG is the altitude of the hard target. Δt is the effective width of the echo pulse. $O{D_m}$ is the dual-path integrated optical depth of aerosols and other atmospheric molecules except for CO2, $O{D_{C{O_2}}}$ is that of CO2. According to the differential absorption of the online and offline wavelengths, the dual-path integrated differential absorption optical depth (DAOD) of CO2 can be deduced and expressed as [23]
$$DAO{D_{C{O_2}}} = 2\int_{{R_G}}^{{R_A}} {{N_{C{O_2}}}(r )\cdot } \Delta {\sigma _{C{O_2}}}({P(r ),T(r )} )dr = \ln \frac{{{P_{off}}{E_{on}}}}{{{P_{on}}{E_{off}}}},$$
$${N_{C{O_2}}}(r )= {\rho _{C{O_2}}}(r )\cdot {N_{dryair}} = \frac{{P(r )\cdot {N_A}}}{{R \cdot T(r )}},$$
where $\Delta {\sigma _{C{O_2}}}$ is the differential absorption cross-section, which is related to the pressure and temperature distributions, and the subscripts of P and E correspond to the online and offline wavelengths, respectively. P is the pressure and T is the temperature. ${N_{C{O_2}}}$ and ${N_{dryair}}$ are the number density of CO2 molecules and dry air, ${\rho _{C{O_2}}}$ is dry air mixing ratio of CO2, ${N_A}$ is Avogadro’s number, and $R$ is the gas constant.

The vertical-path XCO2 is given as:

$$XC{O_2} = \frac{{\frac{1}{2}\ln \frac{{{P_{off}}{E_{on}}}}{{{P_{on}}{E_{off}}}}}}{{IWF}},$$
$$IWF = \int_{{R_G}}^{{R_A}} {\frac{{P(r )\cdot {N_A} \cdot \Delta {\sigma _{C{O_2}}}({P(r ),T(r )} )}}{{R \cdot T(r )\cdot ({1 + {\rho_{{H_2}O}}(r )} )}}} dr,$$
where ${\rho _{{H_2}O}}$ is dry air mixing ratio of water vapor, and IWF is the integrated weight function.

2.2 IPDA LIDAR system configuration

The IPDA LIDAR of ACDL is primarily divided into a laser transmitting system, telescope, photoelectric detection system, data acquisition system, and control electronics system. The laser transmitter system consists of a 1572 nm seed laser, frequency stabilization system, pulsed laser, and optical components, resulting in a 1572 nm double-pulse laser at 20 Hz. Using external frequency modulation technique and optical phase-locked loop (OPLL) of a phase-locked system, the seed laser provides continuous wave online and offline laser signals with frequency stability of less than 0.3 MHz (root mean square (RMS)) [29]. A frequency switching output between the online and offline seed lasers is achieved at 200 µs intervals using a magneto-optical switch. Then, pulsed laser is generated based on seed injection OPO (optical parametric oscillator) technology and a frequency stabilization system. The frequency stability is less than 0.6 MHz (RMS) [30]. Finally, OPA (optical parametric amplifier) amplifies the laser energy to above 75 mJ and 35 mJ, for online and offline radiation respectively. The 1572-nm radiation is then divided into two parts by beam splitter mirror 1 (BSM1) after beam pointing control. One part enters the atmosphere through the beam expander, and the other enters the integrating sphere (IS) energy monitoring module. Backscattered echo signals from targets and monitoring signals from IS are ultimately focused on the same InGaAs APD detector in the photoelectric detection system. A filter F2 with a bandwidth of 0.45 nm is inserted in front of the APD detector to block background light. Figure 1. shows the schematic diagram of the IPDA LIDAR system, and Table 1 lists the main parameters of the spaceborne IPDA LIDAR.

 figure: Fig. 1.

Fig. 1. Schematic diagram of the spaceborne IPDA LIDAR system. RM1, 2: reflecting mirrors; BSM1, 2: beam splitter mirrors; IS: integrating sphere; CL1, 2: collimating lens; LPC: laser pointing control; F1, 2: filter; APD: Avalanche photodiode.

Download Full Size | PDF

Tables Icon

Table 1. Main parameters of spaceborne IPDA LIDAR.

