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

Error analysis for lidar retrievals of atmospheric species from absorption spectra

Open Access Open Access

Abstract

We generalize a model for retrieving atmospheric constituents from lidar absorption spectra measured at any laser frequency channels. Random and systematic retrieval errors from measurement noise and model bias, respectively, are analyzed parametrically and numerically to provide deeper insight. By placing four or more channels symmetrically around the absorption peak, retrieval errors from a common laser frequency shift and spectral baseline tilt can be eliminated. By solving for the frequency shift and spectral baseline tilt, atmospheric retrievals degrade only slightly even when such channels are shifted substantially out of symmetry. An etalon can thus be used for the wavelength stabilization.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

This paper continues our study on lidar retrievals of atmospheric constituents from absorption spectra. In a previous study [1] we established a model for the ideal case where the laser frequency channels are placed symmetrically around the center of the target absorption line profile. The present paper generalizes the model for any distribution of the laser frequency channels across the absorption line profile.

Future lidar missions call for unprecedented high precision for global measurement of atmospheric constituents and parameters [24]. For example, the Active Sensing of CO2 Emissions over Nights, Days, and Seasons (ASCENDS) mission planned by NASA [2,3] aims to measure the global distribution of carbon dioxide (CO2) mixing ratios (∼400 ppm) to ∼1 ppm precision. To meet such stringent requirements, nadir-viewing, direct-detection, and pulsed integrated-path differential-absorption (IPDA) lidar techniques are being developed to measure the two-way optical absorption spectra of the target species from the spacecraft to the surface and back [5,6]. From the absorption spectra and altimetry measurements and other ancillary data of the atmosphere, the dry mixing ratios of the target species can be retrieved [1,710].

Throughout this paper, a candidate ASCENDS IPDA lidar approach [6] being developed at NASA is used as a concrete example. As shown in Fig. 1, this approach uses a pulsed laser to rapidly cycle through multiple laser frequency channels across a single CO2 line at 1572.335 nm [11]. The returned laser pulses are detected directly with a nearly noiseless HgCdTe avalanche photodiode (APD) detector [12], to maximize the signal to noise ratio [13]. To reduce retrieval errors arising from the laser frequency fluctuation, laser pulses at each fixed wavelength are carved from a frequency stabilized continuous-wave (CW) laser [11,14]. Crosstalk from cloud scattering is eliminated by keeping the pulse rate below ∼8 kHz. A fast pulse rate of 8 kHz is chosen to lower the pulse energy requirement (to ∼2 mJ), to average out speckle noise and fast laser frequency noise across more pulses, and to reduce the spectral distortion from surface reflectance variations. A pulse duration of ∼1 µs is chosen to maintain a narrow laser line-width and adequate laser peak power. This approach has also been adopted to scan an O2 absorption line doublet for an atmospheric pressure measurement [15], and to measure atmospheric methane concentrations [16]. On Earth, such a column absorption line is typically broadened to a few GHz wide, mostly due to pressure broadening in the lower troposphere. Narrow line-width lasers are preferred to scan such absorption lines, to simplify the retrievals and to improve the vertical resolution of the retrieved dry mixing ratios [17].

 figure: Fig. 1.

Fig. 1. The IPDA lidar transmits a wavelength-stepped pulse train (left) to repeatedly measure CO2 absorption at multiple laser frequency channels across a single absorption line (right).

Download Full Size | PDF

The classic case of column-averaged retrieval from atmospheric absorption measured at two (online/offline) wavelengths has been analyzed extensively [5,79,1820]. Our previous study addressed multiple-layer retrievals from the absorption spectra measured at multiple pairs of symmetrical laser frequency channels [1]. Based on established inversion methods [21], closed-form formula have been derived that link both the relative random error (RRE) and the relative systematic error (RSE) of the retrieved mixing ratios, arising from the measurement noise and model bias, respectively, to characteristic parameters of the measurement and ancillary data of the atmosphere. By combining the two optical depths (ODs) measured at each pair of symmetrical channels, errors from a common laser frequency drift and spectral baseline tilt (due to, e.g., etalon fringes and surface reflectance variations) can be eliminated. The model includes laser line-shape factor and thus remains accurate even when the laser line-shape is broadened. In the present paper, we generalize this model to quantify the retrieval errors for any laser frequency channels that can be asymmetrical and are no longer combined. By solving for the frequency shift and spectral baseline tilt, atmospheric retrievals degrade only slightly even when such channels are shifted substantially out of symmetry. An optical etalon can thus be used to stabilize the reference laser wavelength.

Our forward model for arbitrary frequency channels is presented in section 2. The retrieval error analysis methods are presented in section 3 with derivation details provided in Appendix C. A numerical example of error estimation for CO2 retrievals is presented in section 4. More considerations for the retrievals and the observing system are discussed in section 5, and further details are provided in the Appendices. Throughout this paper, matrices are denoted by bold face upper case, e.g., K, column vectors by bold face lower case, e.g. b, the transpose by superscript T, e.g., KT. $\hat{{\textbf b}}$ represents an estimate of b, and $\det ({\textbf K})$ the determinant of ${\textbf K}.\;{\mathop{\textrm {Cov}}\nolimits} (x,y)$ represents the covariance between $x$ and $y,\;{\sigma ^2}(x)$ represents the variance, $\sigma (x)$ the standard deviation, $\bar{x}$ the ensemble average of x.

2. Forward model for arbitrary frequency channels

The laser frequency noise contributes to measurement errors through two complementary factors: the energy spectral density (ESD) L(${\nu_{F}}$) of each single laser pulse (as a function of the Fourier frequency ${\nu_{F}}$), and the fluctuation of the line-center frequency ${\nu_{c}}$ of L(${\nu_{F}}$). Both factors are jointly determined by the CW laser frequency noise and the laser pulse amplitude [22]. The present analysis adopts a forward model established previously for uncombined laser frequency channels [1], as summarized in this section.

In this generalized model, the m laser frequency channels are not necessarily distributed symmetrically about the center ${\nu _0}$ of the column absorption line. Let ${\nu _i} \equiv \bar{\nu }_c^i$ denote the ensemble mean of the laser line-center frequency and $\Delta {\nu _i} \equiv {\nu _i} - {\nu _0}$ the frequency offset of channel i (i = 1, 2, …, m). The atmospheric pressure p is conveniently used as the vertical coordinate and $p(r)$ denotes the atmospheric pressure at range r from the spacecraft. The column absorption line is slightly asymmetrical due to the pressure shift and ${\nu _0}$ is taken to be at the absorption peak.

The detected laser pulse energy $W_s^i$ backscattered from the surface is given by [22]

$$W_s^i = E_s^iA_s^i\exp [{ - \tau (\nu_c^i,\,2r_s^i)} ]{\kern 1pt} \,,$$
where $\tau ({\nu _c},\;2r_s^i)$ is the effective two-way OD of the target species, $r_s^i$ the range from the spacecraft to laser beam spot on the actual surface, $E_s^i$ the transmitted laser pulse energy, $A_s^i \equiv \rho T_{atm}^2D/{(r_s^i)^2}$ the lump sum of loss factors excluding $\exp [ - \tau (\nu _c^i,\,2r_s^i)],D$ an instrument parameter, ρ the surface reflectance, and Tatm the one-way transmittance of the atmosphere excluding the target species.

The effective two-way OD is related to the dry mixing ratio ${q_{gas}}(p)$ and the two-way effective weighting function $w({\nu _c},p)$ of the target species by $\tau ({\nu _c},2r)\, = \int_0^{p(r)} {{q_{gas}}(p)w({\nu _c},p)dp}$ (see Appendix A for details). We divide the atmospheric column into nq layers and attempt to solve for the layer-average ${q_j}$ of ${q_{gas}}(p)$ for each layer (j = 1, 2, …, nq, with layer 1 at the bottom). The layer-average of a dry mixing ratio profile is the ratio of the total number of target molecules to the total number of dry air molecules in a column within that layer. ${q_{gas}}(p)$ in layer j is taken to be proportional to a given dry mixing ratio profile $q_j^a(p)$ and the layer average of $q_j^a(p)$ is denoted as $< q_j^a > $ (given by Eq. (20) in Appendix A). The effective two-way OD $\tau ({\nu _c},2r)\,$ is then related to ${q_j}$ and $w({\nu _c},p)$ by $\tau ({\nu _c},2r) \cong \sum\nolimits_{j = 1}^{{n_q}} {({q_j}/ < q_j^a > )\int_{{p_j}}^{{p_{j - 1}}} {q_j^a(p)w({\nu _c},p)dp} }$ where ${p_j}$ is the pressure at the top boundary of layer j, and ${p_0} = p(r)$.

Let $r_G^i$ denote a certain mean of $r_s^i(k)$ averaged across multiple laser pulses (k = 1, 2, …, np) in channel i. ${\tau _i} \equiv \tau ({\nu _i},2r_G^i)$ can be estimated from the following transmittance ${T_i}$ averaged across the np pulses

$$\begin{array}{c} {T_i} \equiv \frac{1}{{{n_p}}}\sum\nolimits_{k = 1}^{{n_p}} {\hat{K}_s^i(k)/[\alpha \hat{E}_s^i(k)A_z^i(k)]} \,,\\ A_z^i(k) \equiv \exp [\tau ({\nu _i},2r_G^i) - \tau ({\nu _i},2r_s^i(k))] \cong \exp [ - ({q_1}/ < q_1^a > )q_1^a(p(r_G^i))\int_{p(r_G^i)}^{p(r_s^i(k))} {w({\nu _i},p)dp} ]\,. \end{array}$$
Here $\hat{E}_s^i(k)$ is the estimated energy of the transmitted laser pulse k, $\hat{K}_s^i(k)$ is the estimated photon count for the received laser pulse, $A_z^i(k)$ accounts for the OD difference $\tau ({\nu _i},2r_s^i(k)) - \tau ({\nu _i},2r_G^i)$, $\alpha \equiv \eta /(h\overline {{\nu _c}} )$ is proportional to the detector quantum efficiency η and h is the Planck constant. $\alpha \hat{E}_s^i(k)$ is the estimated photon count for the transmitted pulse. Errors due to surface height variation can be reduced to negligible levels by incorporating pulse-by-pulse ranging knowledge $r_s^i(k)$ through $A_z^i(k)$. ${\tau _i}$ is related to ${T_i}$ by $- \ln ({{{\bar{T}}_i}} )= {\tau _i} - \ln (A_{av}^i)$ where $A_{av}^i \equiv \sum\nolimits_{i = 1}^{{n_p}} {A_s^i(k)/{n_p}} .$ Here $- \ln ({{{\bar{T}}_i}} )$ can be accurately estimated by ${y_i} \equiv{-} \ln ({{T_j}} )+ C_i^{OD}$ where $C_i^{OD}$ is a bias correction factor (given by Eq. (22) in Appendix B) that becomes negligible when the integrated photon count is sufficiently high. ${y_i}$ is an accurate estimator of the OD ${\tau _i}$ plus an offset $- \ln (A_{av}^i).$ In practice, the OD baseline $- \ln (A_{av}^i)$ is often tilted and can be approximated by ${c_0} + {c_1}\Delta {\nu _i}$ (${c_0}$ and ${c_1}$ are unknown constants). All laser frequency channels could also be shifted from their nominal ensemble mean frequency ${\nu _i}$ by ${\delta _{\nu nslow}}$ plus a common laser frequency bias. Denoting the total frequency shift as ${\delta _{\nu 0}}$ and treating ${\delta _{\nu 0}}$ as another unknown constant, the measurement vector ${\textbf y} = {[{y_1},{y_2},\ldots ,{y_m}]^T}$ can be related to the state vector ${\textbf x} = {[{q_1},{q_2},\ldots ,{q_{{n_q}}},\;{\delta _{\nu 0}},\;{c_1},\;{c_0}]^T}$ by a forward model ${\textbf y} = {\textbf F}({\textbf x},{\textbf b}) + {\boldsymbol \varepsilon }$ as given by
$$\begin{array}{c} {y_i} = {[{\textbf F}({\textbf x},{\textbf b})]_i} + {\varepsilon _i} = \,\sum\nolimits_{j = 1}^{{n_q}} {{{[{{\textbf K}_q}]}_{i,j}}{q_j}} + {c_0} + {c_1}\Delta {\nu _i} + {\varepsilon _i},\\ {[{{\textbf K}_q}]_{i,j}} \equiv {{\int_{{p_j}}^{{p_{j - 1}}} {q_j^a(p)w({\nu _i} + {\delta _{\nu 0}},p)dp} } \mathord{\left/ {\vphantom {{\int_{{p_j}}^{{p_{j - 1}}} {q_j^a(p)w({\nu_i} + {\delta_{\nu 0}},p)dp} } { < q_j^a > }}} \right.} { < q_j^a > }}\,\quad (j \le {n_q}), \end{array}$$
where b is a vector of model parameters that are not to be retrieved (including temperature and water vapor profiles, surface pressure, and spectroscopic data), ${p_0} = p(r_G^i)$, and ${\varepsilon _i}$ represents a measurement error of ${y_i}$. ${y_i}$ depends on ${\delta _{\nu 0}}$ nonlinearly through ${[{{\textbf K}_q}]_{i,j}}.$ Without other prior constraint, the number of unknowns in ${\textbf x}$ is at most equal to m. For column-averaged retrieval of ${\textbf x} = {[q,\;{\delta _{\nu 0}},\;{c_1},\;{c_0}]^T},$ at least four wavelength channels are required. For the narrow line-width case, $\tau ({\nu _c},2r)$ becomes twice the one-way monochromatic OD ${\tau _0}({\nu _F},p(r))$ and $w({\nu _c},p)$ twice the one-way monochromatic weighting function ${w_0}({\nu _F},p)$.

