1. Introduction
As discussed previously in Nalli et al. (2012, 2013a), accurate satellite observations (obs) and calculations (calc) of clear-sky, top-of-atmosphere (TOA) spectral radiances are necessary for retrieval of environmental data records (EDRs) from satellite infrared (IR) sounder and imager remote sensing systems. IR-based EDR physical retrieval algorithms are based upon the minimization of clear-sky obs minus calc (obs − calc, or equivalently from the forward modeling perspective, calc − obs). Therefore, it is important that differences between observation and calculation be minimal under well-characterized conditions over the range of satellite zenith angles θ. A systematic angular dependence in obs − calc may lead to undesirable scan-dependent artifacts and/or errors in the calibration/validation (cal/val) of sensor data records (SDRs) and, consequently, EDRs. Current satellite-based IR sounding systems are based upon the measurement of onboard-calibrated high-spectral-resolution radiances on the order of thousands of channels over the IR spectrum (i.e., “hyperspectral” radiances) (e.g., Smith et al. 2009) and include the Joint Polar Satellite System (JPSS) Cross-Track Infrared Sounder (CrIS) on board the Suomi–National Polar-Orbiting Partnership (SNPP) satellite (Goldberg et al. 2013), the MetOp Infrared Atmospheric Sounding Interferometer (IASI) (Cayla 1993; Hilton et al. 2012) and the Aqua Atmospheric Infrared Sounder (AIRS) (Chahine et al. 2006). Because water-droplet clouds are typically opaque in the IR spectrum, EDR retrievals must assume some degree of cloud-free radiative transfer within the sensor field of view (FOV) or field of regard (FOR). In the case of sounding systems, radiances from microwave (MW) sounders—for example, the Advanced Technology Microwave Sounder (ATMS) on board SNPP (Weng et al. 2012)—are utilized to “cloud clear” the IR spectra within partly cloudy FORs (e.g., Susskind et al. 2003). In the case of narrowband imager systems, a “cloud-mask” algorithm is applied (e.g., Thomas et al. 2004), using radiances from IR channels, as well as solar-spectrum [or “visible” (VIS)] channels over sunlit portions of Earth (i.e., IR/VIS), to screen cloudy FOVs. The remaining “clear sky” FOVs can then be used for EDR retrievals—for example, global sea surface temperature (SST) and aerosol optical depth (AOD).
As discussed in Nalli et al. (2012, 2013a), a fundamental problem in clear-sky calc − obs analyses is the assumption of pure global clear-sky observations, when in reality we only have access to observations obtained from either cloud-clearing (in the case of IR/MW sounder systems) or cloud-masking (in the case of IR/VIS imager systems) algorithms as mentioned above. Because of this, clear-sky radiance products (i.e., clear-sky obs) will be subject to algorithmic errors, something commonly referred to as “cloud contamination.” Generally speaking, a small degree of residual clouds (e.g., subpixel clouds) and aerosols can remain in clear-sky radiances regardless of the algorithm used. This may lead to an observation that is cold-biased relative to a clear-sky calculation for channels sensitive to tropospheric layers below the cloud top (e.g., Benner and Curry 1998; Nalli and Stowe 2002; Sokolik 2002; Maddy et al. 2011; Nalli et al. 2012). Furthermore, global environmental satellite observations are obtained from sensors that scan Earth at oblique local zenith angles θ, and it is well established that apparent cloud cover increases with angle owing to the decrease in probability of clear (cloud free) lines of sight (PCLoS) (e.g., Kauth and Penquite 1967; Lund and Shanklin 1972, 1973). Subpixel cloud contamination in an ensemble of observations (e.g., arising from false negatives in the cloud-mask result) may thus have angular dependence, assuming the occurrence of false negatives does not vary appreciably with view geometry. To quantify IR brightness temperature sensitivity to angularly varying cloud contamination, Nalli et al. (2012) derived simple models for idealized “superwindow” channels assuming that the angular variation in probability for cloud contamination on average would behave inversely as PCLoS for a very small absolute cloud fraction (which is similar to assuming the cloud-mask algorithm has a small, angularly independent fraction of false negatives in regions consisting of broken, subpixel clouds with small absolute cloud fractions). Aerosol contamination was also assumed to behave according to the increased slant path arising from a small AOD. It was found that very small levels of contamination can theoretically lead to measurable angular effects with a “concave up” signal in calc − obs brightness temperatures1 on the order of hundreds of millikelvins or more (cf. Nalli et al. 2012). In the follow-up companion paper (i.e., Nalli et al. 2013a), analyses of satellite hyperspectral sounder and narrowband imager systems were conducted. Based on the sounder cloud-cleared radiances (CCRs) and collocated radiosonde observations (RAOBs), calc − obs analyses were performed that showed a possible impact of clouds on the systematic error (bias) of sounder temperature EDR retrievals near the surface on the order of ≃−1 to −3 K. Wong et al. (2015) have since confirmed cloud contamination biases in lower-troposphere temperature profile EDRs (AIRS version 6) of approximately −2 K based on a thorough analysis against global RAOBs and Moderate Resolution Imaging Spectroradiometer (MODIS) cloud pressure and optical depth estimates.