2.3 CO2 absorption cell configuration

The free space CO2 absorption cell is specially-designed to simulate the LIDAR beam CO2 absorption of the atmospheric column. When the absorption cell is filled, the absorption cross-section is only related to the pressure and temperature inside. The path integral of the DAOD can convert into the product of the differential absorption cross-section and the effective optical length of the absorption cell (L). DAOD of the absorption cell can be expressed as

$$DAO{D_{C{O_2}}} = \frac{{P \cdot {N_A}}}{{R \cdot T}} \cdot \Delta {\sigma _{C{O_2}}} \cdot L,$$

The temperature and pressure of the in-door CO2 absorption cell are controllable to simulate the different DAOD of CO2 in the atmosphere. When the absorption cell is at a fixed temperature, the pressure of pure CO2 is the only variable that causes the change of DAOD. The US standard atmosphere model [31] and the spectroscopy database HITRAN 2020 [32] are used to simulate the DAOD and IWF. Several different XCO2 around 400 ppm in the atmosphere can be equivalent to the various pressures charged to the absorption cell. The inversion values of the equivalent XCO2 at different pressures of the absorption cell can verify the accuracy of IPDA LIDAR.

An absorption cell with a 15.213-m length was designed to simulate the DAOD of CO2 from a spaceborne platform according to optimized parameters. The absorption cell device mainly includes a CO2 absorption cell pipe, an optical path turning structure, a temperature control system, a pressure detection system, a vacuuming system, and a CO2 gas charging system [27]. The absorption cell is folded into three parts via two 45° deflection mirrors to reduce the space. The layout of the CO2 absorption cell system is shown in Fig. 2. Table 2 shows the major parameters of the absorption cell, in which several real-time temperature and pressure control points are set.

 figure: Fig. 2.

Fig. 2. The schematic layout of CO2 absorption cell

Download Full Size | PDF

Tables Icon

Table 2. Parameters of the absorption cell.

3. IPDA LIDAR receiver system calibration experiment

3.1 Receiver calibration experimental equipment

An optical double pulse signal simulator based on the acoustic-optical modulator (AOM) modulation is utilized to verify the performance of IPDA LIDAR. The time interval between the double pulses is 200 µs, and the generated double-pulse laser is divided into two parts with a 2 × 2 fiber optical coupler (FOC). One part is delayed by a 3 km long single mode fiber and then enters the CO2 absorption cell to simulate the total atmosphere layer CO2 absorption of the spaceborne platform. The other part is collimated directly and received by the APD detector of the IPDA LIDAR receiving system through the telescope, which acts as an energy monitoring part. The digital adjustable attenuators of AA1 and AA2 are used to control the optical power of the LIDAR system. The timing relationship of the entire experimental equipment is controlled by a multi-channel signal generator, and the schematic diagram of the whole experimental setup is shown in Fig. 3.

 figure: Fig. 3.

Fig. 3. Schematic diagram of absorption cell calibration equipment for IPDA LIDAR based on an optical signal simulator. AOM, acoustic-optic modulator; FOC, fiber optical coupler; AA1, adjustable attenuator 1; AA2, adjustable attenuator 2; FC1, fiber collimator 1; FC2, fiber collimator 2; RM1, reflecting mirror 1; RM2, reflecting mirror 2; SM fiber, 3 km single mode fiber.

Download Full Size | PDF

When XCO2 changed from 400 to 415 ppm with a step of 5 ppm, DAOD and IWF were calculated according to Eq. (2) and Eq. (5). Table 3 gives the pressure corresponding to different XCO2, and the accuracy of the pressure is 1 hPa. The temperature was fixed at 25°C.

Tables Icon

Table 3. XCO2 and corresponding theoretical pressure.

Since the AOM has shifted the frequency online and offline by about 200 MHz, the corresponding DAOD and IWF have changed. During the experiment, CO2 with different pressures was charged into the absorption cell separately. The theoretical results were compared with the experimental results under the online and offline conditions of frequency shift. Table 4 gives the corresponding theoretical equivalent XCO2 (ppm) under several different pressures. According to the simulation, the change in IWF is ignored.

Tables Icon

Table 4. The actual pressure and corresponding theoretical XCO2 at shifted online and offline wavelengths.

3.2 Receiver calibration experimental results

When the pressure in the absorption cell is almost zero, the four signals detected by the detector are shown in Fig. 4.

 figure: Fig. 4.

Fig. 4. Four signals received by the lidar receiver.

Download Full Size | PDF

Since the spatial resolutions requirements of land and ocean are 50 km and 100 km, respectively, corresponding to an accumulated pulse number of 148 and 296, which last roughly 7 s and 14 s on the time scale, the analysis in this paper is based on the average of 148 pulses and 296 pulses. The relationship between the SNR of a single-shot and that of a multi-shot is

$$SN{R_m} = SN{R_s} \cdot \sqrt {{N_{shots}}} ,$$
where Nshots is the number of pulses and SNRs/m represents the SNR of single-shot/multi-shot.

A double-path DAOD with more than a 10000-shot (500 s) sample is shown in Fig. 5(a), while Fig. 5(b) shows the results of the 148-shot moving average. Without considering the error of IWF, Fig. 5(c) and Fig. 5(d) show the results of the single-shot and the 148-shot moving average, respectively.

 figure: Fig. 5.

