投稿指南
来稿应自觉遵守国家有关著作权法律法规,不得侵犯他人版权或其他权利,如果出现问题作者文责自负,而且本刊将依法追究侵权行为给本刊造成的损失责任。本刊对录用稿有修改、删节权。经本刊通知进行修改的稿件或被采用的稿件,作者必须保证本刊的独立发表权。 一、投稿方式: 1、 请从 我刊官网 直接投稿 。 2、 请 从我编辑部编辑的推广链接进入我刊投审稿系统进行投稿。 二、稿件著作权: 1、 投稿人保证其向我刊所投之作品是其本人或与他人合作创作之成果,或对所投作品拥有合法的著作权,无第三人对其作品提出可成立之权利主张。 2、 投稿人保证向我刊所投之稿件,尚未在任何媒体上发表。 3、 投稿人保证其作品不含有违反宪法、法律及损害社会公共利益之内容。 4、 投稿人向我刊所投之作品不得同时向第三方投送,即不允许一稿多投。 5、 投稿人授予我刊享有作品专有使用权的方式包括但不限于:通过网络向公众传播、复制、摘编、表演、播放、展览、发行、摄制电影、电视、录像制品、录制录音制品、制作数字化制品、改编、翻译、注释、编辑,以及出版、许可其他媒体、网站及单位转载、摘编、播放、录制、翻译、注释、编辑、改编、摄制。 6、 第5条所述之网络是指通过我刊官网。 7、 投稿人委托我刊声明,未经我方许可,任何网站、媒体、组织不得转载、摘编其作品。

Improving Representation of Tropical Cloud (6)

来源:热带地理 【在线投稿】 栏目:期刊导读 时间:2020-12-24
作者:网站采编
关键词:
摘要:The results of this study do, however, provide modelers with a simple and effective approach to more realistically address cloud overlap in tropical regions. This approach acts as another, although no

The results of this study do, however, provide modelers with a simple and effective approach to more realistically address cloud overlap in tropical regions. This approach acts as another, although not perfect, constraint of cloud-radiation interactions, an aspect of GCMs that contributes greatly to uncertainties in simulations. The form of the Lcf-convection relationship is open to further improvement and the application of this method to extended CRM and observational datasets (e.g., datasets from the multi-sensor A-Train constellation) will provide essential evidence of the Lcf-convection relationship from the points of view of both simulation and observation,and produce more robust treatments of Lcf in future studies.

Fig. 13. (a) Longwave and (c) shortwave heating rates for the three cloud overlap treatments (MRO, G2KM, and PARA) and REF over the tropical ascending area and the differences between the three overlap treatments and REF (b and d, respectively).

*Corresponding author: .