In the current work we examine the potential cloud impact on the angular distribution of IR calc − obs using hyperspectral radiance observations obtained from aircraft overflying clear-to-partly-cloudy atmospheric conditions during an ocean-based validation field experiment. We conduct analyses of microwindow calc − obs as a function of sensor zenith angle based on these observations along with concurrent dropsonde observations. The observations were originally intended to be clear sky for the purposes of satellite IR sounder cal/val, but low-level microscale clouds had subsequently formed during the observing period. Because the clouds were difficult to detect in satellite imagery, we realized that the dataset provided sample conditions under which satellite radiance observations may be cloud contaminated. Unlike satellite-based cloud-cleared sounder data, the aircraft-based spectral data are at a high spatial resolution comparable to satellite imager data, with high spectral resolution to allow for careful selection of channels to minimize the impacts of absorbing gas uncertainties in the forward calculation (e.g., Nalli and Smith 2003).
2. Radiative transfer model
This section overviews the radiative transfer model (RTM) used for high-spectral-resolution quasimonochromatic radiance calculations valid for ocean surfaces considered by this work.
a. Microwindow channel selection
To minimize uncertainties arising from gas absorption deviating from atmospheric state parameter inputs (including errors in raob-measured H2O, as well as assumed values for fixed gases), we carefully selected six spectral microwindows (i.e., spectral regions of high transmittance located between absorption lines) minimally impacted by absorbing species in the longwave IR (LWIR) region roughly spanning 650–1200 cm−1 [cf. Nalli and Smith (2003), their Fig. 1]. Although there are more transparent microwindows in the shortwave IR (SWIR) region (roughly spanning 2000–2700 cm−1 and largely due to minimized H2O continuum absorption) the use of daytime data containing sun glint correlated with zenith view angle precluded any possible exploitation of such channels, even with the sun-glint estimation detailed in appendix A.
The peak of the LWIR H2O continuum transmittance is located on the SWIR side of the O3 band (i.e., ≃1080–1200 cm−1). However, for a sensor located at high altitudes it is desirable to avoid the O3 band as much as possible. For satellite sounder analyses, Nalli et al. (2013a) utilized LWIR microwindow channels located just outside of this region to minimize trace gas and H2O continuum uncertainty for a sensor located at the TOA, defined by
b. IR radiative transfer equation
c. SLR
In this paper we employ the “effective emissivity” model (Nalli et al. 2008b,a) that was developed to be used in conjunction with the reflection approximation [Eq. (5)] and has since been implemented within the Community Radiative Transfer Model (CRTM).2 For theoretical details on the sea surface emissivity model, the reader is referred to Nalli et al. (2001, 2008b,a) and references therein.
d. Atmospheric transmittance and radiance model
To ensure accurate high-spectral-resolution forward radiance calculations using Eq. (1) (i.e., calc), quasimonochromatic atmospheric transmittances
3. Aircraft campaign IR microwindow analysis
In an effort to examine more closely the cloud/aerosol angular impact on radiance observations independent of a cloud-clearing or cloud-mask algorithm, we turned to field campaign data that include radiance spectra obtained from an aircraft-based Fourier transform spectrometer (FTS). Specifically, we used data obtained from the Joint Airborne IASI Validation Experiment (JAIVEx) (Newman et al. 2012) during the 29 April 2007 clear-sky overflight of the Gulf of Mexico (e.g., Larar et al. 2010). Ocean-based campaigns such as this provide the tightest control on the lower-boundary (surface) radiometric variables (e.g., Nalli et al. 2006).