Fig. 5. (a) Single-shot DAOD. (b) DAOD of 148-shot moving average. (c) Single-shot XCO2. (d) XCO2 of 148-shot moving average in case of more than a 10000-shot sample.

Download Full Size | PDF

The values of XCO2 with zero pressure inside the CO2 absorption cell are 12.53 ppm with STDs of 0.74 ppm for 148-shot moving average and 0.59 ppm for 296-shot moving average, which is a system bias parameter for subsequence measurement correction. The nonzero DAOD bias is mainly induced by the filter transmission difference for the online and offline signals at different wavelengths and different incident angles. The beam pointing is adjusted in advance and monitored in real-time during the experiment to ensure stable beam pointing. In addition, the testing time is determined to be chosen at midnight when the environment disturbance is weak to ensure the relative stability of the beam pointing. Thus, the system bias can maintain a relatively constant value. But the error caused by the spatial optical path of about 20 m outside the absorption cell is negligible.

After the CO2 absorption cell is operated at different pressures to simulate CO2 absorption in the atmosphere, the measurement results are inverted and analyzed. When charging with different pressures, the SNR of the measurement results for the 148-shot and 296-shot moving average lists in Table 5. The results meet the requirements of SNR on the spaceborne platform [10,21,23].

Tables Icon

Table 5. SNR for different pressures of CO2.

For the measurement data with the lowest SNR at 466.8 kPa pressure, the results of the “pulse peak method (PPM)” and the “pulse integration method (PIM)” were compared. PPM uses the peak of the pulses to represent the pulse power, while PIM uses the integral values of each point. Besides, the methods of “log before averaging (AVS)” and “log after averaging (AVD)” were also compared [24,33,34]. The methods of AVS and AVD can be expressed as

$${\overline {XC{O_2}} ^{AVS}} = \frac{{\frac{1}{2} \cdot \ln \left\langle {\overline {{V_i}} } \right\rangle }}{{\left\langle {IWF} \right\rangle }},$$
$$\overline {{V_i}} = \frac{{\left\langle {{V_{off}}} \right\rangle \cdot \left\langle {{V_{o{n_0}}}} \right\rangle }}{{\left\langle {{V_{on}}} \right\rangle \cdot \left\langle {{V_{of{f_0}}}} \right\rangle }},$$
$${\overline {XC{O_2}} ^{AVD}} = \frac{{\left\langle {DAOD} \right\rangle }}{{\left\langle {IWF} \right\rangle }}.$$
where $\overline {{V_i}} $ is the average of each signal before log, and the triangular bracket notation denotes the arithmetic mean. The subscripts on and off represent the echo signal, and on0 and off0 represent the monitoring signal. In Fig. 6(a) and 6(b), the DAOD and XCO2 calculated by PPM and PIM for single-shot are compared, respectively. From the statistics results, as shown in Table 6, the peak-to-peak values based on PPM are smaller than that of PIM, so the PPM method is selected in subsequent calculations. Figure 6(c) compares the XCO2 calculated by AVS and AVD for 148-shot moving average, respectively, and Table 7 summarizes the results. The difference in STD is negligible, and the methods of PPM and AVS are utilized in the calibration experiments.

 figure: Fig. 6.

Fig. 6. (a) DAOD for single-shot calculation with PPM and PIM methods. (b) XCO2 for single-shot calculation. (c) XCO2 for 148-shot moving average with AVS and AVD methods.

Download Full Size | PDF

Tables Icon

Table 6. Statistical results of PPM and PIM.

Tables Icon

Table 7. Statistical XCO2 of AVS and AVD for 148-shot moving average.

The summary of the absolute errors and STDs is shown in Fig. 7 and Table 8. The absolute errors of XCO2 meet the accuracy of 1 ppm and remain almost stable for different pulse averaging numbers. It confirms that IPDA LIDAR’s receiving system can achieve a measurement accuracy of 1ppm. The STDs are around 1ppm with 148-shot moving average and almost 0.6 ppm with 296-shot moving average. They show that the system STD can remain stable within the spatial resolution. The accuracy of the IPDA LIDAR receiving system is therefore confirmed and the calibration system performance is validated.

 figure: Fig. 7.

Fig. 7. (a) The inversion results XCO2 of 148-shot moving average after bias correction. (b) The inversion results XCO2 of 296-shot moving average after bias correction.

Download Full Size | PDF

Tables Icon

Table 8. Results of calibration experiments of IPDA receiver system.

3.3 System stability analysis

For the absorption cell with CO2 at a fixed pressure of 439 hPa, the Allan deviation of the inversion results was analyzed. Figure 8(a) shows the XCO2 inversion results for a single-shot calculated for 800 s, and Fig. 8(b) is the corresponding Allan deviation. The corresponding Allan deviation for 7 s, 14 s, and 100 s are approximately 1.07 ppm, 0.78 ppm, and 0.25 ppm, respectively. The Allan deviation decreases with time, and the LIDAR receiving system and calibration system are validated to operate stably over a long period.

 figure: Fig. 8.