As needed for the retrievals, the covariance matrix ${{\textbf S}_y}$ of ${\textbf y}$ and the model bias ${\bar{\varepsilon }_i}$ are quantified in [1] and summarized in Appendix B where the OD slope at channel i is denoted as ${\dot{\tau }_i} \equiv \partial {\tau _i}/\partial {\nu _i}$ and $\dot{{\boldsymbol \tau }} \equiv {[{\dot{\tau }_1},{\dot{\tau }_2},\ldots ,{\dot{\tau }_m}]^T}.$ It would be ideal if the slow drifts of all laser frequency channels are the same so the common frequency drift can be solved and corrected for. This can be achieved by phase locking the transmitted laser frequencies to a master laser [11,14]. For this ideal case, ${{\textbf S}_y}$ can be expressed as

$$\begin{array}{c} {{\textbf S}_y} = {{\textbf S}_{y0}} + {\sigma ^2}({{\delta_{\nu nslow}}} )\dot{{\boldsymbol \tau }}{{\dot{{\boldsymbol \tau }}}^T},\\ {[{{\textbf S}_{y0}}]_{i,j}} = \sigma _u^2({{y_i}} ){\delta _{i,j}},\\ \sigma _u^2({{y_i}} )\equiv {\sigma ^2}({{y_i}} )- {{\dot{\tau }}_i}^2{\sigma ^2}({{\delta_{\nu nslow}}} ). \end{array}$$
Here ${\delta _{\nu nslow}}$ is an average of the slow drift ${\delta _{\nu slow}}$ of the laser line-center frequency ${\nu _c}$ across np pulses in a channel (defined by Eq. (23) in Appendix B), and the element $\sigma _u^2({{y_i}} )$ of the diagonal ${{\textbf S}_{y0}}$ is the variance of ${y_i}$ excluding the contribution ${\dot{\tau }_i}^2{\sigma ^2}({{\delta_{\nu nslow}}} )$ from ${\delta _{\nu nslow}}.$ ${\delta _{\nu nslow}}$ remains constant and essentially the same for all channels within each pulse averaging time, but varies slowly over longer time scales. Within each pulse averaging time, all channels are shifted by ${\delta _{\nu nslow}}$ plus a common frequency bias (e.g., the Doppler shift arising from the high-speed cross-wind and the radial component of the spacecraft velocity [18]). ${\sigma ^2}({{\delta_{\nu nslow}}} )$ is only slightly smaller than ${\sigma ^2}({{\delta_{\nu slow}}} ).$ The nonzero off-diagonal elements of ${{\textbf S}_y}$ contained in ${\sigma ^2}({{\delta_{\nu nslow}}} )\dot{{\boldsymbol \tau }}{\dot{{\boldsymbol \tau }}^T}$ provide more information about ${\textbf x}$ (hence less retrieval errors) and it is important not to ignore them. For the ideal case, the retrieval errors from ${\delta _{\nu slow}}$ can be essentially eliminated if the frequency channels are placed symmetrically.

On the other extreme, the transmitted laser frequencies are uncorrelated (with uncorrelated slow drifts) among different channels so that ${{\textbf S}_y}$ becomes diagonal. The diagonal element ${\sigma ^2}({{y_i}} )$ includes the contribution from the slow frequency drift (in channel i) that essentially does not decrease from pulse averaging. As illustrated in section 4, the uncorrelated laser frequency drifts can drive up retrieval errors significantly.

3. Retrieval error analysis for arbitrary frequency channels

The matrix formulas for the retrievals have been described in [1], and are briefly summarized herein. To simplify the analysis, we assume no a priori information and consider the maximum likelihood (ML) retrieval approach. Since the forward model is in general nonlinear, an estimate $\hat{{\textbf x}}$ of ${\textbf x}$ can only be retrieved iteratively. Nevertheless, it is adequate to linearize the forward model ${\textbf F}({\textbf x})$ about an estimated state ${{\textbf x}_i}$ for the retrieval error analysis. Using the Newton-Gauss method, the following iteration ${{\textbf x}_{i + 1}}$ converges to $\hat{{\textbf x}}$ [21]

$$\begin{array}{c} {{\textbf x}_{i + 1}} = {{\textbf x}_i} + {({\textbf K}_i^T{\textbf S}_y^{ - 1}{{\textbf K}_i})^{ - 1}}\{ {\textbf K}_i^T{\textbf S}_y^{ - 1}[{\textbf y}({{\textbf x}_i}) - {\textbf F}({{\textbf x}_i})]\} ,\\ {[{\textbf K}({\textbf x})]_{i,j}} \equiv \frac{{\partial {{[{\textbf F}({\textbf x}) - {\textbf y}]}_i}}}{{\partial {x_j}}} \cong \frac{{\partial {{[{\textbf F}({\textbf x})]}_i}}}{{\partial {x_j}}}, \end{array}$$
where ${\textbf K}({\textbf x})$ the effective Jacobian matrix, ${{\textbf K}_i} = {\textbf K}({{\textbf x}_i}),$ and the weak dependence of ${\textbf y}$ on q1 (through $A_z^i(k)$) can be neglected. It should be noted that all unknowns in ${\textbf x}$ are simultaneously updated in the retrieval iterations. Assuming the narrow laser line-width case, ${{\textbf K}_q}$ becomes independent of the mixing ratios ${q_j}.$ The elements of ${\textbf K}({\textbf x})$ are found to be ${[{\textbf K}]_{i,j}} = {[{{\textbf K}_q}]_{i,j}}\;(j \le {n_q}),$ ${[{\textbf K}]_{i,{n_q} + 1}}\; = {\dot{\tau }_i},$ ${[{\textbf K}]_{i,{n_q} + 2}}\; = \Delta {\nu _i},$ and ${[{\textbf K}]_{i,{n_q} + 3}}\; = 1.$ Here ${x_{{n_q} + 1}} = {\delta _{\nu 0}},{x_{{n_q} + 2}} = {c_1},$ and ${x_{{n_q} + 3}} = {c_0}.$ When ${\delta _{\nu 0}}$ is not included in ${\textbf x}$ and the laser line-width is narrow, the forward model becomes essentially linear. This allows the following linear retrieval ${\hat{{\textbf x}}_L}$ of ${\textbf x}$ without iterations [21]
$${\hat{{\textbf x}}_L} = {({{\textbf K}^T}{\textbf S}_y^{ - 1}{\textbf K})^{ - 1}}{{\textbf K}^T}{\textbf S}_y^{ - 1}{\textbf y}\,.$$
This linear ML solution is equivalent to generalized linear least-square solution with weighting covariance matrix ${\textbf S}_y^{ - 1}.$ It can serve as the start point ${{\textbf x}_0}$ for the iterations. Such a linear fitting is illustrated in Fig. 2 (left) for column-averaged retrieval of ${\textbf x} = {[q,\;{c_0}]^T}$ from ${y_i}$ plotted as a function of ${[{{\textbf K}_q}]_{i,1}}$ across the atmospheric CO2 line at 1572.335 nm. The CO2 absorption spectrum (Fig. 2, right) is computed for US standard atmospheric conditions with a constant dry CO2 mixing ratio of 400 ppm. 8 channels (marked with blue dots) are placed symmetrically at $\Delta {\nu _8} ={-} \Delta {\nu _1} = 15.6\;\textrm{GHz,}$ $\Delta {\nu _7} ={-} \Delta {\nu _2} = 1.7\;\textrm{GHz,}\Delta {\nu _6} ={-} \Delta {\nu _3} = 1.08\;\textrm{GHz,}$ and $\Delta {\nu _5} ={-} \Delta {\nu _4} = 0.5\;\textrm{GHz}\textrm{.}$ The red squares mark the channels when they are shifted by 100 MHz (for illustration purposes). $\hat{q}$ is the slope of the fitting line and ${\hat{c}_0}$ the vertical intercept. The fitting line can be regarded as a rigid bar constrained by data points $({y_i},\;{[{{\textbf K}_q}]_{i,1}})$ with error bars described by ${{\textbf S}_y}.$ Reducing the retrieval errors can be visualized as constraining the rigid bar from rotating. By placing the channels symmetrically, a small common frequency shift causes nearly opposite OD changes in each pair of symmetrical channels without rotating the rigid bar. Similarly, a small OD baseline tilt ${c_1}\Delta {\nu _i}$ will result in opposite OD changes in each pair of symmetrical channels again without rotating the rigid bar. When the channels are shifted substantially out of symmetry, the rigid bar tends to rotate with the common frequency shift or OD baseline tilt. To correct for the shift, non-linear iterations of Eq. (5) can be invoked to solve for ${\delta _{\nu 0}}$ in addition to $q,$ ${c_0},$ and ${c_1}$.

 figure: Fig. 2.

Fig. 2. Linear least-square fitting (left) for retrieving q and c0 from monochromatic OD measurement of atmospheric CO2 line at 1572.335 nm (right). The blue dots mark four pair of symmetrical channels at their ensemble mean frequencies. The red squares mark the channels when they are shifted by 100 MHz.

Download Full Size | PDF

The estimated error covariance matrix $\hat{{\textbf S}}$ of $\hat{{\textbf x}}$ is found to be [21]

$$\hat{{\textbf S}} = {({{\textbf K}^T}{\textbf S}_y^{ - 1}{\textbf K})^{ - 1}}.$$
The retrieval bias towards ${\textbf x}$ can be expressed by [21]
$$\hat{{\textbf x}} - {\textbf x} = \hat{{\textbf S}}{{\textbf K}^T}{\textbf S}_y^{ - 1}[{\bar{{\boldsymbol \varepsilon }} + ({{\textbf F}(\hat{{\textbf x}},{\textbf b}) - {\textbf F}(\hat{{\textbf x}},\hat{{\textbf b}})} )} ].$$
The first term in Eq. (8) is due to the forward model bias $\bar{{\boldsymbol \varepsilon }}$ and the second term due to the errors in the forward model parameters ${\textbf b} - \hat{{\textbf b}}.$ In this paper, we focus on the first term. The retrieval bias due to imperfect forward model parameters is equal to that from a model bias $\bar{{\boldsymbol \varepsilon }} = {\textbf F}(\hat{{\textbf x}},{\textbf b}) - {\textbf F}(\hat{{\textbf x}},\hat{{\textbf b}}),$ and can be evaluated with the formula derived for $\bar{{\boldsymbol \varepsilon }}$ here.

The RRE and RSE of ${\hat{q}_j}$ for multiple-layer retrievals are derived from Eqs. (7) and (8) in Appendix C. Here we examine the simple case of column-averaged retrieval. In these derivations, $< {a_i} > \, \equiv {{\sum\nolimits_{i = 1}^m {{a_i}\sigma _u^{ - 2}({{y_i}} )} } \mathord{\left/ {\vphantom {{\sum\nolimits_{i = 1}^m {{a_i}\sigma_u^{ - 2}({{y_i}} )} } {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({{y_i}} )} }}} \right.} {\sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({{y_i}} )} }}$ denotes an average of the m elements of a vector ${\textbf a}$ weighted by $\sigma _u^{ - 2}({{y_i}} ),$ ${{\mathop{\textrm {Var}}\nolimits} _i}({a_i}) \equiv \, < {({a_i} - < {a_i} > )^2} > $ a variance of ${a_i},$ and ${{\mathop{\textrm {Cov}}\nolimits} _i}({a_i},{b_i}) \equiv \, < ({a_i} - < {a_i} > )({b_i} - < {b_i} > ) > $ a covariance between elements ${a_i}$ and ${b_i}$ across m channels. Also, ${\tau _{ij}} \equiv {[{{\textbf K}_q}]_{i,j}}{\hat{q}_j}$ is the estimated two-way OD of the target species within layer j at channel i, $\Delta {\tau _j} \equiv {[{{\mathop{\textrm {Var}}\nolimits} _i}({\tau _{ij}})]^{1/2}}$ is an effective two-way differential absorption OD (DAOD) within layer j that quantifies the spread of the ODs ${\tau _{ij}}$ across m channels, $\Delta \tau \equiv {[{{\mathop{\textrm {Var}}\nolimits} _i}({\tau _i})]^{1/2}}$ is an effective two-way DAOD for the entire column, and a constant ${c_{\Delta \tau }}$ is defined as $c_{\Delta \tau }^2 \equiv {\sigma ^2}({{\delta_{\nu nslow}}} ){{\mathop{\textrm {Var}}\nolimits} _i}({\dot{\tau }_i})[\sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({y_i})]} .$ We also define the following correlation coefficients: ${r_{\dot{\tau }}} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({\tau _i},{\dot{\tau }_i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({\tau _i}){{\mathop{\textrm {Var}}\nolimits} _i}({\dot{\tau }_i})]^{1/2}}$, ${r_{\Delta \nu }} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({\tau _i},\Delta {\nu _i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({\tau _i}){{\mathop{\textrm {Var}}\nolimits} _i}(\Delta {\nu _i})]^{1/2}}$, ${r_\varepsilon } \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({\tau _i},{\bar{\varepsilon }_i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({\tau _i}){{\mathop{\textrm {Var}}\nolimits} _i}({\bar{\varepsilon }_i})]^{1/2}}$, ${r_{\dot{\tau }\Delta \nu }} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({\dot{\tau }_i},\Delta {\nu _i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({\dot{\tau }_i}){{\mathop{\textrm {Var}}\nolimits} _i}(\Delta {\nu _i})]^{1/2}}$, ${r_{\dot{\tau }\varepsilon }} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({\dot{\tau }_i},{\bar{\varepsilon }_i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({\dot{\tau }_i}){{\mathop{\textrm {Var}}\nolimits} _i}({\bar{\varepsilon }_i})]^{1/2}}.$ They all vary between -1 and 1.

From Eq. (37) in Appendix C, the RRE of $\hat{q}$ for ${\textbf x} = {[q,\;{c_0}]^T}$ is found to be

$$\begin{array}{c} \frac{{\sigma (\hat{q})}}{{\hat{q}}} = \frac{{\sigma (\Delta \tau )}}{{\Delta \tau }},\\ {\sigma ^2}(\Delta \tau ) = {\left[ {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right]^{ - 1}} + \frac{{{\sigma ^2}({{\delta_{\nu nslow}}} ){{{\mathop{\textrm {Var}}\nolimits} }_i}({{\dot{\tau }}_i})r_{\dot{\tau }}^2}}{{1 + c_{\Delta \tau }^2(1 - r_{\dot{\tau }}^2)}}, \end{array}$$
where ${\sigma ^2}(\Delta \tau )$ can be regarded as an effective measurement variance of the effective DAOD $\Delta \tau .$ When $m = 2,$ the fitting line passes through both data points, and ${r_{\dot{\tau }}} = 1.$ For this two-wavelength case, Eq. (9) is reduced to the classic formula $\sigma (\hat{q})/\hat{q} = \sigma ({y_2} - {y_1})/|{y_2} - {y_1}|,$ and $\Delta \tau = |{y_2} - {y_1}|{\sigma _u}({y_1}){\sigma _u}({y_2})/[\sigma _u^2({y_1}) + \sigma _u^2({y_2})]$ is maximized to $|{y_2} - {y_1}|/2$ when $\sigma _u^2({y_1}) = \sigma _u^2({y_2}).$ Eq. (9) indicates that the RRE of $\hat{q}$ is equal to the RRE of $\Delta \tau .$ The first and second terms of ${\sigma ^2}(\Delta \tau )$ come from ${{\textbf S}_{y0}}$ and ${\sigma ^2}({{\delta_{\nu nslow}}} )\dot{{\boldsymbol \tau }}{\dot{{\boldsymbol \tau }}^T},$ respectively. The second term is non-negative and becomes negligible for symmetrical channels whose ${r_{\dot{\tau }}}^2$ approximates zero. The first term ${[\sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({y_i})} ]^{ - 1}}$ combines $\sigma _u^2({y_i})$ in parallel, just like combining m resistors in parallel. From Eq. (39) in Appendix C, the RSE of $\hat{q}$ for ${\textbf x} = {[q,\;{c_0}]^T}$ arising from the model bias ${\bar{\varepsilon }_i}$ is
$$\begin{array}{c} \frac{{{\delta _{\hat{q}}}}}{{\hat{q}}} = \frac{{{\delta _{\Delta \tau }}}}{{\Delta \tau }},\\ {\delta _{\Delta \tau }} = {r_\varepsilon }{[{{\mathop{\textrm {Var}}\nolimits} _i}({{\bar{\varepsilon }}_i})]^{1/2}} + \frac{{c_{\Delta \tau }^2{{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\bar{\varepsilon }}_i})]}^{1/2}}({{r_{\dot{\tau }}}{r_\varepsilon } - {r_{\dot{\tau }\varepsilon }}} ){r_{\dot{\tau }}}}}{{1 + c_{\Delta \tau }^2(1 - r_{\dot{\tau }}^2)}}, \end{array}$$
where ${\delta _{\Delta \tau }}$ can be regarded as an effective measurement bias of $\Delta \tau .$ The two terms of ${\delta _{\Delta \tau }}$ also come from ${{\textbf S}_{y0}}$ and $\dot{{\boldsymbol \tau }}{\dot{{\boldsymbol \tau }}^T}{\sigma ^2}({{\delta_{\nu nslow}}} ),$ respectively. The second term of ${\delta _{\Delta \tau }}$ also becomes negligible when ${r_{\dot{\tau }}} \cong 0.$ For $m = 2,$ we have ${r_{\dot{\tau }}} = {r_\varepsilon } = {r_{\dot{\tau }\varepsilon }} = 1$ and ${\delta _{\hat{q}}}/\hat{q} = ({\bar{\varepsilon }_2} - {\bar{\varepsilon }_1})/({y_2} - {y_1})$.