a. Experimental overview
Radiance observations at high spectral resolution (unapodized, 0.25 cm−1) were obtained during JAIVEx from the NPP Atmospheric Sounder Test Bed-Interferometer (NAST-I) (Smith et al. 2005), an FTS system similar to IASI and CrIS, but designed for high-altitude aircraft in support of JPSS sounder risk reduction and cal/val. NAST-I was flown on board the NASA WB-57 aircraft during JAIVEx over the Gulf of Mexico spanning 1523–1940 UTC (0923–1340 LST) on 29 April 2007. The WB-57 flew at altitudes ranging between
Complementing the NAST-I radiances are a total of 20 Vaisala dropsondes that were launched from a Facility for Airborne Atmospheric Measurements (FAAM) BAe 146 aircraft underflying the WB-57 at an altitude of ≃7–8 km. Additionally, a hemispheric camera was mounted on the WB-57 main fuselage for all-sky visible color imagery capable of resolving microscale cloud features spanning the swath scanned by the NAST-I sensor. A total of 712 JPEG images were obtained covering much (but not all) of the NAST-I sampling period, and these were made available to us courtesy of V. Leslie (MIT Lincoln Laboratory). The file time stamp had an undetermined “offset” from the actual UTC time (V. Leslie 2010, personal communication). We therefore conducted a careful visual inspection of the images looking for known changes in the aircraft heading (using the NAST-I GPS latitude–longitude and heading data), and using the solar-glint disk (located east-southeast to southeast in the morning, south around noon to south-southwest by early afternoon), we determined the time-stamp offset to be approximately
Figure 1 shows the WB-57 flight path and locations of the dropsonde launches during the NAST-I sampling period (1523–1940 UTC) overlying two U.S. National Oceanic and Atmospheric Adminstration (NOAA) Geostationary Operational Environmental Satellite (GOES-12) visible channel (band 1, 0.65 μm) images taken within the same time frame. Given that the occurrence of cloud-free FOVs for a IR sounder such as CrIS or IASI is
b. Cloud-cover information from GOES aerosol data
To obtain quantitative characterizations of these MBL FWC (as well as any potential aerosols/haze), we realized that, barring a coregistered imager operating synergistically with the NAST-I on board the aircraft, visible (reflected solar) data acquired from geostationary Earth orbit (GEO) is the best available alternative option. It was also clear that an algorithm designed for detecting very small backscatter signals would be necessary. Recognizing that solar-spectrum-based AOD EDR retrieval algorithms fall into that category, we utilized the GOES Aerosol/Smoke Product (GASP) developed by K. Knapp and collaborators at NESDIS/Center for Satellite Applications and Research (STAR) (Knapp et al. 2002; Prados et al. 2007). The GASP algorithm retrieves total column AOD from GEO orbit by first removing the invariant “background” solar reflectance signal using an image composite assembled from the previous month, thereby allowing small transient anomalies (i.e., backscatter due to aerosol, smoke, subpixel cloud) to be detected (Knapp et al. 2002). In essence, the GASP algorithm interprets low-signal, solar-spectrum atmospheric backscattering as AOD.
Figure 3 shows GASP AOD imagery at nominal 4 × 4 km2 nadir resolution (Prados et al. 2007) corresponding to the GOES visible channel imagery shown Fig. 1. The AOD distribution was found to span unphysical negative values, so the product distribution was shifted by 0.15 minus the uncorrected mode (which was 0.05), thereby resetting the mode of the distribution to be 0.15, a typical background marine level. It can be seen that the FWC observed by the JAIVEx hemispheric camera appear as intermittent regions of high GASP AOD (
In using GOES aerosol data for estimating “cloudiness” in the NAST-I FOVs, we first note that a space–time interpolation is necessary. The GASP data are not gridded in geographic coordinates, so we interpolated half-hourly GASP AOD fields (AOD means and standard deviation) to the NAST-I FOV latitude–longitude coordinates. However, the GASP data are derived from the FOVs of the GOES-12 imager located at (0°N, 75°W). Strictly speaking, the atmospheric paths observed within the GOES FOVs are not the same as those of the NAST-I FOVs given the different vantage points of the two sensors. To account for the different FOVs, we performed a remapping which then enabled the spatial interpolation mentioned above; our method for doing this is detailed in appendix B. The space-interpolated fields were then linearly interpolated in time to the NAST-I GPS times. Figure 4 shows the resulting interpolated values (without cloud masking) along with the evolution of the FWC cloud fields (as well as any aerosol) throughout the aircraft flight duration. The application of these data is described more in section 3e.