Fig. 8. (a) The inversion results XCO2 of single-shot after bias correction. (b) Allan deviation of the inversion results.

Download Full Size | PDF

The error of the calibration system, regardless of the ACDL IPDA system itself, mainly comes from the optical path. The optical path affects the online and offline beam pointing fluctuations, which results in different transmissions through the filter before the detector. And the error analysis of IPDA LIDAR itself has been analyzed in detail in Ref. [12]. The STD increases from 0.74 ppm to ∼1 ppm (148-shot moving average). Compared to the zero pressure state of the absorption cell and the state with a certain amount of CO2, the airflow in the absorption cell also causes jitter in the beam pointing. In general, the calibration system can meet the measurement accuracy requirements.

4. IPDA LIDAR system calibration experiment

4.1 LIDAR system calibration experimental equipment

After testing the performance of the receiving system, it’s necessary to validate the performance of the entire IPDA LIDAR system using the CO2 absorption cell setup. A relative calibration experiment for the whole IPDA LIDAR system was conducted. Figure 9 shows the schematic diagram of the IPDA LIDAR calibration system. in which reflecting mirror 1 (RM1) is a total-reflection mirror, RM2 is an anti-reflection (AR) mirror, and RM3 is a high-reflection (HR) mirror. The beam splitter (BSM) divides the pulsed laser into two parts. One part enters the absorption cell through RM1, RM2, and RM3 to simulate the absorption of CO2 in the atmosphere, and the other passes through the IS for transmitting energy picking. The output of the IS is delivered via a 1 km multimode fiber (MM fiber), which produced a 5 µs delay for monitoring pulse to separate the echo and monitoring pulse in time. Both optical signals are received by the same APD detector of the LIDAR system. In the LIDAR system, the IS for energy monitoring is devoted to reducing the impact of pointing jitter and speckle, and an error of less than 0.2 ppm is validated [20,35].

 figure: Fig. 9.

Fig. 9. Schematic diagram of IPDA LIDAR calibration system. LA, light absorber; RM1, 2, 3, 4, 5, reflecting mirror; MM fiber, 1 km multimode fiber; IS, integrating sphere; BSM, beam splitter mirror.

Download Full Size | PDF

4.2 LIDAR system calibration experiment results

First, the theoretical XCO2 based on the actual pressure was calculated. Table 9 shows the relationship between the charged pressure and the theoretically corresponding equivalent XCO2. The original four pulse signals are shown in Fig. 10.

Tables Icon

Table 9. The actual pressure and corresponding theoretical XCO2 at 1572.024 nm and 1572.085 nm.

 figure: Fig. 10.

Fig. 10. Four signals received by the APD detection.

Download Full Size | PDF

 figure: Fig. 11.

Fig. 11. (a) The inversion results XCO2 of 148-shot moving average after bias correction. (b) The inversion results XCO2 of 296-shot moving average after bias correction.

Download Full Size | PDF

PPM and AVS methods are adopted to invert the equivalent XCO2. The beam pointing is adjusted to be constant in the experiment and monitored in real-time by other optical systems, so the bias under zero pressure acts as a fixed bias. The statistics of the final inversion results of 148-shot moving average and 296-shot moving average are shown in Fig. 11. The inversed XCO2 increases with the increasing CO2 pressure in the absorption cell equipment. The bias and STDs are counted in the XCO2 inversion results and shown in Table 10. For the CO2 absorption cell charged with different pressures, the differences between the inversion results and the simulated theoretical values are less than 1 ppm. Under the accuracy requirements of 50 km resolution (148 shots average) on the land and 100 km resolution (296 shots average) on the ocean, it can be satisfied.

Tables Icon

Table 10. Results of calibration experiments of IPDA LIDAR system.

Comparing the calibration results of the receiving system and the LIDAR system, they have in common the absolute error of less than 1ppm, and the STD of 148-shot moving average is almost close to 1 ppm. The absolute bias verifies the absorption cell’s ability as a standard calibration device and the IPDA LIDAR measurement accuracy. STDs are less than 1.2 ppm with 148-pulse averaged. The main reason for the different STDs is the effect of the spatial optical path. The mirrors used in the calibration system cause inevitably stray light interference, which affects the stability of the beam pointing. In calibration experiments, it is important to reduce the parasitic laser pulse interference.