When $m \ge 3,{\delta _{\nu 0}}$ can be included in ${\textbf x}$ (i.e., ${\textbf x} = {[q,{\delta _{\nu 0}},\;{c_0}]^T}$). The RRE and RSE of $\hat{q}$ and ${\sigma ^2}({\delta _{\nu 0}})$ are found to be

$$\frac{{\sigma (\hat{q})}}{{\hat{q}}} = \frac{{{{\left[ {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right]}^{ - 1/2}}}}{{{{(1 - r_{\dot{\tau }}^2)}^{1/2}}\Delta \tau }},$$
$$\frac{{{\delta _{\hat{q}}}}}{{\hat{q}}} = \frac{{({r_\varepsilon } - {r_{\dot{\tau }\varepsilon }}{r_{\dot{\tau }}}){{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\bar{\varepsilon }}_i})]}^{1/2}}}}{{(1 - r_{\dot{\tau }}^2)\Delta \tau }},$$
$${\sigma ^2}({\hat{\delta }_{\nu 0}}) = \frac{{{{\left[ {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right]}^{ - 1}}}}{{{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\dot{\tau }}_i})(1 - r_{\dot{\tau }}^2)}} + {\sigma ^2}({\delta _{\nu nslow}}).$$
The second terms vanish in Eqs. (11) and (12) as if the slow frequency drift ${\delta _{\nu nslow}}$ were removed. Since ${\delta _{\nu 0}}$ is the sum of the frequency bias and ${\delta _{\nu nslow}},$ the first term of ${\sigma ^2}({\hat{\delta }_{\nu 0}})$ in Eq. (13) is the variance for the frequency bias estimation and the second term is the variance for ${\delta _{\nu nslow}}.$ As to be illustrated in the next section, the RRE and RSE of $\hat{q}$ and the first term of ${\sigma ^2}({\hat{\delta }_{\nu 0}})$ increase only slightly when the channels are shifted out of symmetry by ${\delta _{\nu 0}}$ within ± 1 GHz.

Similarly, the RRE and RSE for ${\textbf x} = {[q,{\delta _{\nu 0}},\;{c_1},{c_0}]^T}$ can also be derived when $m \ge 4.$ When the laser frequency channels are nearly symmetrically distributed so that ${r_{\dot{\tau }}} \cong 0$ and ${r_{\Delta \nu }} \cong 0,$ the RREs and RSEs of $\hat{q}$ for ${\textbf x} = {[q,{\delta _{\nu 0}},\;{c_1},{c_0}]^T},$ ${\textbf x} = {[q,{\delta _{\nu 0}},\;{c_0}]^T},$ and ${\textbf x} = {[q,{c_1},{c_0}]^T}$ are also minimized and become essentially the same as those for ${\textbf x} = {[q,\;{c_0}]^T}$ (${[\sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({y_i})} ]^{ - 1/2}}/\Delta \tau$ and ${r_\varepsilon }{[{{\mathop{\textrm {Var}}\nolimits} _i}({\bar{\varepsilon }_i})]^{1/2}}/\Delta \tau ,$ respectively). This means that ${\delta _{\nu 0}}$ and ${c_1}$ can be solved and corrected for without degrading the RRE and RSE of $\hat{q}$ when the frequency channels are nearly symmetrical. When the channels are symmetrical, the first term of ${\sigma ^2}({\delta _{\nu 0}})$ for ${\textbf x} = {[q,{\delta _{\nu 0}},\;{c_1},{c_0}]^T}$ equals to that for ${\textbf x} = {[q,{\delta _{\nu 0}},\;{c_0}]^T}$ multiplied $1/(1 - r_{\dot{\tau }\Delta \nu }^2).$ For the symmetrical channels illustrated in Fig. 2, $r_{\dot{\tau }\Delta \nu }^2$ approximates zero (as shown in section 4) so that fitting for OD baseline tilt does not degrade fitting for ${\delta _{\nu 0}}$ either. If the frequency channels are symmetrically combined, the information for ${\delta _{\nu 0}}$ and ${c_1}$ is lost, but the information for q and ${c_0}$ is preserved. As shown in Eqs. (11) and (13), solving for ${\delta _{\nu 0}}$ with asymmetrical frequency channels will increase the RRE of $\hat{q}$ by $1/{(1 - r_{\dot{\tau }}^2)^{1/2}}$ and the first term of ${\sigma ^2}({\delta _{\nu 0}})$ by $1/(1 - r_{\dot{\tau }}^2).$ As shown in section 4, these increases are negligible even when the symmetrical channels are shifted out of symmetry by ${\delta _{\nu 0}}$ within ± 1 GHz.

The noise contributions to ${\sigma ^2}(\Delta \tau )$ from sources other than ${\delta _{\nu slow}}$ (such as the signal shot noise, speckle noise, and fast frequency noise) are uncorrelated among different pulses and can be reduced by pulse averaging. The shot-noise limited $\sigma (\Delta \tau ) = {[\sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({y_i})} ]^{ - 1/2}}$ (and hence the retrieval RRE) is inversely proportional to the square root of the accumulated photon count of the pulses from all m channels.

To restrain the rotation of the rigid bar (hence the RRE and RSE of $\hat{q}$) as illustrated in Fig. 2, it is desirable to “clamp” the bar at its two ends and keep the two ends far apart (i.e., to increase $\Delta \tau$). To reduce the rotational “play” of the rigid bar (hence the RRE of $\hat{q}$), the error bars of ${y_i}$ need to be reduced, particularly at the two extreme ends. The online and offline channels are equivalent in their contributions to $\Delta \tau$ and $\sigma (\Delta \tau ).$ $\Delta \tau$ can be increased by minimizing the absorption of offline channels and shifting online channels towards the absorption peak. Minimizing the absorption of offline channels also reduces their error bars (and hence $\sigma (\Delta \tau )$). Increasing the absorption of online channels increases their error bars (and hence $\sigma (\Delta \tau )$) and could end up reducing $\sigma (\Delta \tau )/\Delta \tau$ and even $\Delta \tau .$ It is often desirable to place some online points on the sides of the absorption line to uniformly sense concentrations in the lower troposphere. This may decrease $\Delta \tau$ and increase the RRE and RSE of $\hat{q}.$ Further error reduction considerations will be discussed along with the numerical example in the next section.

4. Numerical estimation of retrieval errors

We now estimate the errors of atmospheric CO2 retrievals from the lidar absorption spectra measured with a narrow line-width laser across the 1572.335 nm CO2 absorption line. The present calculations are based on the same surface reflectance data [23] and essentially the same realistic parameters (listed in Table 1) employed in [1]. The pulses of all channels are assumed to have the same transmitted pulse energy ${E_s}$ of 2 mJ, smaller than the 4mJ assumed in [1]. The detector specifications listed in Table 1 are rather conservative compared to the nearly noiseless HgCdTe APD detectors employed in NASA's IPDA lidars [12]. The additional model bias and measurement noise due to imperfect ranging are negligible for a ranging bias ≤ 0.66 m and a ranging precision ≤ 20 m [22], and are neglected hereafter. The speckle noise is neglected due to the large telescope diameter of 1.5 m and large laser beam spot size of 50 m on the surface [5,24,25]. Unless noted otherwise, the results illustrated in the following figures are for an IPDA lidar with 8 symmetrical channels (as shown in Fig. 2) having a common slow drift ${\delta _{\nu nslow}}$ with $\sigma ({\delta _{\nu nslow}}) = 3\;\textrm{MHz}$.

Tables Icon

Table 1. Parameters used for numerical estimation of retrieval errors

 figure: Fig. 3.

Fig. 3. (left) The measurement noise σ(yi) for atmospheric CO2 (solid black) as a function of 2τ0 (for frequency offset < 0), computed using parameters listed in Table 1. The blue dots mark the channels ν1 to ν4. Also plotted are partial contributions to σ(yi) from the signal shot noise (solid grey), slow laser line-center frequency drift (dashed red), solar background (dotted brown), receiver circuitry noise (dash-dotted green), and detector dark count (long-dashed blue). (right) The RRE of $\hat{q}$ (solid black), ${r_{\dot{\tau }}}$ (dotted red), ${r_{\Delta \nu }}$ (dash-dotted green), and ${r_{\dot{\tau }\Delta \nu }}$ (dashed blue) as functions of a common frequency shift ${\delta _{\nu 0}}.$ The RRE of $\hat{q}$ becomes much larger (solid grey) when the transmitted laser frequencies become uncorrelated.

Download Full Size | PDF

The measurement noise $\sigma ({y_i})$ and ${\sigma _u}({y_i})$ are computed from Eq. (24) in Appendix B and plotted as functions of two-way OD $2{\tau _0}$ in Fig. 3 (left). The same figure also shows partial contributions to $\sigma ({y_i})$ from the signal shot noise, slow laser line-center frequency drift ${\delta _{\nu slow}},$ solar background, receiver circuitry noise, and detector dark count. The effective fast laser line-center frequency noise ${\sigma ^2}({\delta _{\nu nfast}})$ is averaged down to 28 kHz and its contribution to $\sigma ({y_i})$ is negligible. On the other hand, the noise contribution from ${\delta _{\nu slow}}$ becomes dominant for online channels. The RRE of $\hat{q}$ is computed from Eq. (11) and plotted in Fig. 3 (right) as a function of a common frequency shift ${\delta _{\nu 0}}$ added to the 8 nominal channels illustrated in Fig. 2. Computing from Eq. (9) yields essentially the same RRE. The correlation coefficients ${r_{\dot{\tau }}},$ ${r_{\Delta \nu }},$ and ${r_{\dot{\tau }\Delta \nu }}$ shown in Fig. 3 (right) are ≤ 0.2 so that their squares approximate zero. Even when the channels are shifted out of symmetry by ± 1 GHz, The RRE of $\hat{q}$ does not increase significantly from its minimum of 0.052% at ${\delta _{\nu 0}} = 0.$ Also the square root of the first term of ${\sigma ^2}({\delta _{\nu 0}})$ in Eq. (13) (i.e., the standard deviation of the frequency bias estimation, not shown in Fig. 3) is ≤ 0.53 MHz when ${\delta _{\nu 0}}$ varies within ± 1 GHz and is minimized to 0.41 MHz at ${\delta _{\nu 0}} = 0.$ When the laser channel frequencies are uncorrelated (while all other conditions being the same), the RRE of $\hat{q}$ (shown in Fig. 3 (right)) increases significantly to 0.126% even at ${\delta _{\nu 0}} = 0$.

Figure 4 (left) shows the RREs of $\hat{q}$ for IPDA lidars using two channels (solid grey) and four symmetrical channels (solid black) as functions of $|\Delta {\nu _2}|,$ the absolute value of the online frequency offset. For two channels, the offline frequency offset is fixed at $\Delta {\nu _1} ={-} 15.6\;\textrm{GHz}$ while $\Delta {\nu _2}$ varies from 0 to -2 GHz. For four channels, $- \Delta {\nu _1} = \Delta {\nu _4} = 15.6\;\textrm{GHz}$ and $- \Delta {\nu _2} = \Delta {\nu _3}$. For fair comparisons, the pulse rate for each channel is 4 kHz for two channels and 2 kHz for four channels (while all other parameters being the same). The RRE of $\hat{q}$ jumps to ≥ 0.19% for the traditional two-wavelength lidar. In comparison, adding two more channels symmetrically reduces the RRE to a minimum of 0.048% when $- \Delta {\nu _2} = \Delta {\nu _3} = 0.78\;\textrm{GHz}\textrm{.}$ Also plotted are the partial RSEs of $\hat{q}$ for the two-channel lidar due to a frequency bias of 3 MHz (dashed red) and a small OD slope of $3.3 \times {10^{ - 4}}/\textrm{GHz}$ (dotted green). When the channels are nearly symmetrical (such as the four symmetrical channels), such partial retrieval biases from the frequency shift and OD baseline tilt are essentially eliminated. Figure 4 (right) shows the RSEs of $\hat{q}$ for 1-s averaging time calculated from the surface reflectance data as functions of the starting position along the path for the 8 symmetrical channels (grey), 4 symmetrical channels (blue), and 2-channels (red). The peak RSE of $\hat{q}$ does not change significantly when the 8 (or four) channels are shifted out of symmetry within ± 1 GHz. To simplify the calculation, each beam spot is in effect taken to be a uniformly illuminated square (aligned to the track), and variations of surface reflectance perpendicular to the track are neglected. The computed peak RSE arising from surface reflectance variations is the smallest (∼8×10−6) for four-channels, larger (∼2.4×10−5) for 8 channels, and largest for two-channels (∼1.1×10−4). In reality, the RSEs will be somewhat higher due to non-uniform illumination of the beam spots and cross-track variations of surface reflectance.

 figure: Fig. 4.