c. Methodology for computing calc − obs
After employing this procedure for obtaining profiles at the NAST-I FOVs, we then kept only those FOVs within 100 km and 1 h of the launches, which eliminated a small fraction of the FOVs (
Surface parameters were specified as follows. Surface wind speeds
d. Modeled LWIR impact of broken FWC clouds
We can estimate the cloud-base height from Eq. (12),
Because the dropsonde data show a stable capping inversion above a shallow mixed layer (Fig. 5), we would expect these FWC clouds to be “forced” (i.e., forced primarily by the original mixed-layer thermal) as opposed to “active” (i.e., cloud buoyancy beyond the thermal due to latent heat release) according to the cloud classification of Stull (1985). Translating to World Meteorological Organization (WMO) cloud definitions, forced clouds are typically cumulus humilis (i.e., slight vertical extent) but also possibly cumulus mediocris (i.e., moderate vertical extent) (WMO 1956, 1987; Stull 1985). However, the shadow analysis conducted in appendix C suggests
e. Results and discussion
To examine the impact of the observed FWC clouds (and residual ambient aerosols) on the angular variation of calc − obs analyses, we placed data in angular bins centered on the NAST-I nadir scan angles
Figure 7 provides box plot statistical summaries of calc − obs results for the
Calculations using the GOES IR SST (Figs. 8c, 9c, and 10c) are more consistent with the modeled results (Figs. 8d, 9d, and 10d), especially in the sense that there is a positive bias throughout the scan as would be expected. It is not clear why the results using RSS blended SSTs show near-zero bias (even slightly negative) at the smaller angles, although it may be related to unresolved spatiotemporal features in the dataset. It can be seen that results for these microwindows are very similar to one another, thus providing us greater confidence in our calculations.3 Regardless of the SST dataset, there are distinct concave-up signals in the double-difference plots (Figs. 8a,c, 9a,c, and 10a,c) ranging from ≃0.2 to 0.4 K. These are consistent in magnitude (albeit somewhat larger) with the
We note these results are based upon a total of
In terms of EDR impact, assuming near-unity transmittance in surface-sensitive microwindow channels (e.g., the AIRS 2616 cm−1 “superwindow”), the angular variation of bias of surface skin/air temperature EDR retrievals from undetected clouds in clear-sky radiance products would be on the order of the inverse of those shown in Figs. 8–10 for calc − obs double differences (Figs. 8a,c, 9a,c, and 10a,c) and modeled PCLoS sensitivity (Figs. 8b,d, 9b,d, and 10b,d)—namely, from
Finally, it is also worth noting that our papers on the angular effect of clouds have methodically extended the application of the PCLoS model, including three general cloud shapes, from visual-based remote sensing and radiative flux applications (e.g., Taylor and Ellingson 2008) to passive IR remote sensing applications. In this paper we have extended our applications of the PCLoS model toward the estimation of cloud aspect ratios from cloud shadows, which we envision may be useful for future aircraft campaigns equipped with all-sky cameras, or in analyses of clouds within satellite visible imagery, or in future applications where cloud shadowing might be of interest (e.g., radiative transfer in atmospheres with one than more cloud layer).
Acknowledgments
This research was supported by the NOAA Joint Polar Satellite System (JPSS-STAR) Office (L. Zhou) and the NOAA/NESDIS/STAR Satellite Meteorology and Climatology Division (SMCD) (F. Weng). We are grateful to the following individuals for their contributions in support of this work: V. Leslie (MIT Lincoln Lab) for providing the all-sky camera images and for discussions pertaining to the image time-stamp offset, S. Kondragunta and C. Xu (NESDIS/STAR) for providing the NOAA GASP aerosol EDR product for April 2007, A. Ignatov and X. Liang (NESDIS/STAR) for bringing our attention to the possibility of sun-glint contamination in LWIR channels, and C. D. Barnet (STC, formerly NESDIS/STAR) for his support of the earlier related papers. We also thank the three anonymous reviewers who carefully reviewed two manuscript iterations and provided constructive feedback that we used to strengthen the quality of the final paper. Determination of the local solar zenith and azimuth angles for the JAIVEx glint analysis was made possible via code developed by V. Roy and distributed at the MATLAB Central File Exchange (sun_position.m). The views, opinions, and findings contained in this paper are those of the authors and should not be construed as an official NOAA or U.S. Government position, policy, or decision.
APPENDIX A
Sun-Glint Calculation
As discussed above in section 2c, sun glint is an issue in the JAIVEx dataset that must be taken into consideration. It is clearly evident in the hemispheric camera images (e.g., Fig. 2) that the reflected solar disk systematically persists near nadir throughout the observing period. Because the aircraft flight path assumed a predominately north–south course, angular scanning was predominately east–west over a continuous limited space–time domain. This means that, unlike the satellite-based analyses in Nalli et al. (2013a), the magnitude of LWIR sun glint in the JAIVEx NAST-I data will have a systematic angular dependence. We accounted for this problem by calculating the reflected solar radiance as described below.