4.3 System stability analysis

For the absorption cell with CO2 at a fixed pressure of 467.2 hPa, the Allan deviation of the XCO2 inversion result was analyzed. As the results in Fig. 12(b) show, the Allan deviation decreases monotonically, indicating that the calibration system can maintain stability over time. The corresponding Allan deviations for 7 s, 14 s, and 100 s are 1.27 ppm, 0.91 ppm, and 0.35 ppm, respectively.

 figure: Fig. 12.

Fig. 12. (a) The inversion results XCO2 of single-shot after bias correction. (b) Allan variance of the inversion results XCO2.

Download Full Size | PDF

The absolute errors of both calibration systems are less than 1ppm. It proved that the calibration systems based on a CO2 absorption cell with changed pressure functioned with high accuracy. In addition, it can be an effective setup for IPDA LIDAR calibration on the ground. The spatial optical path of the calibration system needs to be further optimized to reduce the jitter of the beam pointing.

5. Conclusions

In this study, a specially-customized CO2 absorption cell in the laboratory was utilized to calibrate and validate the performance of the spaceborne IPDA LIDAR. It can simulate the atmospheric CO2 absorption of the spaceborne platform. Using the CO2 absorption cell, the calibration experiments for the receiving system and the entire IPDA LIDAR system were constructed. The measurement concentrations set in the experiment ranged from 400 to 415 ppm with a step size of 5 ppm. By establishing the relationship between XCO2 concentrations and the pressure inside the CO2 absorption cell, the performance of IPDA LIDAR was analyzed.

In the calibration experiment of the receiving system, an optical double-pulse signal simulator is used to simulate a 20 Hz double-pulse laser. The equivalent XCO2 concentration that acted as system bias is 12.53 ppm. The STDs are 0.74 ppm and 0.59 ppm with 148-shot and 296-shot moving averages for the experiments, respectively. The PPM and AVS methods were utilized in inversions. The inversion results show that the absolute bias between the inversion results and the theoretical values is less than 1 ppm for all measurements. The STDs are less than 1.1 ppm and 0.8 ppm using 148-shot and 296-shot moving averages, respectively. The XCO2 Allan deviation calculated for 100 s is 0.25 ppm, which proves the long-term stability of the IPDA LIDAR receiving system.

In the calibration experiment of the entire IPDA LIDAR system, the experimental results of different CO2 concentrations were validated. The absolute errors are less than 1 ppm for all measurements, and the STDs are less than 1.2 ppm and 0.9 ppm using 148-shot and 296-shot moving averages, respectively. The XCO2 Allan deviation for an accumulation time of 100 s is validated to be 0.35 ppm for the ACDL IPDA system.

The accuracy and long-term stability of the IPDA LIDAR system used for XCO2 measurements are validated. In addition, a CO2 absorption cell was creatively designed and customized to simulate the CO2 absorption of the spaceborne platform, which can provide a new method IPDA LIDAR of calibration for relevant researchers. In future studies, we plan to reduce the influence of beam pointing in the spatial optical path of the experimental system.

Funding

ACDL LIDAR project; China National Space Administration.

Disclosures

The authors declare no conflicts of interest.

Data availability

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

References

1. V. MassonDelmotte, P. Zhai, A. Pirani, S. L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R. Matthews, T. K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou, eds., IPCC, 2021: Summary for Policymakers. In: Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. (Cambridge University, In Press).

2. T. F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S. K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, and P. M. Midgley, eds., IPCC, 2013: Summary for Policymakers. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. (Cambridge University, 2013).

3. D. J. Jacob, Introduction to Atmospheric Chemistry. (Princeton University, 1999).

4. T. Yokota, Y. Yoshida, N. Eguchi, Y. Ota, T. Tanaka, H. Watanabe, and S. Maksyutov, “Global concentrations of CO2 and CH4 retrieved from GOSAT: First preliminary results,” SOLA 5, 160–163 (2009). [CrossRef]  

5. C. Frankenberg, R. Pollock, R. Lee, R. Rosenberg, J. F. Blavier, D. Crisp, C. O’Dell, G. Osterman, C. Roehl, P. Wennberg, and D. Wunch, “The Orbiting Carbon Observatory (OCO-2): Spectrometer performance evaluation using pre-launch direct sun measurements,” Atmos. Meas. Tech. Discuss. 7, 7641–7670 (2014). [CrossRef]  

6. S. Basu, S. Guerlet, A. Butz, S. Houweling, O. Hasekamp, I. Aben, P. Krummel, P. Steele, R. Langenfelds, M. Torn, S. Biraud, B. Stephens, A. Andrews, and D. Worthy, “Global CO2 fluxes estimated from GOSAT retrievals of total column CO2,” Atmos. Chem. Phys. 13(17), 8695–8717 (2013). [CrossRef]  

7. A. Butz, S. Guerlet, O. Hasekamp, et al., “Toward accurate CO2 and CH4 observations from GOSAT,” Geophys. Res. Lett. 38(14), L14812 (2011). [CrossRef]  