Fig. 4. (left) The RREs of $\hat{q}$ for two-channel (solid grey) and four-channel (solid black) lidars as functions of the absolute value of the online frequency offset. Δν1 is fixed at -15.6 GHz for two channels. For four channels, -Δν1 = Δν4 = 15.6 GHz and -Δν2 = Δν3. Also plotted for two channels are the partial RSEs of $\hat{q}$ due to a frequency bias of 3 MHz (dashed red) and a small OD slope of 3.3×10−4/GHz (dotted green). (right) RSEs of $\hat{q}$ for 1-s averaging time calculated from the surface reflectance data as functions of the starting position along the path for the 8 symmetrical channels (grey), four symmetrical channels (blue), and two-channels (red).

Download Full Size | PDF

5. Discussions

5.1. Retrievals from transmittance measurement

It should be noted that the same state vector ${\textbf x} = {[{q_1},{q_2},\ldots ,{q_{{n_q}}},\;{\delta _{\nu 0}},\;{c_1},\;{c_0}]^T}$ can also be retrieved from the transmittance measurement vector ${\textbf t} = {[{T_1},{T_2},\ldots ,{T_m}]^T}$ by using the following forward model

$${T_i} = {[{{\textbf F}_t}({\textbf x},{\textbf b})]_i} = \exp \,\left( { - \left[ {\sum\nolimits_{j = 1}^{{n_q}} {{{[{{\textbf K}_q}]}_{i,j}}{q_j}} + {c_0} + {c_1}\Delta {\nu_i}} \right]} \right) + \varepsilon _i^t.$$
Here the transmittance measurement error $\varepsilon _i^t$ is related to the OD error ${\varepsilon _i}$ by ${\varepsilon _i} ={-} \ln (1 + \varepsilon _i^t/{\bar{T}_i}).$ The OD measurement noise ${\varepsilon _i} - {\bar{\varepsilon }_i}$ can then be expressed as
$${\varepsilon _i} - {\bar{\varepsilon }_i} \cong{-} \frac{{(\varepsilon _i^t - \bar{\varepsilon }_i^t)}}{{{{\bar{T}}_i}}}\left( {1 - \frac{{\bar{\varepsilon }_i^t}}{{{{\bar{T}}_i}}}} \right) + \frac{{{{(\varepsilon _i^t - \bar{\varepsilon }_i^t)}^2} - \overline {{{(\varepsilon _i^t - \bar{\varepsilon }_i^t)}^2}} }}{{2\bar{T}_i^2}} \cong{-} \frac{{(\varepsilon _i^t - \bar{\varepsilon }_i^t)}}{{{{\bar{T}}_i}}}\left( {1 - \frac{{\bar{\varepsilon }_i^t}}{{{{\bar{T}}_i}}}} \right).$$
Here $[{(\varepsilon _i^t - \bar{\varepsilon }_i^t)^2} - \overline {{{(\varepsilon _i^t - \bar{\varepsilon }_i^t)}^2}} ]/(2\bar{T}_i^2)$ is neglected so that $\; {\varepsilon _i} - {\bar{\varepsilon }_i}$ becomes proportional to $(\varepsilon _i^t - \bar{\varepsilon }_i^t)$ and thus has the same type of probability distribution as $(\varepsilon _i^t - \bar{\varepsilon }_i^t).$ Since $\bar{\varepsilon }_i^t/{\bar{T}_i} \ll 1,$ the covariance ${{\textbf S}_t}$ of $\varepsilon _i^t - \bar{\varepsilon }_i^t$ becomes ${[{{\textbf S}_t}]_{i,j}} = {T_i}{T_j}{[{{\textbf S}_y}]_{i,j}}$ and the corresponding effective Jacobian matrix ${{\textbf K}_t}({\textbf x})$ becomes ${[{{\textbf K}_t}({\textbf x})]_{i,j}} = {T_i}{[{\textbf K}({\textbf x})]_{i,j}}.$ Equivalently, we have ${{\textbf S}_t} = {{\textbf I}_t}{{\textbf S}_y}{{\textbf I}_t}$ and ${{\textbf K}_t}({\textbf x}) = {{\textbf I}_t}{\textbf K}({\textbf x}),$ where ${{\textbf I}_t}$ is diagonal with ${[{{\textbf I}_t}]_{i,i}} = {\bar{T}_i}.$ The corresponding covariance matrix $\hat{{\textbf S}}$ of ${\textbf x}$ is $\hat{{\textbf S}} = {\textbf K}_t^T{\textbf S}_t^{ - 1}{{\textbf K}_t} = ({{\textbf K}^T}{{\textbf I}_t})({\textbf I}_t^{ - 1}{\textbf S}_y^{ - 1}{\textbf I}_t^{ - 1})({{\textbf I}_t}{\textbf K}) = {{\textbf K}^T}{\textbf S}_y^{ - 1}{\textbf K},$ the same as that for OD measurement as given by Eq. (7). Similarly, the retrieval bias from $\bar{\varepsilon }_i^t$ and ${\textbf b} - \hat{{\textbf b}}$ is $\delta {\textbf x} = {\hat{\textbf{S}}K}_t^T{\textbf S}_t^{ - 1}[{{{\bar{{\boldsymbol \varepsilon }}}^t} + ({{{\textbf F}_t}(\hat{{\textbf x}},{\textbf b}) - {{\textbf F}_t}(\hat{{\textbf x}},\hat{{\textbf b}})} )} ]= \hat{{\textbf S}}({{\textbf K}^T}{{\textbf I}_t})({\textbf I}_t^{ - 1}{\textbf S}_y^{ - 1}{\textbf I}_t^{ - 1})[{{{\textbf I}_t}\bar{{\boldsymbol \varepsilon }} + {{\textbf I}_t}({{\textbf F}(\hat{{\textbf x}},{\textbf b}) - {\textbf F}(\hat{{\textbf x}},\hat{{\textbf b}})} )} ],$ the same as given by Eq. (8) for OD measurement.

When the measurement noise is small so that Eq. (15) is valid, it is essentially equivalent to retrieve ${\textbf x}$ from either the transmittance measurement or the OD measurement. The same error formula derived for retrievals from OD measurement (i.e., Eqs. (9)–(13), (33), (36)–(40)) can also be derived from $\hat{{\textbf S}} = {\textbf K}_t^T{\textbf S}_t^{ - 1}{{\textbf K}_t}$ and $\delta {\textbf x} = {\hat{\textbf{S}}K}_t^T{\textbf S}_t^{ - 1}{\bar{{\boldsymbol \varepsilon }}^t}$ for the transmittance measurement (with proper substitutions such as ${[{{\textbf K}_t}({\textbf x})]_{i,j}} = {T_i}{[{\textbf K}({\textbf x})]_{i,j}}$ and ${[{{\textbf S}_t}]_{i,j}} = {T_i}{T_j}{[{{\textbf S}_y}]_{i,j}}$). It is more efficient to retrieve from OD measurement because the forward model becomes nearly linear with respect to ${\textbf x}$ (except for ${\delta _{\nu 0}}$). When retrieving within a short averaging time so that the integrated photon count becomes insufficient, the OD bias arising from the nonlinearity of the log function becomes significant [1]. In this case, it is more accurate to retrieve from the transmittance measurement.

5.2. Multiple-layer retrievals of atmospheric species

For multiple-layer retrievals, the RRE of ${\hat{q}_j}$ equals to the RRE of the effective DAOD $\Delta {\tau _j}$ for layer j (see Eq. (37) in Appendix C). More layers means smaller $\Delta {\tau _j}$ and larger $\sigma (\Delta {\tau _j}),$ and hence larger RRE of ${\hat{q}_j}$. $\sigma (\Delta {\tau _j})$ becomes larger than $\sigma (\Delta \tau )$ due to strong correlation of ${\tau _{ij}}$ among different layers [1]. When the laser linewidth is broadened, such correlation becomes even stronger, leading to increased $\sigma (\Delta {\tau _j})$ and hence even larger RRE of ${\hat{q}_j}$ [17]. Similarly, RSE of ${\hat{q}_j}$ becomes worse with more layers (see Eq. (39) in Appendix C). Even when the column is divided into just two layers, the RRE of CO2 mixing ratio in the boundary layer (2 km high) is 15-times higher than that for the entire column [1]. It is therefore impractical to derive the vertical profile $q_j^a(p)$ accurately from the absorption measurement itself. Such vertical profiles need to be provided as ancillary data of the atmosphere for the retrievals from lidar absorption spectra.

5.3. More IPDA lidar considerations

It is highly desirable to probe the target absorption line with more than two wavelengths so that ${\delta _{\nu 0}}$ can also be solved for. It has been shown in Eq. (39) that the second term of $\hat{{\textbf x}} - {\textbf x}$ arising from $\dot{{\boldsymbol \tau }}{\dot{{\boldsymbol \tau }}^T}{\sigma ^2}({{\delta_{\nu nslow}}} )$ vanishes if ${\delta _{\nu 0}}$ is included in ${\textbf x}.$ Since ${{\textbf x}_{i + 1}} - {{\textbf x}_i}$ in Eq. (5) is related to ${\textbf y}({{\textbf x}_i}) - {\textbf F}({{\textbf x}_i})$ by the same matrix $\hat{{\textbf S}}{{\textbf K}^T}{\textbf S}_y^{ - 1}$ as $\hat{{\textbf x}} - {\textbf x}$ is to $\bar{{\boldsymbol \varepsilon }}$ in Eq. (8), ${\textbf x}$ (with ${\delta _{\nu 0}}$ included) can be retrieved iteratively without the knowledge of ${\sigma ^2}({\delta _{\nu nslow}}).$ By solving for ${\delta _{\nu 0}},$ this unknown shift does not degrade the RRE and RSE of ${\hat{q}_j}$ significantly. In other words, an unknown frequency shift can be tolerated without causing significant retrieval degradation, as long as the drift is bound to a small fraction of the target atmospheric absorption line-width. This allows an optical reference (such as an optical etalon or a ring resonator) other than a gas absorption cell to be used to lock the master laser wavelength. Such an optical reference can be tailored to any reference wavelengths, and can be made more compact and reliable than a long path-length gas cell needed for weak gas absorption lines. The optical reference needs to be stabilized to remain virtually stationary within each wavelength sweep cycle (∼1 ms) and the slow drift needs to be bound to less than ± 1 GHz, preferably within ± 200 MHz. Such requirements are not too difficult to be met. As illustrated in section 4, the atmospheric lidar absorption measurement across the target absorption line provides an absolute wavelength calibration with high precision of ∼0.5 MHz. By solving for ${\delta _{\nu 0}},$ the optical reference can be better controlled to reduce its drift.

It is desirable for the IPDA lidar to transmit a single-spatial-mode beam. In contrast, transmitting a multi-mode beam (or multiple beams shooting at the same surface spot) requires larger transmitter aperture to illuminate the same surface spot and may lead to variations of beam intensity distribution on the surface among different laser frequency channels. Such wavelength-dependent variations may introduce unwanted spectral distortion for absorption measurement. On the other hand, the slow detector responsivity drift in either the lidar receiver or the transmitted pulse energy monitor does not degrade the retrievals as long as the detector responsivity remains virtually constant during each wavelength sweep cycle [1].

As quantified in Eq. (24) in Appendix B, the speckle noise contribution to ${\sigma ^2}({y_i})$ is $1/({n_p}{M_{sp}}{M_t}),$ and thus can be reduced by increasing the number of spatial correlation cells Msp within the receiver aperture, the number of coherent intervals Mt within a pulse duration, and the number of pulses np within the averaging time. Msp can be increased by enlarging the receiver aperture and the divergent angle of the transmitted beam. In order to reduce the speckle noise contribution to a small fraction of the shot noise contribution ${F_e}/({n_p}{K_s}),$ the single pulse photon count Ks needs to be limited to the same small fraction of ${M_{sp}}{M_t}/{F_e}.$ Unlike Ks, Msp and Mt are independent of the laser power. It is thus desirable to choose a fast pulse rate (e.g., 8 kHz) to reduce Ks and to increase np for the same average laser power. On the other hand, the peak power of the laser pulses needs to be kept high enough to overcome the solar background noise and the detector noise. The speckle correlation cells during one pulse are taken to be uncorrelated to those during another pulse because the pulse period is longer than the laser coherent time ${\tau _c}$ in practice. For a Gaussian laser line-shape with a FWHM line-width of $\Delta {\nu _{FWHM}},$ ${\tau _c} = 0.664/\Delta {\nu _{FWHM}}$ [24]. For a typical $\Delta {\nu _{FWHM}}$ of 30 MHz [11], ${\tau _c} \cong 22\;\textrm{ns}$ and ${M_t} \cong 45$ for a pulse duration of 1 µs.

6. Summary

A generalized model is presented for retrieving atmospheric constituents from lidar absorption spectra measured at any laser frequency channels. Random and systematic retrieval errors arising from measurement noise and model bias, respectively, are analyzed parametrically and numerically to provide deeper insight. By placing four or more channels symmetrically around the absorption peak, retrieval errors from a common laser frequency shift and spectral baseline tilt can be eliminated. By solving for the frequency shift and spectral baseline tilt, atmospheric retrievals degrade only slightly even when such channels are shifted substantially out of symmetry. An etalon can thus be used for the wavelength stabilization. If the laser frequencies are uncorrelated, the uncorrelated frequency drifts can drive up retrieval errors significantly. Retrieving from the transmittance measurement becomes more accurate than from the OD measurement only when the integrated photon count becomes insufficient. The vertical profile $q_j^a(p)$ needs to be provided and cannot be derived accurately from the absorption measurement itself. It is desirable to transmit a single-mode laser beam to avoid unwanted measurement errors.

Appendices

A. Effective optical depth and weighting function

The laser pulse energy E(l) attenuated by the absorption of the target species over a path length l is found to be [22]

$$\begin{array}{c} E(l) = E(0)\exp [{ - \tau ({\nu_c},l)} ]{\kern 1pt} \,,\\ \tau ({\nu _c},l) \equiv \int_0^l {{\sigma _{eff}}({\nu _c},l^{\prime}){N_{gas}}(l^{\prime})dl^{\prime}} ,\\ {\sigma _{eff}}({\nu _c},l) \equiv \int_0^\infty {{\sigma _0}({\nu _F},l){L_N}({\nu _F},l)d{\nu _F}} . \end{array}$$
Here $\tau ({\nu _c},l)$ is the effective OD of the target species accumulated along path length l, ${N_{gas}}(l)$ the number density of the target molecules, and ${\sigma _{eff}}({\nu _c},l)$ the effective absorption cross-section, ${L_N}({\nu _F},l) \equiv L({\nu _F},l)/\int_0^\infty {L({\nu _F},l)d} {\nu _F}$ the normalized laser pulse ESD, and ${\sigma _0}({\nu _F},l)$ the monochromatic absorption cross-section of the target species. To account for two-way absorption as the laser pulse travels from the spacecraft to the surface (at range $r$) and back to the spacecraft, the path length l is taken to be the accumulated distance that the laser pulse has traveled, running from 0 to $2r.$ In general, a broadened ${L_N}({\nu _F},l)$ is distorted progressively along the outgoing and return paths. This leads to two different values of ${\sigma _{eff}}({\nu _c},l)$ for each vertical position (designated by pressure p): $\sigma _{eff}^f({\nu _c},p)$ for the outgoing laser pulse and $\sigma _{eff}^b({\nu _c},p)$ for the returned laser pulse as given by
$$\begin{array}{c} \sigma _{eff}^{f,\;b}({\nu _c},p) = \frac{{\int_0^\infty {{\sigma _0}({\nu _F},p)\exp [{ - \tau_0^{f,\;b}({\nu_F},p)} ]L({\nu _F},0)d{\nu _F}} }}{{\int_0^\infty {\exp [{ - \tau_0^{f,\;b}({\nu_F},l)} ]L({\nu _F},0)d{\nu _F}} }},\\ \tau _0^f({\nu _F},p) = {\tau _0}({\nu _F},p) = \int_0^p {{\sigma _0}({\nu _F},p){N_{gas}}(p)(dr/dp)dp} ,\\ \tau _0^b({\nu _F},p) = 2{\tau _0}({\nu _F},p(r)) - {\tau _0}({\nu _F},p), \end{array}$$
where ${\tau _0}({\nu _F},p(r))$ is the one-way monochromatic OD for the column.