APPENDIX B
Mapping GEO Cloud/Aerosol Layer Data to Aircraft Sensor Footprints
The aerosol or cloud signal observed within the FOVs of an aircraft sensor (e.g., NAST-I) geolocated to the surface “footprint”
We first performed the mapping from
APPENDIX C
Estimating Cloud Aspect Ratios from Cloud Shadows
Figure C1 shows a zoom on an image of the broken FWC clouds observed during JAIVEx near the edge of the sun-glint region (the zoom is on the center image shown in Fig. 2), which is annotated to show estimates of the cloud horizontal dimensions versus the shadows they cast, these being visible over the sun-glint region. From geometric considerations discussed below, we consider only the shadows observed near the nadir angle of the observer (i.e., those near the center of the image), which are found near the edge of the sun-glint region owing to surface wave slopes oriented to reflect sun rays from the solar zenith angle to the observer. The observed cloud shadows are approximately in the range of 0.5–1.0 times the size of the cloud horizontal dimensions. We may then obtain the cloud aspect ratios
Here we restrict our consideration to the idealized cloud shapes considered previously by Nalli et al. (2012) based on the PCLoS cloud shape factors compiled by Taylor and Ellingson (2008), which include three basic forms—namely, isosceles trapezoids (including rectangles and triangles), spheroids (ellipsoids with equal horizontal semiaxes,
For simplicity, in the current work we assume clouds close enough to the surface to disregard Earth curvature (as would be the case for most opaque water-droplet clouds and certainly FWC); the observer is assumed to be near zenith, and both the sun and observer are considered far enough away such that rays may be assumed parallel. Figure C2 shows the two-dimensional (2D) geometry for an assumed hemispheroidal (semiellipsoidal) cloud shape, which also applies to the other idealized cloud shapes without loss of generality, hemispheroidal cloud shapes being considered good for representing FWC (as implied in the above paragraph). The figure specifically shows the semiellipsoidal cross section in the x–z plane with the solar azimuth oriented along the x axis. Given the geometrical optics limit valid for visible spectrum wavelengths, the width of a shadow cast by an opaque cloud is equivalent to the 2D cross section of the cloud perpendicular to the incoming solar rays; for a nonspherical cloud shape, the linear cross section will vary with the solar incidence angle. Nalli et al. (2012) derived equivalent expressions for calculating mean slant paths through semitransparent clouds that can be utilized for the present application—namely,
Thus, given Eq. (C3), along with either of Eq. (C4), (C6), or (C7), we can calculate the cloud shadow projected on Earth’s surface visible to an observer near zenith,
REFERENCES
Arndt, D. S., M. O. Baringer, and M. R. Johnson, 2010: State of the Climate in 2009. Bull. Amer. Meteor. Soc., 91, s1–s222, doi:10.1175/BAMS-91-7-StateoftheClimate.
Benner, T. C., and J. A. Curry, 1998: Characteristics of small tropical cumulus clouds and their impact on the environment. J. Geophys. Res., 103, 28 753–28 767, doi:10.1029/98JD02579.
Bréon, F. M., and N. Henriot, 2006: Spaceborne observations of ocean glint reflection and modeling of wave slope distributions. J. Geophys. Res., 111, C06005, doi:10.1029/2005JC003343.
Cayla, F.-R., 1993: IASI infrared interferometer for operations and research. High Spectral Resolution Infrared Remote Sensing for Earth’s Weather and Climate Studies, A. Chedin, M. T. Chahine, and N. A. Scott, Eds., NATO ASI Series, Vol. 19, Springer-Verlag, 9–19.
Chahine, M. T., and Coauthors, 2006: AIRS: Improving weather forecasting and providing new data on greenhouse gases. Bull. Amer. Meteor. Soc., 87, 911–926, doi:10.1175/BAMS-87-7-911.
Clough, S. A., M. W. Shephard, E. J. Mlawer, J. S. Delamere, M. J. Iacono, K. Cady-Pereira, S. Boukabara, and P. D. Brown, 2005: Atmospheric radiative transfer modeling: A summary of the AER codes. J. Quant. Spectrosc. Radiat. Transfer, 91, 233–244, doi:10.1016/j.jqsrt.2004.05.058.
Cox, C., and W. Munk, 1955: Some problems in optical oceanography. J. Mar. Res., 14, 63–78.
Cox, C., 1974: Refraction and reflection of light at the sea surface. Optical Aspects of Oceanography, N. Jerlov and E. Nielsen, Eds., Academic Press, 51–75.