8. D. Wunch, P. O. Wennberg, G. Osterman, et al., “Comparisons of the Orbiting Carbon Observatory-2 (OCO-2) XCO2 measurements with TCCON,” Atmos. Meas. Tech. 10(6), 2209–2238 (2017). [CrossRef]  

9. J. B. Abshire, A. Ramanathan, H. Riris, J. Mao, G. R. Allan, W. E. Hasselbrack, C. J. Weaver, and E. V. Browell, “Airborne Measurements of CO2 Column Concentration and Range Using a Pulsed Direct-Detection IPDA Lidar,” Remote Sens. 6(1), 443–469 (2013). [CrossRef]  

10. G. Han, H. Xu, W. Gong, J. Liu, J. Du, X. Ma, and A. Liang, “Feasibility Study on Measuring Atmospheric CO2 in Urban Areas Using Spaceborne CO2-IPDA LIDAR,” Remote Sens. 10(7), 985 (2018). [CrossRef]  

11. R. T. Menzies and D. M. Tratt, “Differential Laser absorption spectrometry for global profiling of tropospheric carbon dioxide: selection of optimum sounding frequencies for high-precision measurements,” Appl. Opt. 42(33), 6569–6577 (2003). [CrossRef]  

12. S. Kawa, J. Abshire, D. Baker, E. Browell, D. Crisp, S. Crowell, J. Hyon, J. Jacob, K. W. Jucks, B. Lin, R. Menzies, L. Ott, and S. Zaccheo, “Active Sensing of CO2 Emissions over Nights, Days, and Seasons (ASCENDS): Final Report of the ASCENDS Ad Hoc Science Definition Team,” NASA/TP-2018-219034 (2018).

13. J. B. Abshire, A. K. Ramanathan, H. Riris, G. R. Allan, X. Sun, W. E. Hasselbrack, J. Mao, S. Wu, J. Chen, K. Numata, S. R. Kawa, M. Y. M. Yang, and J. DiGangi, “Airborne measurements of CO2 column concentrations made with a pulsed IPDA lidar using a multiple-wavelength-locked laser and HgCdTe APD detector,” Atmos. Meas. Tech. 11(4), 2001–2025 (2018). [CrossRef]  

14. J. T. Dobler, F. W. Harrison, E. V. Browell, B. Lin, D. McGregor, S. Kooi, Y. Choi, and S. Ismail, “Atmospheric CO2 column measurements with an airborne intensity-modulated continuous wave 1.57 mu m fiber laser lidar,” Appl. Opt. 52(12), 2874–2892 (2013). [CrossRef]  

15. B. Lin, A. R. Nehrir, F. W. Harrison, E. V. Browell, S. Ismail, M. D. Obland, J. Campbell, J. Dobler, B. Meadows, T.-F. Fan, and S. Kooi, “Atmospheric CO2 column measurements in cloudy conditions using intensity-modulated continuous-wave lidar at 1.57 micron,” Opt. Express 23(11), A582–A593 (2015). [CrossRef]  

16. B. Lin, E. Browell, S. Kooi, J. Dobler, J. Campbell, T. Fan, S. Pal, C. O’Dell, M. Obland, W. Erxleben, D. McGregor, I. Gordon, R. Kochanov, J. DiGangi, K. Davis, and Y. Choi, “Assessment of MFLL column CO2 measurements obtained during the ACT-America field campaigns,” 2017 American Geophysical Union Fall Meeting, (2017).

17. J. F. Campbell, B. Lin, J. Dobler, S. Pal, K. Davis, M. D. Obland, W. Erxleben, D. McGregor, C. O’Dell, E. Bell, B. Weir, T.-F. Fan, S. Kooi, I. Gordon, A. Corbett, and R. Kochanov, “Field Evaluation of Column CO2 Retrievals From Intensity-Modulated Continuous-Wave Differential Absorption Lidar Measurements During the ACT-America Campaign,” Earth Space Sci. 7(12), e2019EA000847 (2020). [CrossRef]  

18. T. F. Refaat, U. N. Singh, J. Yu, M. Petros, R. Remus, and S. Ismail, “Double-pulse 2 µm integrated path differential absorption lidar airborne validation for atmospheric carbon dioxide measurement,” Appl. Opt. 55(15), 4232 (2016). [CrossRef]  

19. A. Amediek, G. Ehret, A. Fix, M. Wirth, C. Büdenbender, M. Quatrevalet, C. Kiemle, and C. Gerbig, “CHARM-F-a new airborne integrated-path differential-absorption lidar for carbon dioxide and methane observations: measurement performance and quantification of strong point source emissions,” Appl. Opt. 56(18), 5182–5197 (2017). [CrossRef]  