The two-way effective OD $\tau ({\nu _c},2r)$ is the sum of the two one-way effective ODs ${\tau ^f}({\nu _c},p(r))$ and ${\tau ^b}({\nu _c},p(r))$ for the outgoing and return path, respectively. ${\tau ^f}({\nu _c},p(r))$ (or ${\tau ^b}({\nu _c},p(r))$) can be linked to the dry mixing ratio ${q_{gas}}(p)$ and the corresponding one-way effective weighting function ${w^f}({\nu _c},p)$ (or ${w^b}({\nu _c},p)$) of the target species by [1]

$$\begin{array}{c} {\tau ^{f,\;b}}({\nu _c},p(r)) = \int_0^{p(r)} {\sigma _{eff}^{f,\;b}({\nu _c},p){N_{gas}}(p)(dr/dp)dp} = \int_0^{p(r)} {{q_{gas}}(p){w^{f,\;b}}({\nu _c},p)dp} ,\\ {w^{f,\;b}}({\nu _c},p) = \frac{{\sigma _{eff}^{f,\;b}({\nu _c},p)}}{{{m_{dryair}}g}}\frac{1}{{1 + {q_{{H_2}O}}{m_{{H_2}O}}/{m_{dryair}}}}\,. \end{array}$$
Here ${w^f}({\nu _c},p)$ and ${w^b}({\nu _c},p)$ are derived from the hydrostatic equation for p, ${m_{{H_2}O}}$ and ${m_{dryair}}$ denote the average masses of one H2O and one dry air molecule, g the gravitational acceleration of the Earth. As shown in Eqs. (16)–(17), the effective absorption cross-section ${\sigma _{eff}}({\nu _c},l)$ (and hence the effective weighting function ${w^f}({\nu _c},p)$ or ${w^b}({\nu _c},p)$) is a weighted average of its monochromatic counterpart, weighted by the normalized laser pulse ESD ${L_N}({\nu _F},l)$. Finally, the two-way effective OD $\tau ({\nu _c},2r)$ is related to the dry mixing ratio ${q_{gas}}(p)$ and the two-way effective weighting function $w({\nu _c},p)$ of the target species by
$$\begin{array}{c} \tau ({\nu _c},2r)\, = \int_0^{p(r)} {{q_{gas}}(p)w({\nu _c},p)dp} ,\\ w({\nu _c},p) \equiv {w^f}({\nu _c},p) + {w^b}({\nu _c},p). \end{array}$$
For broadened laser line-shape, the effective weighting function $w({\nu _c},p)$ also depends on ${q_{gas}}(p)$ so that $\tau ({\nu _c},2r)$ is no longer linear with respect to ${q_{gas}}(p).$ It can be shown that the layer-average $< q_j^a > $ of the dry mixing ratio profile $q_j^a(p)$ in layer j is
$$< q_j^a > \; = \frac{{\int_{{p_j}}^{{p_{j - 1}}} {q_j^a(p){{[g(1 + {q_{{H_2}O}}{m_{{H_2}O}}/{m_{dryair}})]}^{ - 1}}dp} }}{{\int_{{p_j}}^{{p_{j - 1}}} {{{[g(1 + {q_{{H_2}O}}{m_{{H_2}O}}/{m_{dryair}})]}^{ - 1}}dp} }}.$$

B. Additional details for forward model

We first summarize the noise contributions to the variance of $\hat{K}_s^i(k).$ $\hat{K}_s^i(k)$ is estimated from the total count $K_{tot}^i(k)$ within a pulse duration $\Delta t$ minus a derived background count $\hat{K}_{bgd}^i$ arising from background solar radiation, detector dark count and receiver circuitry noise. To reduce the background variance, the background count is measured in a longer duration βΔt between the pulse measurements, and scaled to $\hat{K}_{bgd}^i$ within Δt. The variance of $\hat{K}_s^i(k)$ is found to be [22]

$$\begin{array}{c} {\sigma ^2}(\hat{K}_s^i(k))\textrm{ } \equiv {F_e}\overline {\hat{K}_s^i(k)} + {\alpha ^2}{\sigma ^2}(W_s^i(k)) + {\lambda _{bgd}}\Delta t,\\ {\lambda _{bgd}} \equiv \left[ {2\alpha {N_{bgd}}{B_o}{F_e}\textrm{ + }{F_d}{\lambda_d} + {{\int_{ - \infty }^\infty {{S_\delta }_i(f)} \sin{{\textrm{c}}^2}(f\Delta t)\Delta tdf} \mathord{\left/ {\vphantom {{\int_{ - \infty }^\infty {{S_\delta }_i(f)} sin{c^2}(f\Delta t)\Delta tdf} {{{({{M_e}e} )}^2}}}} \right.} {{{({{M_e}e} )}^2}}}} \right]({1 + 1/\beta } ). \end{array}$$
Here ${F_e}$ is the detector’s excess noise factor, $2{N_{bgd}}$ the power spectral density (PSD) of the background solar radiation, Bo the optical filter bandwidth, λd an equivalent dark count rate, Fd an effective dark-count excess noise factor, ${S_{\delta i}}(f)$ the PSD of the equivalent input noise current of the circuit, Me the mean internal gain of the detector, and e the electron charge. The laser frequency noise and speckle noise enter through the second term ${\alpha ^2}{\sigma ^2}(W_s^i(k)).$ The error contributions from the transmitted pulse energy measurement can be reduced to negligible levels in practice [22,26].

The speckle noise contribution to ${\sigma ^2}(W_s^i)$ is ${(W_s^i)^2}/({M_{sp}}{M_t}),$ where ${M_{sp}}$ is the number of spatial correlation cells falling on the receiver aperture and ${M_t}$ is the number of coherent intervals contained within a pulse width $\Delta t$[24,25]. ${M_{sp}} = 2({A_R}/{A_{sp}})/(1 + {P^2})$ when the receiver aperture area ${A_R}$ is larger than the speckle correlation area ${A_{sp}},$ and ${M_{sp}}$ is doubled when the degree of polarization P of the returned signal is reduced from 1 (fully polarized) to 0 (fully depolarized). ${M_t} = \Delta t/{\tau _c}$ when the pulse width $\Delta t$ is longer than the laser coherent time ${\tau _c},$ and ${M_t} = 1$ when $\Delta t \le {\tau _c}.$ When the surface is illuminated by a Gaussian laser beam, the area ${A_{sp}}$ of the speckle correlation cell is the same as the area of the beam waist located at the transmitter aperture [25].

The bias correction factor $C_i^{OD}$ to OD measurement is given by [22]

$$C_i^{OD} \equiv{-} \frac{{{F_e}}}{2}\left( {\frac{{S_{NNK}^i}}{{n_p^2{T_i}^2}}} \right) - \frac{{{\lambda _{bgd}}\Delta t}}{2}\left( {\frac{{S_{NN}^i}}{{n_p^2{T_i}^2}}} \right)\,,$$
where $S_{NNK}^i \equiv \sum\nolimits_{k = 1}^{{n_p}} {\hat{K}_s^i(k)/[\alpha \hat{E}_s^i(k)A_z^i(k)]} {\,^2},$ and $S_{NN}^i \equiv {\sum\nolimits_{k = 1}^{{n_p}} {1/[\alpha \hat{E}_s^i(k)A_z^i(k)]} ^2}$.

Next, we summarize the covariance matrix ${{\textbf S}_y}$ of ${\textbf y}$ and the model bias ${\bar{\varepsilon }_i}$ quantified in [1]. The measurement error ${\varepsilon _i}$ of ${y_i}$ depends linearly on the following effective laser line-center frequency noise averaged across np pulses in channel i [22]

$$\delta _{\nu n}^i(t) \equiv \frac{1}{{\sum\nolimits_{k = 1}^{{n_p}} {A_s^i(k)} }}\sum\limits_{k = 1}^{{n_p}} {[{A_s^i(i)\delta_{\nu 1}^i({t + (k - 1){t_p}} )} ]\,} ,$$
where $\delta _{\nu 1}^i \equiv \nu _c^i - {\nu _i}$ is the fluctuation of the laser line-center frequency for channel i. For each channel, the laser pulses are assumed to have the same pulse shape and pulse duration $\Delta t,$ and a period of ${t_p}.$ In general, $\delta _{\nu 1}^i$ can be approximately divided into two components $\delta _{\nu 1}^i(t) = \delta _{\nu slow}^i(t) + \delta _{\nu fast}^i(t)$ [22]. The fast frequency noise component $\delta _{\nu fast}^i(t)$ can be treated as uncorrelated among different pulses and uncorrelated to the slow frequency noise component $\delta _{\nu slow}^j(t)$ of any channels. From Eq. (23), $\delta _{\nu n}^i$ can be split into $\delta _{\nu n}^i = \delta _{\nu nslow}^i + \delta _{\nu nfast}^i$, where $\delta _{\nu nslow}^i$ or $\delta _{\nu nfast}^i$ is the average of $\delta _{\nu slow}^i$ or $\delta _{\nu fast}^i$, respectively. The variance of ${y_i}$ is found to be
$$\begin{array}{c} {\sigma ^2}({y_i}) = \frac{{{F_e}}}{{\overline {S_K^i} }} + \frac{1}{{{n_p}{M_{sp}}{M_t}}} + \frac{{{n_p}{\lambda _{bgd}}\Delta t}}{{{{\overline {S_K^i} }^2}}} + [{{\sigma^2}(\delta_{\nu nfast}^i) + {\sigma^2}(\delta_{\nu slow}^i)} ]{{\dot{\tau }}_i}^2 + {({\sigma_{\tau r}^i} )^2}\,,\\ {(\sigma _{\tau r}^i)^2} \equiv {[{2{\sigma_{eff}}({\nu_i},r_G^i){N_{gas}}(p(r_G^i))\sigma_r^i} ]^2} < W_{ii}^{cov}(f) > \;. \end{array}$$
Here $S_K^i \equiv \sum\nolimits_{k = 1}^{{n_p}} {\hat{K}_s^i(k)}$ is the integrated photon count. The first term in Eq. (24) arises from the signal shot noise, the second term from the speckle noise, the third term from background solar radiation, receiver circuitry noise and the detector dark count, the fourth term from the laser line-center frequency noise, and the last term ${({\sigma_{\tau r}^i} )^2}$ from the random altimetry error. $\sigma _r^i$ is the measurement standard deviation for $r_s^i(k)$ and is assumed to be the same for each of the np pulses. $< W_{ii}^{cov}(f) > $ is the averaged height of a window function $W_{ii}^{cov}(f)$ [1] and is only slightly higher than 1/np. The variance of $\delta _{\nu nfast}^i$ is ${\sigma ^2}(\delta _{\nu nfast}^i) \cong {\sigma ^2}(\delta _{\nu fast}^i) < W_{ii}^{cov}(f) > \; \cong {\sigma ^2}(\delta _{\nu fast}^i)/{n_p}$ [22], and thus can be reduced to a negligible level by pulse averaging.

The covariance ${\mathop{\textrm {Cov}}\nolimits} ({y_i},{y_j})(i \ne j)$ depends only on the cross-channel correlation of $\delta _{\nu slow}^i.$ When the transmitted slave laser frequencies are phase-locked to a master laser, $\delta _{\nu slow}^i$ arises solely from the master laser frequency drift ${\delta _{\nu slow}}(t),$ and thus is essentially the same for all channels (see experimental results shown in Fig. 3 of [22]). ${\delta _{\nu slow}}(t)$ essentially remains unchanged within each wavelength sweep, but varies slowly over longer time scales. Consequently, $\delta _{\nu nslow}^i$ is essentially the same for all channels within the same pulse averaging period and is denoted as ${\delta _{\nu nslow}}.$ Using $\overline {\hat{K}_s^i(k)\hat{K}_s^j(k^{\prime})} = {\alpha ^2}\overline {W_s^i(k)W_s^j(k^{\prime})} \;(i \ne j)$ [22] and $\ln (x) - \overline {\ln (x)} \cong (x - \overline x )/\overline x$, ${\mathop{\textrm {Cov}}\nolimits} ({y_i},{y_j})$ is found to be

$${[{{\textbf S}_y}]_{i,j}} \cong {\mathop{\textrm {Cov}}\nolimits} [{\ln ({{T_i}} ),\ln ({{T_j}} )} ]\cong {\dot{\tau }_i}{\dot{\tau }_j}{\mathop{\textrm {Cov}}\nolimits} ({\delta_{\nu nslow}^i,\delta_{\nu nslow}^j} )\cong {\dot{\tau }_i}{\dot{\tau }_j}{\sigma ^2}({{\delta_{\nu nslow}}} )\quad (i \ne j).$$
Based on experimental data, it has been shown that the approximation ${\mathop{\textrm {Cov}}\nolimits} ({\delta_{\nu nslow}^i,\delta_{\nu nslow}^j} )\cong {\sigma ^2}({{\delta_{\nu nslow}}} )$ is accurate within 0.001% and ${\sigma ^2}({{\delta_{\nu nslow}}} )$ is only slightly reduced from ${\sigma ^2}({{\delta_{\nu slow}}} )$ by < 2% [1]. When the transmitted laser frequencies are not locked to a common master laser so that $\delta _{\nu slow}^i$ of different channels are uncorrelated, ${y_i}$ becomes uncorrelated among different channels and ${{\textbf S}_y}$ becomes diagonal.