Goldberg, M. D., H. Kilcoyne, H. Cikanek, and A. Mehta, 2013: Joint Polar Satellite System: The United States next generation civilian polar-orbiting environmental satellite system. J. Geophys. Res. Atmos., 118, 13 463–13 475, doi:10.1002/2013JD020389.
Goody, R. M., and Y. L. Yung, 1989: Atmospheric Radiation: Theoretical Basis. 2nd ed. Oxford University Press, 519 pp.
Hilton, F., and Coauthors, 2012: Hyperspectral Earth observation from IASI: Five years of accomplishments. Bull. Amer. Meteor. Soc., 93, 347–370, doi:10.1175/BAMS-D-11-00027.1.
Inman, R. L., 1969: Computation of temperature at the lifted condensation level. J. Appl. Meteor., 8, 155–158, doi:10.1175/1520-0450(1969)008<0155:COTATL>2.0.CO;2.
Jackson, C. R., and W. Alpers, 2010: The role of the critical angle in brightness reversals on sunglint images of the sea surface. J. Geophys. Res., 115, C09019, doi:10.1029/2009JC006037.
Kauth, R. J., and J. L. Penquite, 1967: The probability of clear lines of sight through a cloudy atmosphere. J. Appl. Meteor., 6, 1005–1017, doi:10.1175/1520-0450(1967)006<1005:TPOCLO>2.0.CO;2.
Kay, S., J. D. Hedley, and S. Lavender, 2009: Sun glint correction of high and low spatial resolution images of aquatic scenes: A review of methods for visible and near-infrared wavelengths. Remote Sens., 1, 697–730, doi:10.3390/rs1040697.
Knapp, K. R., T. H. Vonder Haar, and Y. J. Kaufman, 2002: Aerosol optical depth retrieval from GOES-8: Uncertainty study and retrieval validation over South America. J. Geophys. Res., 107, doi:10.1029/2001JD000505.
Larar, A. M., W. L. Smith, D. K. Zhou, X. Liu, H. Revercomb, J. P. Taylor, S. M. Newman, and P. Schlüssel, 2010: IASI spectral radiance validation inter-comparisons: Case study assessment from the JAIVEx field campaign. Atmos. Chem. Phys., 10, 411–430, doi:10.5194/acp-10-411-2010.
Lawrence, M. G., 2005: The relationship between relative humidity and the dewpoint temperature in moist air: A simple conversion and applications. Bull. Amer. Meteor. Soc., 86, 225–233, doi:10.1175/BAMS-86-2-225.
Liu, Q., and S. Boukabara, 2014: Community Radiative Transfer Model (CRTM) applications in supporting the Suomi National Polar-orbiting Partnership (SNPP) mission validation and verification. Remote Sens. Environ., 140, 744–754, doi:10.1016/j.rse.2013.10.011.
Lund, I. A., and M. D. Shanklin, 1972: Photogrammetrically determined cloud-free lines-of-sight through the atmosphere. J. Appl. Meteor., 11, 773–782, doi:10.1175/1520-0450(1972)011<0773:PDCFLO>2.0.CO;2.
Lund, I. A., and M. D. Shanklin, 1973: Universal methods for estimating probabilities of cloud-free lines-of-sight through the atmosphere. J. Appl. Meteor., 12, 28–35, doi:10.1175/1520-0450(1973)012<0028:UMFEPO>2.0.CO;2.
Maddy, E. S., and Coauthors, 2011: Using MetOp-A AVHRR clear-sky measurements to cloud-clear MetOp-A IASI column radiances. J. Atmos. Oceanic Technol., 28, 1104–1116, doi:10.1175/JTECH-D-10-05045.1.
Mandel, J., 1984: The Statistical Analysis of Experimental Data. Dover Books on Mathematics, Dover Publications, 432 pp.
Martin, S., 2004: An Introduction to Ocean Remote Sensing. 1st ed. Cambridge University Press, 426 pp.
Masuda, K., 2006: Infrared sea surface emissivity including multiple reflection effect for isotropic Gaussian slope distribution model. Remote Sens. Environ., 103, 488–496, doi:10.1016/j.rse.2006.04.011.
Merchant, C. J., A. R. Harris, E. Maturi, O. Embury, S. N. MacCallum, J. Mittaz, and C. P. Old, 2009: Sea surface temperature estimation from the Geostationary Operational Environmental Satellite-12 (GOES-12). J. Atmos. Oceanic Technol., 26, 570–581, doi:10.1175/2008JTECHO596.1.