20. A. Fix, M. Quatrevalet, A. Amediek, and M. Wirth, “Energy calibration of integrated path differential absorption lidars,” Appl. Opt. 57(26), 7501–7514 (2018). [CrossRef]  

21. G. Han, X. Ma, A. Liang, T. Zhang, Y. Zhao, M. Zhang, and W. Gong, “Performance Evaluation for China’s Planned CO2-IPDA,” Remote Sens. 9(8), 768 (2017). [CrossRef]  

22. S. Wang, J. Ke, S. Chen, Z. Zheng, C. Cheng, B. Tong, J. Liu, D. Liu, and W. Chen, “Performance Evaluation of Spaceborne Integrated Path Differential Absorption Lidar for Carbon Dioxide Detection at 1572 nm,” Remote Sens. 12(16), 2570 (2020). [CrossRef]  

23. Y. Zhu, J. Liu, X. Chen, X. Zhu, D. Bi, and W. Chen, “Sensitivity analysis and correction algorithms for atmospheric CO2 measurements with 1.57-microm airborne double-pulse IPDA LIDAR,” Opt. Express 27(22), 32679–32699 (2019). [CrossRef]  

24. Y. Zhu, J. Yang, X. Chen, X. Zhu, J. Zhang, S. Li, Y. Sun, X. Hou, D. Bi, L. Bu, Y. Zhang, J. Liu, and W. Chen, “Airborne Validation Experiment of 1.57-µm Double-Pulse IPDA LIDAR for Atmospheric Carbon Dioxide Measurement,” Remote Sens. 12(12), 1999 (2020). [CrossRef]  

25. Y. Zhu, J. Yang, X. Zhang, J. Liu, X. Zhu, H. Zang, T. Xia, C. Fan, X. Chen, Y. Sun, X. Hou, and W. Chen, “Performance Improvement of Spaceborne Carbon Dioxide Detection IPDA LIDAR Using Linearty Optimized Amplifier of Photo-Detector,” Remote Sens. 13(10), 2007 (2021). [CrossRef]  

26. D. Sakaizawa, S. Kawakami, M. Nakajima, Y. Sawa, and H. Matsueda, “Ground-based demonstration of a CO2 remote sensor using a 1.57 µm differential laser absorption spectrometer with direct detection,” J. Appl. Remote Sens. 4(1), 043548 (2010). [CrossRef]  

27. J. Q. Liu, W. Y. Hu, W. B. Chen, X. Hou, X. P. Zhu, and D. C. Bi, “Carbon dioxide concentration measurement and calibration system,” China patent CN201811190918.8 (12 October 2018).

28. T. F. Refaat, U. N. Singh, J. Yu, M. Petros, S. Ismail, M. J. Kavaya, and K. J. Davis, “Evaluation of an airborne triple-pulsed 2 µm IPDA lidar for simultaneous and independent atmospheric water vapor and carbon dioxide measurements,” Appl. Opt. 54(6), 1387–1398 (2015). [CrossRef]  

29. J. Du, Y. Sun, D. Chen, Y. Mu, M. Huang, Z. Yang, J. Liu, D. Bi, X. Hou, and W. Chen, “Frequency-stabilized laser system at 1572 nm for space-borne CO2 detection LIDAR,” Chin. Opt. Lett. 15(3), 031401 (2017). [CrossRef]  

30. X. Chen, X.-L. Zhu, S.-G. Li, X.-H. Ma, W. Xie, J.-Q. Liu, W.-B. Chen, and R. Zhu, “Frequency Stabilization of Pulsed Injection-Seeded OPO Based on Optical Heterodyne Technique,” Chin. Phys. Lett. 35(2), 024201 (2018). [CrossRef]  

31. G. Anderson, S. Clough, F. Kneizys, J. Chetwynd, and E. Shettle, “AFGL Atmospheric Constituent Profiles (0-120 km),” Air Force Geophysics Lab. 46 (1986).

32. I. E. Gordon, L. S. Rothman, R. J. Hargreaves, et al., “The HITRAN2020 molecular spectroscopic database,” J. Quant. Spectrosc. Radiat. Transfer 277, 107949 (2022). [CrossRef]  

33. M. J. T. Milton and P. T. Woods, “Pulse averaging methods for a laser remote monitoring system using atmospheric backscatter,” Appl. Opt. 26(13), 2598–2603 (1987). [CrossRef]  

34. E. P. MacKerrow, M. J. Schmitt, and D. C. Thompson, “Effect of speckle on lidar pulse–pair ratio statistics,” Appl. Opt. 36(33), 8650–8669 (1997). [CrossRef]  

35. W. Hu, J. Liu, Y. Zhu, J. Dong, X. Ma, S. Li, J. Zhang, X. Zhu, and W. Chen, “Analysis of energy monitoring for a double-pulsed CO2 integrated path differential absorption lidar at 1.57 µm,” Appl. Opt. 58(3), 616–625 (2019). [CrossRef]  

Data availability

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

Cited By

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

Alert me when this article is cited.