The model bias ${\bar{\varepsilon }_i}$ (evaluated for deterministic $A_{av}^i$) is found to be

$$\begin{array}{c} {{\bar{\varepsilon }}_i} \cong{-} [{\ln (A_{av}^i) + {c_0} + {c_1}\Delta {\nu_i}} ]+ b_{\tau \nu }^i + b_{\tau C}^i + b_{\tau r}^i,\\ b_{\tau \nu }^i \equiv \frac{1}{2}{\sigma ^2}(\delta _{\nu n}^i){({{\dot{\tau }}_i})^2} - \ln \left[ {1 + \frac{1}{2}{\sigma^2}(\delta_{\nu 1}^i)\left( {{{\dot{\tau }}_i}^2 - \frac{{{d^2}\tau ({\nu_i},2r_G^i)}}{{{{(d{\nu_i})}^2}}}} \right)} \right],\\ b_{\tau C}^i \equiv{-} \frac{1}{2}{({{F_e}/\overline {S_{K^{\prime}}^i} } )^2} - \frac{3}{2}{F_e}{n_p}{\lambda _{bgd}}\Delta t/{\overline {S_{K^{\prime}}^i} ^3},\\ b_{\tau r}^i \equiv 2{\sigma _{eff}}({\nu _i},r_G^i){N_{gas}}(p(r_G^i))\delta _r^i - 2{[{{\sigma_{eff}}({\nu_i},r_G^i){N_{gas}}(p(r_G^i))\sigma_r^i} ]^2}(1 - < W_{ii}^{cov}(f) > \;)\,, \end{array}$$
where $\delta _r^i$ is the measurement bias for $r_s^i(k)$. Here the first bias term arises from uncorrected variations of $A_{av}^i$ among m channels, the second term $b_{\tau \nu }^i$ from the laser line-center frequency noise, the third term $b_{\tau C}^i$ is the residual bias for estimation of $- \ln ({\overline {A_{av}^i} } )$ with $- \ln ({A_{av}^i} )+ C_i^{OD},$ and the last term $b_{\tau r}^i$ from the altimetry bias and variance. All these bias terms can be reduced to negligible levels in practice.

C. Derivation of retrieval RRE and RSE

In this appendix, we derive the RRE and RSE for multiple-layer retrievals. Let ${n_x}$ denote the number of unknowns in the state vector ${\textbf x}$ where ${\delta _{\nu 0}}$ or ${c_1}$ may or may not be included. Correspondingly, ${n_x}$ varies from ${n_x} = {n_q} + 3$ (when ${\textbf x} = {[{q_1},{q_2},\ldots ,{q_{{n_q}}},{\kern 1pt} {\kern 1pt} {\delta _{\nu 0}},{c_1},{c_0}]^T}$) to ${n_x} = {n_q} + 1$) (when ${\textbf x} = {[{q_1},{q_2},\ldots ,{q_{{n_q}}},{\kern 1pt} {\kern 1pt} {c_0}]^T}$). The covariance ${{\textbf S}_y}$ in Eq. (4) can be inverted using the Sherman–Morrison formula

$$\begin{array}{c} {\textbf S}_y^{ - 1} = {\textbf S}_{y0}^{ - 1} - {{{\sigma ^2}({{\delta_{\nu nslow}}} )({\textbf S}_{y0}^{ - 1}\dot{{\boldsymbol \tau }}){{({\textbf S}_{y0}^{ - 1}\dot{{\boldsymbol \tau }})}^T}} \mathord{\left/ {\vphantom {{{\sigma^2}({{\delta_{\nu nslow}}} )({\textbf S}_{y0}^{ - 1}\dot{{\boldsymbol \tau }}){{({\textbf S}_{y0}^{ - 1}\dot{{\boldsymbol \tau }})}^T}} {[{1 + {\sigma^2}({{\delta_{\nu nslow}}} ){{\dot{{\boldsymbol \tau }}}^T}{\textbf S}_{y0}^{ - 1}\dot{{\boldsymbol \tau }}} ]}}} \right.} {[{1 + {\sigma^2}({{\delta_{\nu nslow}}} ){{\dot{{\boldsymbol \tau }}}^T}{\textbf S}_{y0}^{ - 1}\dot{{\boldsymbol \tau }}} ]}} = {\textbf S}_{y0}^{ - 1} - c_u^2{\textbf u}{{\textbf u}^T},\\ {[{\textbf S}_{y0}^{ - 1}]_{i,j}} = {{{\delta _{i,j}}} \mathord{\left/ {\vphantom {{{\delta_{i,j}}} {\sigma_u^2({y_i})}}} \right.} {\sigma _u^2({y_i})}},\\ {\textbf u} \equiv {[{{{{\dot{\tau }}_1}} \mathord{\left/ {\vphantom {{{{\dot{\tau }}_1}} {\sigma_u^2({y_1})}}} \right.} {\sigma _u^2({y_1})}},\;{{{{\dot{\tau }}_2}} \mathord{\left/ {\vphantom {{{{\dot{\tau }}_2}} {\sigma_u^2({y_2})}}} \right.} {\sigma _u^2({y_2})}},\;\ldots \;,{{{{\dot{\tau }}_m}} \mathord{\left/ {\vphantom {{{{\dot{\tau }}_m}} {\sigma_u^2({y_m})}}} \right.} {\sigma _u^2({y_m})}}]^T},\\ c_u^2 \equiv {{{\sigma ^2}({{\delta_{\nu nslow}}} )} \mathord{\left/ {\vphantom {{{\sigma^2}({{\delta_{\nu nslow}}} )} {\left[ {1 + {\sigma^2}({{\delta_{\nu nslow}}} )< \dot{\tau }_i^2 > \sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right].}}} \right.} {\left[ {1 + {\sigma^2}({{\delta_{\nu nslow}}} )< \dot{\tau }_i^2 > \sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right].}} \end{array}$$
From Eq. (7), ${\hat{{\textbf S}}^{ - 1}}$ is found to be
$$\begin{array}{c} {{\hat{{\textbf S}}}^{ - 1}} = {{\textbf K}^T}{\textbf S}_y^{ - 1}{\textbf K} = \hat{{\textbf S}}_0^{ - 1} - c_u^2{\textbf v}{{\textbf v}^T},\\ {\textbf v} \equiv \left( {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right){[ < {[{\textbf K}]_{i,1}}{{\dot{\tau }}_i} > ,\; < {[{\textbf K}]_{i,2}}{{\dot{\tau }}_i} > ,\;\ldots \;, < {[{\textbf K}]_{i,{n_x} - 1}}{{\dot{\tau }}_i} > ,\; < {{\dot{\tau }}_i} > ]^T}, \end{array}$$
where ${\textbf S}_0^{ - 1} = {{\textbf K}^T}{\textbf S}_{y0}^{ - 1}{\textbf K}$ can be expressed as
$${\textbf S}_0^{ - 1} = \sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({y_i})} \left( {\begin{array}{{cccc}} { < [{\textbf K}]_{i,1}^2 > }& \cdots &{ < {{[{\textbf K}]}_{i,1}}{{[{\textbf K}]}_{i,{n_x} - 1}} > }&{ < {{[{\textbf K}]}_{i,1}} > }\\ \vdots & \vdots & \vdots & \vdots \\ { < {{[{\textbf K}]}_{i,{n_x} - 1}}{{[{\textbf K}]}_{i,1}} > }& \cdots &{ < [{\textbf K}]_{i,{n_x} - 1}^2 > }&{ < {{[{\textbf K}]}_{i,{n_x} - 1}} > }\\ { < {{[{\textbf K}]}_{i,1}} > }& \cdots &{ < {{[{\textbf K}]}_{i,{n_x} - 1}} > }&1 \end{array}} \right).$$
Invoking the Sherman–Morrison formula again, $\hat{{\textbf S}}$ is found to be
$$\begin{array}{c} \hat{{\textbf S}} = {{\hat{{\textbf S}}}_0} + \Delta \hat{{\textbf S}},\\ {{\hat{{\textbf S}}}_0} = {{{\mathop{\textrm {adj}}\nolimits} ({\textbf S}_0^{ - 1})} \mathord{\left/ {\vphantom {{{\mathop{\textrm {adj}}\nolimits} ({\textbf S}_0^{ - 1})} {\det }}} \right.} {\det }}({\textbf S}_0^{ - 1}),\\ {{\Delta \hat{{\textbf S}} \equiv c_u^2({{\hat{{\textbf S}}}_0}{\textbf v}){{({{\hat{{\textbf S}}}_0}{\textbf v})}^T}} \mathord{\left/ {\vphantom {{\Delta \hat{{\textbf S}} \equiv c_u^2({{\hat{{\textbf S}}}_0}{\textbf v}){{({{\hat{{\textbf S}}}_0}{\textbf v})}^T}} {(1 - c_u^2{{\textbf v}^T}{{\hat{{\textbf S}}}_0}{\textbf v}}}} \right.} {(1 - c_u^2{{\textbf v}^T}{{\hat{{\textbf S}}}_0}{\textbf v}}}), \end{array}$$
where ${\mathop{\textrm {adj}}\nolimits} ({\textbf S}_0^{ - 1})$ is the adjugate of ${\textbf S}_0^{ - 1}.$ $\det ({\textbf S}_0^{ - 1})$ is found to be
$$\begin{array}{c} \det ({\textbf S}_0^{ - 1}) = {\left( {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right)^{{n_x}}}\det ({\textbf R})\prod\limits_{j = 1}^{{n_x} - 1} {{{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}})} ,\\ {[{\textbf R}]_{j,l}} \equiv \frac{{{{{\mathop{\textrm {Cov}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}},{{[{\textbf K}]}_{i,l}})}}{{{{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}}){{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,l}})]}^{1/2}}}}\quad (j,{\kern 1pt} \;l \le {n_x} - 1). \end{array}$$
Here $\det ({\textbf S}_0^{ - 1})$ is derived from Eq. (29) by subtracting from column l ($l = 1,\;2,\;..\;,\;{n_x} - 1$) the product of the last column and $< {[{\textbf K}]_{i,l}} > ,$ and then dividing column l and row j ($j = 1,\;2,\;..\;,\;{n_x} - 1$) by ${[{{\mathop{\textrm {Var}}\nolimits} _i}({[{\textbf K}]_{i,l}})]^{1/2}}$ and ${[{{\mathop{\textrm {Var}}\nolimits} _i}({[{\textbf K}]_{i,j}})]^{1/2}},$ respectively. Following similar steps, ${[{\mathop{\textrm {adj}}\nolimits} ({\textbf S}_0^{ - 1})]_{j,l}}$ (for $j,{\kern 1pt} \;l \le {n_x} - 1$) is found to be
$${[{\mathop{\textrm {adj}}\nolimits} ({\textbf S}_0^{ - 1})]_{j,l}} = {( - 1)^{j + l}}{\left( {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right)^{{n_x} - 1}}\frac{{\prod\limits_{j^{\prime} = 1}^{{n_x} - 1} {{{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j^{\prime}}})} }}{{{{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}}){{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,l}})]}^{1/2}}}}{M_{l,j}},$$
where ${M_{l,j}}$ is the minor of the element ${[{\textbf R}]_{l,j}}$ of matrix ${\textbf R}$ (i.e., the determinant of a matrix formed by removing from ${\textbf R}$ its lth row and jth column). We arrives at
$${[{\hat{{\textbf S}}_0}]_{j,l}} = {( - 1)^{j + l}}\frac{{{{\left[ {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right]}^{ - 1}}{M_{l,j}}}}{{{{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}}){{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,l}})]}^{1/2}}\det ({\textbf R})}}\quad (j,{\kern 1pt} \;l \le {n_x} - 1).$$
We now derive the additional covariance $\Delta \hat{{\textbf S}}$ in Eq. (30). Following the argument leading to the Cramer's rule, ${\hat{{\textbf S}}_0}{\textbf v}$ is found to be
$$\begin{array}{c} {[{{\hat{{\textbf S}}}_0}{\textbf v}]_j} = {\left( {\frac{{{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\dot{\tau }}_i})}}{{{{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}})}}} \right)^{1/2}}\frac{{\det ({\textbf R}_j^{\dot{\tau }})}}{{\det ({\textbf R})}}\quad (j \le {n_x} - 1),\\ {[{{\hat{{\textbf S}}}_0}{\textbf v}]_{{n_x}}} = \; < {{\dot{\tau }}_i} > - \sum\limits_{j = 1}^{{n_x} - 1} {({{{[{{\hat{{\textbf S}}}_0}{\textbf v}]}_j} < {{[{\textbf K}]}_{i,j}} > } )} , \end{array}$$
where matrix ${\textbf R}_j^{\dot{\tau }}$ is formed by replacing the jth column of ${\textbf R}$ by a column vector ${{\textbf r}_{\dot{\tau }}}$ defined as ${[{{\textbf r}_{\dot{\tau }}}]_l} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({[{\textbf K}]_{i,l}},{\dot{\tau }_i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({[{\textbf K}]_{i,l}}){{\mathop{\textrm {Var}}\nolimits} _i}({\dot{\tau }_i})]^{1/2}}\;(l \le {n_x} - 1).$ $1 - c_u^2{{\textbf v}^T}{\hat{{\textbf S}}_0}{\textbf v}$ is then found to be
$$\begin{array}{c} 1 - c_u^2{{\textbf v}^T}{{\hat{{\textbf S}}}_0}{\textbf v} = \frac{{1 + c_{\Delta \tau }^2{{\det ({{\textbf R}_{ext}})} \mathord{\left/ {\vphantom {{\det ({{\textbf R}_{ext}})} {\det ({\textbf R})}}} \right.} {\det ({\textbf R})}}}}{{1 + {\sigma ^2}({{\delta_{\nu nslow}}} )< \dot{\tau }_i^2 > \sum\nolimits_{i = 1}^m {\sigma _u^{ - 2}({y_i})} }},\\ {{\textbf R}_{ext}} \equiv \left( {\begin{array}{{cc}} {\textbf R}&{{{\textbf r}_{\dot{\tau }}}}\\ {{\textbf r}_{\dot{\tau }}^T}&1 \end{array}} \right), \end{array}$$
where ${{\textbf R}_{ext}}$ is a matrix of correlation coefficients extended from ${\textbf R}$ by adding a column ${{\textbf r}_{\dot{\tau }}}$ and a row ${\textbf r}_{\dot{\tau }}^T$, and a diagonal element of 1. The additional contribution to ${\sigma ^2}({x_j})$ from $\Delta \hat{{\textbf S}}$ is
$${[\Delta \hat{{\textbf S}}]_{j,j}} = \frac{{{\sigma ^2}({{\delta_{\nu nslow}}} ){{{\mathop{\textrm {Var}}\nolimits} }_i}({{\dot{\tau }}_i}){{\det }^2}({\textbf R}_j^{\dot{\tau }})}}{{[{\det ({\textbf R}) + c_{\Delta \tau }^2\det ({{\textbf R}_{ext}})} ]\det ({\textbf R}){{{\mathop{\textrm {Var}}\nolimits} }_i}({{[{\textbf K}]}_{i,j}})}}\quad (j \le {n_x} - 1).$$
From Eqs. (33) and (36), the RRE of ${\hat{q}_j}$ is found to be
$$\begin{array}{c} \frac{{\sigma ({{\hat{q}}_j})}}{{{{\hat{q}}_j}}} = \frac{{\sigma (\Delta {\tau _j})}}{{\Delta {\tau _j}}}\quad (j \le {n_q}),\\ {\sigma ^2}(\Delta {\tau _j}) \equiv \frac{{{{\left[ {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right]}^{ - 1}}{M_{j,j}}}}{{\det ({\textbf R})}} + \frac{{{\sigma ^2}({{\delta_{\nu nslow}}} ){{{\mathop{\textrm {Var}}\nolimits} }_i}({{\dot{\tau }}_i}){{\det }^2}({\textbf R}_j^{\dot{\tau }})}}{{[{\det ({\textbf R}) + c_{\Delta \tau }^2\det ({{\textbf R}_{ext}})} ]\det ({\textbf R})}}. \end{array}$$
Since the determinant of any covariance matrix (such as ${\textbf R}$ and ${{\textbf R}_x}$) is non-negative, we have ${[\Delta {\textbf S}]_{j,j}} \ge 0$.