Nalli, N. R., and L. L. Stowe, 2002: Aerosol correction for remotely sensed sea surface temperatures from the National Oceanic and Atmospheric Administration advanced very high resolution radiometer. J. Geophys. Res., 107, 3172, doi:10.1029/2001JC001162.
Nalli, N. R., and W. L. Smith, 2003: Retrieval of ocean and lake surface temperatures from hyperspectral radiance observations. J. Atmos. Oceanic Technol., 20, 1810–1825, doi:10.1175/1520-0426(2003)020<1810:ROOALS>2.0.CO;2.
Nalli, N. R., W. L. Smith, and B. Huang, 2001: Quasi-specular model for calculating the reflection of atmospheric emitted infrared radiation from a rough water surface. Appl. Opt., 40, 1343–1353, doi:10.1364/AO.40.001343.
Nalli, N. R., and Coauthors, 2006: Ship-based measurements for infrared sensor validation during Aerosol and Ocean Science Expedition 2004. J. Geophys. Res., 111, D09S04, doi:10.1029/2005JD006385.
Nalli, N. R., P. J. Minnett, E. Maddy, W. W. McMillan, and M. D. Goldberg, 2008a: Emissivity and reflection model for calculating unpolarized isotropic water surface leaving radiance in the infrared. 2: Validation using Fourier transform spectrometers. Appl. Opt., 47, 4649–4671, doi:10.1364/AO.47.004649.
Nalli, N. R., P. J. Minnett, and P. van Delst, 2008b: Emissivity and reflection model for calculating unpolarized isotropic water surface leaving radiance in the infrared. 1: Theoretical development and calculations. Appl. Opt., 47, 3701–3721, doi:10.1364/AO.47.003701.
Nalli, N. R., C. D. Barnet, E. S. Maddy, and A. Gambacorta, 2012: On the angular effect of residual clouds and aerosols in clear-sky infrared window radiance observations: Sensitivity analyses. J. Geophys. Res., 117, D12208, doi:10.1029/2012JD017667.
Nalli, N. R., C. D. Barnet, A. Gambacorta, E. S. Maddy, H. Xie, T. S. King, E. Joseph, and V. R. Morris, 2013a: On the angular effect of residual clouds and aerosols in clear-sky infrared window radiance observations 2. Satellite experimental analyses. J. Geophys. Res. Atmos., 118, 1420–1435, doi:10.1029/2012JD018260.
Nalli, N. R., and Coauthors, 2013b: Validation of satellite sounder environmental data records: Application to the Cross-track Infrared Microwave Sounder Suite. J. Geophys. Res. Atmos., 118, 13 628–13 643, doi:10.1002/2013JD020436.
Newman, S. M., and Coauthors, 2012: The Joint Airborne IASI Validation Experiment: An evaluation of instrument and algorithms. J. Quant. Spectrosc. Radiat. Transfer, 113, 1372–1390, doi:10.1016/j.jqsrt.2012.02.030.
NOAA/NESDIS, 2003: GOES level 3 6km near real time SST 1 hour. PO.DAAC, accessed 3 December 2010, doi:10.5067/GOES3-1HOUR.
Prados, A. I., S. Kondragunta, P. Ciren, and K. R. Knapp, 2007: GOES Aerosol/Smoke Product (GASP) over North America: Comparisons to AERONET and MODIS observations. J. Geophys. Res., 112, D15201, doi:10.1029/2006JD007968.
Remote Sensing Systems, 2008: GHRSST level 4 mw_ir_OI global foundation sea surface temperature analysis. PO.DAAC, accessed 13 April 2011, doi:10.5067/GHMWI-4FR01.
Smith, W. L., Sr., D. K. Zhou, A. M. Larar, S. A. Mango, H. B. Howell, R. O. Knuteson, H. E. Revercomb, and W. L. Smith Jr., 2005: The NPOESS Airborne Sounding Testbed Interferometer—Remotely sensed surface and atmospheric conditions during CLAMS. J. Atmos. Sci., 62, 1118–1134, doi:10.1175/JAS3384.1.
Smith, W. L., Sr., and Coauthors, 2009: Technical note: Evolution, current capabilities, and future advance in satellite nadir viewing ultra-spectral IR sounding of the lower atmosphere. Atmos. Chem. Phys., 9, 5563–5574, doi:10.5194/acp-9-5563-2009.
Sokolik, I. N., 2002: The spectral radiative signature of wind-blown mineral dust: Implications for remote sensing in the thermal IR region. Geophys. Res. Lett., 29, 2154, doi:10.1029/2002GL015910.