Figures (12)

Fig. 1.
Fig. 1. Schematic diagram of the spaceborne IPDA LIDAR system. RM1, 2: reflecting mirrors; BSM1, 2: beam splitter mirrors; IS: integrating sphere; CL1, 2: collimating lens; LPC: laser pointing control; F1, 2: filter; APD: Avalanche photodiode.
Fig. 2.
Fig. 2. The schematic layout of CO2 absorption cell
Fig. 3.
Fig. 3. Schematic diagram of absorption cell calibration equipment for IPDA LIDAR based on an optical signal simulator. AOM, acoustic-optic modulator; FOC, fiber optical coupler; AA1, adjustable attenuator 1; AA2, adjustable attenuator 2; FC1, fiber collimator 1; FC2, fiber collimator 2; RM1, reflecting mirror 1; RM2, reflecting mirror 2; SM fiber, 3 km single mode fiber.
Fig. 4.
Fig. 4. Four signals received by the lidar receiver.
Fig. 5.
Fig. 5. (a) Single-shot DAOD. (b) DAOD of 148-shot moving average. (c) Single-shot XCO2. (d) XCO2 of 148-shot moving average in case of more than a 10000-shot sample.
Fig. 6.
Fig. 6. (a) DAOD for single-shot calculation with PPM and PIM methods. (b) XCO2 for single-shot calculation. (c) XCO2 for 148-shot moving average with AVS and AVD methods.
Fig. 7.
Fig. 7. (a) The inversion results XCO2 of 148-shot moving average after bias correction. (b) The inversion results XCO2 of 296-shot moving average after bias correction.
Fig. 8.
Fig. 8. (a) The inversion results XCO2 of single-shot after bias correction. (b) Allan deviation of the inversion results.
Fig. 9.
Fig. 9. Schematic diagram of IPDA LIDAR calibration system. LA, light absorber; RM1, 2, 3, 4, 5, reflecting mirror; MM fiber, 1 km multimode fiber; IS, integrating sphere; BSM, beam splitter mirror.
Fig. 10.
Fig. 10. Four signals received by the APD detection.
Fig. 11.
Fig. 11. (a) The inversion results XCO2 of 148-shot moving average after bias correction. (b) The inversion results XCO2 of 296-shot moving average after bias correction.
Fig. 12.
Fig. 12. (a) The inversion results XCO2 of single-shot after bias correction. (b) Allan variance of the inversion results XCO2.

Tables (10)

Tables Icon

Table 1. Main parameters of spaceborne IPDA LIDAR.

Tables Icon

Table 2. Parameters of the absorption cell.

Tables Icon

Table 3. XCO2 and corresponding theoretical pressure.

Tables Icon

Table 4. The actual pressure and corresponding theoretical XCO2 at shifted online and offline wavelengths.

Tables Icon

Table 5. SNR for different pressures of CO2.

Tables Icon

Table 6. Statistical results of PPM and PIM.

Tables Icon

Table 7. Statistical XCO2 of AVS and AVD for 148-shot moving average.

Tables Icon

Table 8. Results of calibration experiments of IPDA receiver system.

Tables Icon

Table 9. The actual pressure and corresponding theoretical XCO2 at 1572.024 nm and 1572.085 nm.

Tables Icon

Table 10. Results of calibration experiments of IPDA LIDAR system.

Equations (10)

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

P e ( λ o n , o f f , R G ) = η O A ( R A R G ) 2 E ( λ o n , o f f ) Δ t ( λ o n , o f f ) ρ π exp [ ( O D m + O D C O 2 ( λ o n , o f f ) ) ]  ,
D A O D C O 2 = 2 R G R A N C O 2 ( r ) Δ σ C O 2 ( P ( r ) , T ( r ) ) d r = ln P o f f E o n P o n E o f f ,
N C O 2 ( r ) = ρ C O 2 ( r ) N d r y a i r = P ( r ) N A R T ( r ) ,
X C O 2 = 1 2 ln P o f f E o n P o n E o f f I W F ,
I W F = R G R A P ( r ) N A Δ σ C O 2 ( P ( r ) , T ( r ) ) R T ( r ) ( 1 + ρ H 2 O ( r ) ) d r ,
D A O D C O 2 = P N A R T Δ σ C O 2 L ,
S N R m = S N R s N s h o t s ,
X C O 2 ¯ A V S = 1 2 ln V i ¯ I W F ,
V i ¯ = V o f f V o n 0 V o n V o f f 0 ,
X C O 2 ¯ A V D = D A O D I W F .
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.