From Eq. (8), $\delta {\textbf x}$ arising from measurement bias $\bar{{\boldsymbol \varepsilon }}$ can be derived as

$$\begin{aligned} \delta {\textbf x} \equiv \hat{{\textbf x}} - {\textbf x} &= \hat{{\textbf S}}[{{{\textbf K}^T}({\textbf S}_{y0}^{ - 1} - c_u^2{\textbf u}{{\textbf u}^T})\bar{{\boldsymbol \varepsilon }}} ]= ({{\hat{{\textbf S}}}_0} + \Delta \hat{{\textbf S}})[{\textbf w} - \left( {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right)c_u^2 < {{\dot{\tau }}_i}{{\bar{\varepsilon }}_i} > {\textbf v}]\\ &= {{\hat{{\textbf S}}}_0}{\textbf w} + \frac{{c_u^2}}{{1 - c_u^2{{\textbf v}^T}{{\hat{{\textbf S}}}_0}{\textbf v}}}\left\{ {\left[ {{{({{\hat{{\textbf S}}}_0}{\textbf v})}^T}{\textbf w} - \left( {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right) < {{\dot{\tau }}_i}{{\bar{\varepsilon }}_i} > } \right]} \right\}{{\hat{{\textbf S}}}_0}{\textbf v},\\ {\textbf w} &\equiv {{\textbf K}^T}{\textbf S}_{y0}^{ - 1}\bar{{\boldsymbol \varepsilon }} = \left( {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right){[{ < {{[{{\textbf K}_q}]}_{i,1}}{{\bar{\varepsilon }}_i} > ,\; \cdots ,\; < {{[{{\textbf K}_q}]}_{i,{n_x} - 1}}{{\bar{\varepsilon }}_i} > ,\; < {{\bar{\varepsilon }}_i} > } ]^T}. \end{aligned}$$
Again following the argument leading to the Cramer's rule, the retrieval RSE is found to be
$$\frac{{{\delta _{\hat{q}j}}}}{{{{\hat{q}}_j}}} = \frac{{{{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\bar{\varepsilon }}_i})]}^{1/2}}\det ({\textbf R}_j^\varepsilon )}}{{\det ({\textbf R})\Delta {\tau _j}}} + \frac{{c_{\Delta \tau }^2{{[{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\bar{\varepsilon }}_i})]}^{1/2}}\det ({\textbf R}_j^{\dot{\tau }})\left( {\frac{{\sum\nolimits_{l = 1}^{{n_x} - 1} {\det ({\textbf R}_l^{\dot{\tau }}){{[{{\textbf r}_\varepsilon }]}_l}} }}{{\det ({\textbf R})}} - {r_{\dot{\tau }\varepsilon }}} \right)}}{{[{\det ({\textbf R}) + c_{\Delta \tau }^2\det ({{\textbf R}_{ext}})} ]\Delta {\tau _j}}},$$
where matrix ${\textbf R}_j^\varepsilon$ is formed by replacing the jth column of ${\textbf R}$ by a column vector ${{\textbf r}_\varepsilon }$ defined as ${[{{\textbf r}_\varepsilon }]_j} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({[{\textbf K}]_{i,j}},{\bar{\varepsilon }_i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({[{\textbf K}]_{i,j}}){{\mathop{\textrm {Var}}\nolimits} _i}({\bar{\varepsilon }_i})]^{1/2}}$.

When ${\delta _{\nu 0}}$ is included in ${\textbf x},$ column ${n_q} + 1$ of ${\textbf R}$ is ${{\textbf r}_{\dot{\tau }}}$ so that $\det ({{\textbf R}_{ext}}) = 0,$ $\det ({\textbf R}_j^{\dot{\tau }}) = 0\;(j \ne {n_q} + 1),$ and $\det ({\textbf R}_{{n_q} + 1}^{\dot{\tau }}) = \det ({\textbf R}).$ The variance of ${\delta _{\nu 0}}$ then becomes

$${\sigma ^2}({\delta _{\nu 0}}) = \frac{{{{\left[ {\sum\nolimits_{i = 1}^m {\sigma_u^{ - 2}({y_i})} } \right]}^{ - 1}}{M_{{n_q} + 1,{n_q} + 1}}}}{{{{{\mathop{\textrm {Var}}\nolimits} }_i}({{\dot{\tau }}_i})\det ({\textbf R})}} + {\sigma ^2}({\delta _{\nu nslow}}).$$
Furthermore, the second term of ${\sigma ^2}(\Delta {\tau _j})$ in Eq. (37) and the second term of RSE in Eq. (39) vanish. This is also true even if ${\delta _{\nu 0}}$ is excluded from ${\textbf x},$ provided that the laser frequency channels are symmetrically distributed so that ${[{{\textbf r}_{\dot{\tau }}}]_j} \cong 0\;(j \le {n_q}),{[{{\textbf r}_{\Delta \nu }}]_j} \equiv {{\mathop{\textrm {Cov}}\nolimits} _i}({[{\textbf K}]_{i,j}},\Delta {\nu _i})/{[{{\mathop{\textrm {Var}}\nolimits} _i}({[{\textbf K}]_{i,j}}){{\mathop{\textrm {Var}}\nolimits} _i}(\Delta {\nu _i})]^{1/2}} \cong 0\;(j \le {n_q}),$ and $\det ({\textbf R}_j^{\dot{\tau }}) \cong 0\;(j \le {n_q}).$ When the laser frequency channels are symmetrically distributed, it can be shown that the RRE and RSE of ${\hat{q}_j}$ are essentially not affected by including ${\delta _{\nu 0}}$ and ${c_1}$ in ${\textbf x},$ but the first term of ${\sigma ^2}({\delta _{\nu 0}})$ is scaled by $1/(1 - r_{\dot{\tau }\Delta \nu }^2)$ if ${c_1}$ is included.

Funding

NASA Earth Sciences Division (ACT-17-0036, ATI-QRS-15-0001).

Acknowledgments

The authors gratefully acknowledge Dr. J. Mao, Dr. X. Sun and Dr. M. A. Stephen of NASA Goddard Space Flight Center for fruitful discussions. The authors are also indebted to other members of the Goddard CO2 sounder team for their support, and Dr. A. Amediek of DLR for sharing surface reflectance measurement data.

Disclosures

The authors declare no conflicts of interest.

References

1. J. R. Chen, K. Numata, and S. T. Wu, “Error reduction in retrievals of atmospheric species from symmetrically measured lidar sounding absorption spectra,” Opt. Express 22(21), 26055–26075 (2014). [CrossRef]  

2. Space Studies Board, National Research Council, Earth Science and Applications from Space: National Imperatives for the Next Decade and Beyond (National Academies, 2007).

3. S. R. Kawa, J. B. Abshire, D. F. Baker, E. V. Browell, D. Crisp, S. Crowell, J. J. Hyon, J. C. Jacob, K. W. Jucks, B. Lin, R. T. Mezies, L. E. Ott, and T. 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).

4. “A-SCOPE—advanced space carbon and climate observation of planet earth, report for assessment,” ESA-SP1313/1(European Space Agency, 2008), http://esamultimedia.esa.int/docs/SP1313-1_ASCOPE.pdf.

5. G. Ehret, C. Kiemle, M. Wirth, A. Amediek, A. Fix, and S. Houweling, “Space-borne remote sensing of CO2, CH4, and N2O by integrated path differential absorption lidar: a sensitivity analysis,” Appl. Phys. B: Lasers Opt. 90(3-4), 593–608 (2008). [CrossRef]  

6. 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]  

7. R. T. Menzies and M. T. Chahine, “Remote sensing with an airborne laser absorption spectrometer,” Appl. Opt. 13(12), 2840–2849 (1974). [CrossRef]  

8. J. B. Abshire, H. Riris, G. R. Allan, C. J. Weaver, J. Mao, X. Sun, W. E. Hasselbrack, S. R. Kawa, and S. Biraud, “Pulsed airborne lidar measurements of atmospheric CO2 column absorption,” Tellus B 62(5), 770–783 (2010). [CrossRef]  

9. J. Caron and Y. Durand, “Operating wavelengths optimization for a spaceborne lidar measuring atmospheric CO2,” Appl. Opt. 48(28), 5413–5422 (2009). [CrossRef]  

10. G. Han, H. Xu, W. Gong, X. Ma, and A. Liang, “Simulations of a multi-wavelength differential absorption lidar method for CO2 measurement,” Appl. Opt. 56(30), 8532–8540 (2017). [CrossRef]  

11. K. Numata, J. R. Chen, and S. T. Wu, “Precision and fast wavelength tuning of a dynamically phase-locked widely-tunable laser,” Opt. Express 20(13), 14234–14243 (2012). [CrossRef]  

12. X. Sun, J. B. Abshire, J. D. Beck, P. Mitra, K. Reiff, and G. Yang, “HgCdTe avalanche photodiode detectors for airborne and spaceborne lidar at infrared wavelengths,” Opt. Express 25(14), 16589–16602 (2017). [CrossRef]  

13. X. Sun and J. B. Abshire, “Comparison of IPDA lidar receiver sensitivity for coherent detection and for direct detection using sine-wave and pulsed modulation,” Opt. Express 20(19), 21291–21304 (2012). [CrossRef]  

14. K. Numata, J. R. Chen, S. T. Wu, J. B. Abshire, and M. A. Krainak, “Frequency stabilization of distributed-feedback laser diodes at 1572 nm for lidar measurements of atmospheric carbon dioxide,” Appl. Opt. 50(7), 1047–1056 (2011). [CrossRef]  

15. H. Riris, M. Rodriguez, G. R. Allan, W. Hasselbrack, J. Mao, M. Stephen, and J. Abshire, “Pulsed airborne lidar measurements of atmospheric optical depth using the Oxygen A-band at 765 nm,” Appl. Opt. 52(25), 6369–6382 (2013). [CrossRef]  

16. H. Riris, K. Numata, S. Li, S. Wu, A. Ramanathan, M. Dawsey, J. Mao, R. Kawa, and J. B. Abshire, “Airborne measurements of atmospheric methane column abundance using a pulsed integrated-path differential absorption lidar,” Appl. Opt. 51(34), 8296–8305 (2012). [CrossRef]  

17. J. R. Chen, K. Numata, and S. T. Wu, “Impact of broadened laser line-shape on retrievals of atmospheric species from lidar sounding absorption spectra,” Opt. Express 23(3), 2660–2675 (2015). [CrossRef]  

18. J. Caron, Y. Durand, J. L. Bezy, and R. Meynart, “Performance modeling for A-SCOPE, a spaceborne lidar measuring atmospheric CO2,” Proc. SPIE 7479, 747901 (2009). [CrossRef]  

19. 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]  

20. J. B. Abshire, H. Riris, C. J. Weaver, J. Mao, G. R. Allan, W. E. Hasselbrack, and E. V. Browell, “Airborne measurements of CO2 column absorption and range using a pulsed direct-detection integrated path differential absorption lidar,” Appl. Opt. 52(19), 4446–4461 (2013). [CrossRef]  

21. C. D. Rodgers, Inverse Methods for Atmospheric Sounding: Theory and Practice (World Scientific, 2000), Vol. 2.

22. J. R. Chen, K. Numata, and S. T. Wu, “Error reduction methods for integrated-path differential-absorption lidar measurements,” Opt. Express 20(14), 15589–15609 (2012). [CrossRef]  

23. A. Amediek, A. Fix, G. Ehret, J. Caron, and Y. Durand, “Airborne lidar reflectance measurements at 1.57 µm in support of the A-SCOPE mission for atmospheric CO2,” Atmos. Meas. Tech. 2(2), 755–772 (2009). [CrossRef]  

24. J. W. Goodman, Statistical Optics (John Wiley & Sons, 1985).

25. J. W. Goodman, “Some effects of target-induced scintillation on optical radar performance,” Proc. IEEE 53(11), 1688–1700 (1965). [CrossRef]  

26. 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]  

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 (4)

Fig. 1.
Fig. 1. The IPDA lidar transmits a wavelength-stepped pulse train (left) to repeatedly measure CO2 absorption at multiple laser frequency channels across a single absorption line (right).
Fig. 2.
Fig. 2. Linear least-square fitting (left) for retrieving q and c0 from monochromatic OD measurement of atmospheric CO2 line at 1572.335 nm (right). The blue dots mark four pair of symmetrical channels at their ensemble mean frequencies. The red squares mark the channels when they are shifted by 100 MHz.
Fig. 3.
Fig. 3. (left) The measurement noise σ(yi) for atmospheric CO2 (solid black) as a function of 2τ0 (for frequency offset < 0), computed using parameters listed in Table 1. The blue dots mark the channels ν1 to ν4. Also plotted are partial contributions to σ(yi) from the signal shot noise (solid grey), slow laser line-center frequency drift (dashed red), solar background (dotted brown), receiver circuitry noise (dash-dotted green), and detector dark count (long-dashed blue). (right) The RRE of $\hat{q}$ (solid black), ${r_{\dot{\tau }}}$ (dotted red), ${r_{\Delta \nu }}$ (dash-dotted green), and ${r_{\dot{\tau }\Delta \nu }}$ (dashed blue) as functions of a common frequency shift ${\delta _{\nu 0}}.$ The RRE of $\hat{q}$ becomes much larger (solid grey) when the transmitted laser frequencies become uncorrelated.
Fig. 4.
Fig. 4. (left) The RREs of $\hat{q}$ for two-channel (solid grey) and four-channel (solid black) lidars as functions of the absolute value of the online frequency offset. Δν1 is fixed at -15.6 GHz for two channels. For four channels, -Δν1 = Δν4 = 15.6 GHz and -Δν2 = Δν3. Also plotted for two channels are the partial RSEs of $\hat{q}$ due to a frequency bias of 3 MHz (dashed red) and a small OD slope of 3.3×10−4/GHz (dotted green). (right) RSEs of $\hat{q}$ for 1-s averaging time calculated from the surface reflectance data as functions of the starting position along the path for the 8 symmetrical channels (grey), four symmetrical channels (blue), and two-channels (red).