Stephens, G. L., 1994: Remote Sensing of the Lower Atmosphere: An Introduction. Oxford University Press, 523 pp.
Stull, R. B., 1985: A fair-weather cumulus cloud classification scheme for mixed-layer studies. J. Climate Appl. Meteor., 24, 49–56, doi:10.1175/1520-0450(1985)024<0049:AFWCCC>2.0.CO;2.
Stull, R. B., 1988: An Introduction to Boundary Layer Meteorology. Kluwer Academic Publishers, 666 pp.
Sun, B., A. Reale, D. J. Seidel, and D. C. Hunt, 2010: Comparing radiosonde and cosmic atmospheric profile data to quantify differences among radiosonde types and the effects of imperfect collocation on comparison statistics. J. Geophys. Res., 115, D23104, doi:10.1029/2010JD014457.
Susskind, J., C. D. Barnet, and J. M. Blaisdell, 2003: Retrieval of atmospheric and surface parameters from AIRS/AMSU/HSB data in the presence of clouds. IEEE Trans. Geosci. Remote Sens., 41, 390–409, doi:10.1109/TGRS.2002.808236.
Taylor, P. C., and R. G. Ellingson, 2008: A study of the probability of clear line of sight through single-layer cumulus cloud fields in the tropical western Pacific. J. Atmos. Sci., 65, 3497–3512, doi:10.1175/2008JAS2620.1.
Thomas, G. E., and K. Stamnes, 1999: Radiative Transfer in the Atmosphere and Ocean. Cambridge Atmospheric and Space Science, Cambridge University Press, 517 pp.
Thomas, S. M., A. K. Heidinger, and M. J. Pavolonis, 2004: Comparison of NOAA’s operational AVHRR-derived cloud amount to other satellite-derived cloud climatologies. J. Climate, 17, 4805–4822, doi:10.1175/JCLI-3242.1.
Weng, F., X. Zou, X. Wang, S. Yang, and M. D. Goldberg, 2012: Introduction to Suomi national polar-orbiting partnership advanced technology microwave sounder for numerical weather prediction and tropical cyclone applications. J. Geophys. Res., 117, D19112, doi:10.1029/2012JD018144.
WMO, 1956: International cloud Atlas: Abridged Atlas. World Meteorological Organization, 62 pp.
WMO, 1987: International Cloud Atlas. Vol. II. World Meteorological Organization WMO-407, 212 pp. [Available online at http://library.wmo.int/pmb_ged/wmo_407_en-v2.pdf.]
Wong, S., E. J. Fetzer, M. Schreier, G. Manipon, E. F. Fishbein, B. H. Kahn, Q. Yue, and F. W. Irion, 2015: Cloud-induced uncertainties in AIRS and ECMWF temperature and specific humidity. J. Geophys. Res. Atmos., 120, 1880–1901, doi:10.1002/2014JD022440.
Wu, X., and W. L. Smith, 1997: Emissivity of rough sea surface for 8–13μm: Modeling and validation. Appl. Opt., 36, 2609–2619, doi:10.1364/AO.36.002609.
Yao, Z., J. Li, and J. Li, 2012: Sunglint impact on atmospheric soundings from hyperspectral resolution infrared radiances. Adv. Atmos. Sci., 29, 455–463, doi:10.1007/s00376-011-1013-8.
Závody, A. M., and A. R. Birks, 2004: Sun glint contamination in ATSR-2 data: Comparison of observations and values calculated from the measured 1.6-μm reflectivities. J. Atmos. Oceanic Technol., 21, 787–798, doi:10.1175/1520-0426(2004)021<0787:SGCIAD>2.0.CO;2.
Zhang, H., and M. Wang, 2010: Evaluation of sun glint models using MODIS measurements. J. Quant. Spectrosc. Radiat. Transfer, 111, 492–506, doi:10.1016/j.jqsrt.2009.10.001.
Zhao, T. X.-P., P. K. Chan, and A. K. Heidinger, 2013: A global survey of the effect of cloud contamination on the aerosol optical thickness and its long-term trend derived from operational AVHRR satellite observations. J. Geophys. Res. Atmos., 118, 2849–2857, doi:10.1002/jgrd.50278.
As in our previous papers, we use the term “concave up” to describe an increasing positive bias in calc − obs with
The CRTM was developed at the Joint Center for Satellite Data Assimilation (JCSDA) in the United States in support of satellite radiance assimilation for numerical weather prediction (NWP), satellite product retrievals, and radiance validation for satellite programs including JPSS (Liu and Boukabara 2014).
We also performed calculations for additional LWIR microwindows, namely