Tables (1)

Tables Icon

Table 1. Parameters used for numerical estimation of retrieval errors

Equations (40)

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

W s i = E s i A s i exp [ τ ( ν c i , 2 r s i ) ] ,
T i 1 n p k = 1 n p K ^ s i ( k ) / [ α E ^ s i ( k ) A z i ( k ) ] , A z i ( k ) exp [ τ ( ν i , 2 r G i ) τ ( ν i , 2 r s i ( k ) ) ] exp [ ( q 1 / < q 1 a > ) q 1 a ( p ( r G i ) ) p ( r G i ) p ( r s i ( k ) ) w ( ν i , p ) d p ] .
y i = [ F ( x , b ) ] i + ε i = j = 1 n q [ K q ] i , j q j + c 0 + c 1 Δ ν i + ε i , [ K q ] i , j p j p j 1 q j a ( p ) w ( ν i + δ ν 0 , p ) d p / p j p j 1 q j a ( p ) w ( ν i + δ ν 0 , p ) d p < q j a > < q j a > ( j n q ) ,
S y = S y 0 + σ 2 ( δ ν n s l o w ) τ ˙ τ ˙ T , [ S y 0 ] i , j = σ u 2 ( y i ) δ i , j , σ u 2 ( y i ) σ 2 ( y i ) τ ˙ i 2 σ 2 ( δ ν n s l o w ) .
x i + 1 = x i + ( K i T S y 1 K i ) 1 { K i T S y 1 [ y ( x i ) F ( x i ) ] } , [ K ( x ) ] i , j [ F ( x ) y ] i x j [ F ( x ) ] i x j ,
x ^ L = ( K T S y 1 K ) 1 K T S y 1 y .
S ^ = ( K T S y 1 K ) 1 .
x ^ x = S ^ K T S y 1 [ ε ¯ + ( F ( x ^ , b ) F ( x ^ , b ^ ) ) ] .
σ ( q ^ ) q ^ = σ ( Δ τ ) Δ τ , σ 2 ( Δ τ ) = [ i = 1 m σ u 2 ( y i ) ] 1 + σ 2 ( δ ν n s l o w ) Var i ( τ ˙ i ) r τ ˙ 2 1 + c Δ τ 2 ( 1 r τ ˙ 2 ) ,
δ q ^ q ^ = δ Δ τ Δ τ , δ Δ τ = r ε [ Var i ( ε ¯ i ) ] 1 / 2 + c Δ τ 2 [ Var i ( ε ¯ i ) ] 1 / 2 ( r τ ˙ r ε r τ ˙ ε ) r τ ˙ 1 + c Δ τ 2 ( 1 r τ ˙ 2 ) ,
σ ( q ^ ) q ^ = [ i = 1 m σ u 2 ( y i ) ] 1 / 2 ( 1 r τ ˙ 2 ) 1 / 2 Δ τ ,
δ q ^ q ^ = ( r ε r τ ˙ ε r τ ˙ ) [ Var i ( ε ¯ i ) ] 1 / 2 ( 1 r τ ˙ 2 ) Δ τ ,
σ 2 ( δ ^ ν 0 ) = [ i = 1 m σ u 2 ( y i ) ] 1 Var i ( τ ˙ i ) ( 1 r τ ˙ 2 ) + σ 2 ( δ ν n s l o w ) .
T i = [ F t ( x , b ) ] i = exp ( [ j = 1 n q [ K q ] i , j q j + c 0 + c 1 Δ ν i ] ) + ε i t .
ε i ε ¯ i ( ε i t ε ¯ i t ) T ¯ i ( 1 ε ¯ i t T ¯ i ) + ( ε i t ε ¯ i t ) 2 ( ε i t ε ¯ i t ) 2 ¯ 2 T ¯ i 2 ( ε i t ε ¯ i t ) T ¯ i ( 1 ε ¯ i t T ¯ i ) .
E ( l ) = E ( 0 ) exp [ τ ( ν c , l ) ] , τ ( ν c , l ) 0 l σ e f f ( ν c , l ) N g a s ( l ) d l , σ e f f ( ν c , l ) 0 σ 0 ( ν F , l ) L N ( ν F , l ) d ν F .
σ e f f f , b ( ν c , p ) = 0 σ 0 ( ν F , p ) exp [ τ 0 f , b ( ν F , p ) ] L ( ν F , 0 ) d ν F 0 exp [ τ 0 f , b ( ν F , l ) ] L ( ν F , 0 ) d ν F , τ 0 f ( ν F , p ) = τ 0 ( ν F , p ) = 0 p σ 0 ( ν F , p ) N g a s ( p ) ( d r / d p ) d p , τ 0 b ( ν F , p ) = 2 τ 0 ( ν F , p ( r ) ) τ 0 ( ν F , p ) ,
τ f , b ( ν c , p ( r ) ) = 0 p ( r ) σ e f f f , b ( ν c , p ) N g a s ( p ) ( d r / d p ) d p = 0 p ( r ) q g a s ( p ) w f , b ( ν c , p ) d p , w f , b ( ν c , p ) = σ e f f f , b ( ν c , p ) m d r y a i r g 1 1 + q H 2 O m H 2 O / m d r y a i r .
τ ( ν c , 2 r ) = 0 p ( r ) q g a s ( p ) w ( ν c , p ) d p , w ( ν c , p ) w f ( ν c , p ) + w b ( ν c , p ) .
< q j a > = p j p j 1 q j a ( p ) [ g ( 1 + q H 2 O m H 2 O / m d r y a i r ) ] 1 d p p j p j 1 [ g ( 1 + q H 2 O m H 2 O / m d r y a i r ) ] 1 d p .
σ 2 ( K ^ s i ( k ) )   F e K ^ s i ( k ) ¯ + α 2 σ 2 ( W s i ( k ) ) + λ b g d Δ t , λ b g d [ 2 α N b g d B o F e  +  F d λ d + S δ i ( f ) sin c 2 ( f Δ t ) Δ t d f / S δ i ( f ) s i n c 2 ( f Δ t ) Δ t d f ( M e e ) 2 ( M e e ) 2 ] ( 1 + 1 / β ) .
C i O D F e 2 ( S N N K i n p 2 T i 2 ) λ b g d Δ t 2 ( S N N i n p 2 T i 2 ) ,
δ ν n i ( t ) 1 k = 1 n p A s i ( k ) k = 1 n p [ A s i ( i ) δ ν 1 i ( t + ( k 1 ) t p ) ] ,
σ 2 ( y i ) = F e S K i ¯ + 1 n p M s p M t + n p λ b g d Δ t S K i ¯ 2 + [ σ 2 ( δ ν n f a s t i ) + σ 2 ( δ ν s l o w i ) ] τ ˙ i 2 + ( σ τ r i ) 2 , ( σ τ r i ) 2 [ 2 σ e f f ( ν i , r G i ) N g a s ( p ( r G i ) ) σ r i ] 2 < W i i c o v ( f ) > .
[ S y ] i , j Cov [ ln ( T i ) , ln ( T j ) ] τ ˙ i τ ˙ j Cov ( δ ν n s l o w i , δ ν n s l o w j ) τ ˙ i τ ˙ j σ 2 ( δ ν n s l o w ) ( i j ) .
ε ¯ i [ ln ( A a v i ) + c 0 + c 1 Δ ν i ] + b τ ν i + b τ C i + b τ r i , b τ ν i 1 2 σ 2 ( δ ν n i ) ( τ ˙ i ) 2 ln [ 1 + 1 2 σ 2 ( δ ν 1 i ) ( τ ˙ i 2 d 2 τ ( ν i , 2 r G i ) ( d ν i ) 2 ) ] , b τ C i 1 2 ( F e / S K i ¯ ) 2 3 2 F e n p λ b g d Δ t / S K i ¯ 3 , b τ r i 2 σ e f f ( ν i , r G i ) N g a s ( p ( r G i ) ) δ r i 2 [ σ e f f ( ν i , r G i ) N g a s ( p ( r G i ) ) σ r i ] 2 ( 1 < W i i c o v ( f ) > ) ,
S y 1 = S y 0 1 σ 2 ( δ ν n s l o w ) ( S y 0 1 τ ˙ ) ( S y 0 1 τ ˙ ) T / σ 2 ( δ ν n s l o w ) ( S y 0 1 τ ˙ ) ( S y 0 1 τ ˙ ) T [ 1 + σ 2 ( δ ν n s l o w ) τ ˙ T S y 0 1 τ ˙ ] [ 1 + σ 2 ( δ ν n s l o w ) τ ˙ T S y 0 1 τ ˙ ] = S y 0 1 c u 2 u u T , [ S y 0 1 ] i , j = δ i , j / δ i , j σ u 2 ( y i ) σ u 2 ( y i ) , u [ τ ˙ 1 / τ ˙ 1 σ u 2 ( y 1 ) σ u 2 ( y 1 ) , τ ˙ 2 / τ ˙ 2 σ u 2 ( y 2 ) σ u 2 ( y 2 ) , , τ ˙ m / τ ˙ m σ u 2 ( y m ) σ u 2 ( y m ) ] T , c u 2 σ 2 ( δ ν n s l o w ) / σ 2 ( δ ν n s l o w ) [ 1 + σ 2 ( δ ν n s l o w ) < τ ˙ i 2 > i = 1 m σ u 2 ( y i ) ] . [ 1 + σ 2 ( δ ν n s l o w ) < τ ˙ i 2 > i = 1 m σ u 2 ( y i ) ] .
S ^ 1 = K T S y 1 K = S ^ 0 1 c u 2 v v T , v ( i = 1 m σ u 2 ( y i ) ) [ < [ K ] i , 1 τ ˙ i > , < [ K ] i , 2 τ ˙ i > , , < [ K ] i , n x 1 τ ˙ i > , < τ ˙ i > ] T ,
S 0 1 = i = 1 m σ u 2 ( y i ) ( < [ K ] i , 1 2 > < [ K ] i , 1 [ K ] i , n x 1 > < [ K ] i , 1 > < [ K ] i , n x 1 [ K ] i , 1 > < [ K ] i , n x 1 2 > < [ K ] i , n x 1 > < [ K ] i , 1 > < [ K ] i , n x 1 > 1 ) .
S ^ = S ^ 0 + Δ S ^ , S ^ 0 = adj ( S 0 1 ) / adj ( S 0 1 ) det det ( S 0 1 ) , Δ S ^ c u 2 ( S ^ 0 v ) ( S ^ 0 v ) T / Δ S ^ c u 2 ( S ^ 0 v ) ( S ^ 0 v ) T ( 1 c u 2 v T S ^ 0 v ( 1 c u 2 v T S ^ 0 v ) ,
det ( S 0 1 ) = ( i = 1 m σ u 2 ( y i ) ) n x det ( R ) j = 1 n x 1 Var i ( [ K ] i , j ) , [ R ] j , l Cov i ( [ K ] i , j , [ K ] i , l ) [ Var i ( [ K ] i , j ) Var i ( [ K ] i , l ) ] 1 / 2 ( j , l n x 1 ) .
[ adj ( S 0 1 ) ] j , l = ( 1 ) j + l ( i = 1 m σ u 2 ( y i ) ) n x 1 j = 1 n x 1 Var i ( [ K ] i , j ) [ Var i ( [ K ] i , j ) Var i ( [ K ] i , l ) ] 1 / 2 M l , j ,
[ S ^ 0 ] j , l = ( 1 ) j + l [ i = 1 m σ u 2 ( y i ) ] 1 M l , j [ Var i ( [ K ] i , j ) Var i ( [ K ] i , l ) ] 1 / 2 det ( R ) ( j , l n x 1 ) .
[ S ^ 0 v ] j = ( Var i ( τ ˙ i ) Var i ( [ K ] i , j ) ) 1 / 2 det ( R j τ ˙ ) det ( R ) ( j n x 1 ) , [ S ^ 0 v ] n x = < τ ˙ i > j = 1 n x 1 ( [ S ^ 0 v ] j < [ K ] i , j > ) ,
1 c u 2 v T S ^ 0 v = 1 + c Δ τ 2 det ( R e x t ) / det ( R e x t ) det ( R ) det ( R ) 1 + σ 2 ( δ ν n s l o w ) < τ ˙ i 2 > i = 1 m σ u 2 ( y i ) , R e x t ( R r τ ˙ r τ ˙ T 1 ) ,
[ Δ S ^ ] j , j = σ 2 ( δ ν n s l o w ) Var i ( τ ˙ i ) det 2 ( R j τ ˙ ) [ det ( R ) + c Δ τ 2 det ( R e x t ) ] det ( R ) Var i ( [ K ] i , j ) ( j n x 1 ) .
σ ( q ^ j ) q ^ j = σ ( Δ τ j ) Δ τ j ( j n q ) , σ 2 ( Δ τ j ) [ i = 1 m σ u 2 ( y i ) ] 1 M j , j det ( R ) + σ 2 ( δ ν n s l o w ) Var i ( τ ˙ i ) det 2 ( R j τ ˙ ) [ det ( R ) + c Δ τ 2 det ( R e x t ) ] det ( R ) .
δ x x ^ x = S ^ [ K T ( S y 0 1 c u 2 u u T ) ε ¯ ] = ( S ^ 0 + Δ S ^ ) [ w ( i = 1 m σ u 2 ( y i ) ) c u 2 < τ ˙ i ε ¯ i > v ] = S ^ 0 w + c u 2 1 c u 2 v T S ^ 0 v { [ ( S ^ 0 v ) T w ( i = 1 m σ u 2 ( y i ) ) < τ ˙ i ε ¯ i > ] } S ^ 0 v , w K T S y 0 1 ε ¯ = ( i = 1 m σ u 2 ( y i ) ) [ < [ K q ] i , 1 ε ¯ i > , , < [ K q ] i , n x 1 ε ¯ i > , < ε ¯ i > ] T .
δ q ^ j q ^ j = [ Var i ( ε ¯ i ) ] 1 / 2 det ( R j ε ) det ( R ) Δ τ j + c Δ τ 2 [ Var i ( ε ¯ i ) ] 1 / 2 det ( R j τ ˙ ) ( l = 1 n x 1 det ( R l τ ˙ ) [ r ε ] l det ( R ) r τ ˙ ε ) [ det ( R ) + c Δ τ 2 det ( R e x t ) ] Δ τ j ,
σ 2 ( δ ν 0 ) = [ i = 1 m σ u 2 ( y i ) ] 1 M n q + 1 , n q + 1 Var i ( τ ˙ i ) det ( R ) + σ 2 ( δ ν n s l o w ) .
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.