?The Chinese Meteorological Society and Springer-Verlag Berlin Heidelberg 2018

1. IntroductionThe simulation of clouds has been a major source of uncertainty in projections of future climate using general circulation models (GCMs) (Stephens, 2005; Li et al.,2009; Bony et al., 2015). One limitation of cloud simulation is the coarse spatial resolution of GCMs (tens of kilometers to 100-200 km), which leaves clouds smaller than grid size unresolved (Barker et al., 2003; Randall et al., 2003). Consequently, clouds in GCMs usually cover only part of a grid layer and the overlap of fractional clouds in the vertical layers has to be addressed artificially in radiation calculations by imposing overlap assumptions (Tompkins and Di Giuseppe, 2015; Zhang and Jing, 2016). For a given vertical distribution of cloud fractions, the assumption of cloud overlap determines the total cloud cover or total cloud fraction (Ctot), which has a considerable influence on solar and terrestrial radiative transfer (Wang et al., 2016).The cloud overlap assumption most widely used in recent decades is the maximum random overlap (MRO) assumption (Morcrette and Fouquart, 1986; Tian and Curry, 1989), in which clouds within layers that are vertically continuous are assumed to have a maximum overlap, whereas those that are separated by cloud-free layers are considered to overlap randomly. Such treatment is insufficient to represent the realistic features of cloud overlap as observed by ground-based radar (Hogan and Illingworth, 2000; Mace and Benson-Troth, 2002) and depends largely on the vertical resolution of the host model (Bergman and Rasch, 2002).In contrast with the simple, crude cloud overlap treatments such as the MRO assumption, Liang and Wang(1997) were among the first to explicitly depict the subgrid distribution of clouds of distinct physical types and to apply different treatments of vertical overlap for different types of clouds. This sophisticated, physically based treatment of cloud overlap (termed “mosaic”) has been demonstrated to improve cloud radiative forcing and radiative heating in both cloud-resolving model(CRM) domains (Liang and Wu, 2005; Wu and Liang,2005a, b) and climate simulations (Zhang F. et al., 2013).Another ingenious approach is the analytical representation of cloud overlap proposed by Hogan and Illingworth (2000) and Mace and Benson-Troth (2002) based on radar observations. This method is called general overlap (GenO). In GenO, for two layers of clouds at heights ofZkandZlwith cloud fractions ofCkandCl, respectively,Ctotis defined asLcfis the decorrelation length (in km) representing theEqs. (1) and (2), the extent of overlap degrades exponentially from maximum overlap to random overlap as the vertical separation of clouds increases. This relationship of decreasing overlap with increasing distance has been reported from both radar observations and simulations by CRMs (Oreopoulos and Khairoutdinov, 2003). The merits of GenO are two-fold: (1) it realistically depicts the distance-related feature of cloud overlap and (2) it is independent of the vertical resolution of the host model and thus more widely applicable among models with various vertical GenO, the extent of cloud overlap is determined byLcf. For given fractional clouds in a vertical column, the use of larger values ofLcfresults in smaller values ofCtot(prone to maximum overlap) and smaller values ofLcfresult in larger values ofCtot(prone to random overlap).The parameterLcfis highly variable both spatially and temporally because of variations in the shapes and formation processes of clouds. Therefore, when applying GenO, one challenge is to determine an optimum value ofLcffor each GCM grid point. Various attempts have been made to obtain detailed information aboutLcf(e.g.,Di Giuseppe, 2005; Kato et al., 2010; Shonk et al., 2010;Oreopoulos et al., 2012; Peng et al., 2013; Zhang H. et al., 2013). It has been demonstrated thatLcfis related to the cloud type and atmospheric dynamics (Naud et al.,2008; Di Giuseppe and Tompkins, 2015; Li et al., 2015)and that it has a global median value of approximately 2 km (Barker, 2008). Simplified expressions have also been extracted to representLcfin GCMs, either as a function of latitude and/or season without interannual variations (Shonk et al., 2010, Oreopoulos et al., 2012; Jing et al., 2016) or as a function of cloud type, which is affected by the limited cloud classification schemes of the host models (Zhang et al., 2014). These approaches either lack a direct link betweenLcfand the instant largescale meteorological conditions that foster the clouds or address the dynamic (e.g., wind shear) impact on cloud overlap over the globe without considering the very different circulation conditions in different formation and evolution of clouds are essentially associated with large-scale circulation (Bony et al.,1997). Therefore one physically robust approach to describeLcfis to establish a direct connection betweenLcfand large-scale circulation conditions. CRMs, because of their ability to simulate cloud micro- and macro-physical structures as well as meteorological conditions in detail,have long been used as a tool to explore cloud physics and to obtain parameterizations applicable in GCMs(GEWEX Cloud System Science Team, 1993; Randall et al., 2003; Wu and Li, 2008). This study uses simulation results from a global CRM to explore the relationship betweenLcfand atmospheric circulation. Unlike previous studies that attempted to explore such a relationship over the whole globe, we focus on the tropical region and vertical motion only, considering that there are large uncertainties in cloud radiative forcing due to vertical overlap treatment in the intertropical convective zone (ITCZ)(Barker and R?is?nen, 2005; Zhang and Jing, 2010;Lauer and Hamilton, 2013) and that the formation and maintenance of clouds in this particular region are closely related to vertical convection. We will attempt to establish a statistical, mathematical description of theLcf-convection connection, which is a novel application in GCMs, and then evaluate its effectiveness in improving the GCM-scale cloud cover and radiation CRM data and the method used to deriveLcfare described in Section 2. Section 3 presents the analysis ofLcfversus atmospheric convection and evaluates the established representation ofLcf. The discussion and conclusions are presented in Section 4.2. Data and dataThe global CRM data are from a simulation of the Nonhydrostatic Icosahedral Atmospheric Model (NICAM)developed at Japan Agency for Marine-Earth Science and Technology and the University of Tokyo (Tomita and Satoh, 2004; Satoh et al., 2008, 2014). The cloud microphysics scheme of Grabowski (1998) is adopted and no convective scheme is used. This scheme is simpler than other cloud microphysics schemes, but convective circulation is explicitly calculated so that the associations between convection and large-scale atmospheric states are consistently represented. The boundary layer scheme with moist processes is implemented (Nakanishi and Niino, 2006). The vertical resolution of the model is 40 levels with Lorenz grids (Satoh et al., 2008, 2014)stretching from the surface up to about 40 km. The vertical interval of the half-level (ΔZ) increases with height from about 170 m at the bottom to about 3 km near the top (ΔZ< 1 km below 10 km).NICAM can be run using different horizontal resolutions depending on the grid division level used. In this study, data from a simulation using a grid division level of 11 (corresponding to a grid size of about 3.5 km)(Miura et al., 2007) are adopted; the cloud characteristics of this dataset have been extensively analyzed by Inoue et al. (2008, 2010), Masunaga et al. (2008), Sato et al. (2009), and Satoh et al. (2010). The simulation was started from 0000 UTC 25 December 2006 and integrated for 7 days. The results were stored as instantaneous snapshots at 0000 UTC on each day. Figures 1 and 2 show the simulated total cloud fraction and the zonal mean vertical cloud fraction profile, respectively, compared with those from the 2B-GEOPROF product of CloudSat observations (Marchand et al., 2008) during the same period. The NICAM simulation captures both the geographical distribution of cloud systems in the deep convective regions (Fig. 1) and the vertical cloud profiles in the tropics (Fig. 2). The cloud top height of NICAM in Fig. 2 is slightly higher than the CloudSat observations,which is because CloudSat does not detect optically thin clouds in the topmost layers (Stephens et al., 2008).The variables used in this study are the mass mixing ratios of liquid and ice water condensates (qcandqi, respectively), in-cloud precipitation (qr) and snow (qs), and vertical velocity (w). of LcfTo obtainLcfat the resolution of the GCM and to establish the GCM-oriented relationship betweenLcfand atmospheric convection, the original NICAM output is averaged based on a 2.8° × 2.8° (latitude × longitude)grid division, which is close to a T42 GCM grid mean fields are then used to obtainLcfusing a stochastic cloud generator (R?is?nen et al., 2004) as ) Diagnose the occurrence of cloud fromqc,qi,qr,andqs. Ifqc+qi+qr+qs> 0.01 g kg-1in a CRM grid cell, then this grid cell is regarded as cloudy (i.e., cloud fraction = 1), otherwise the grid cell is regarded as clear(i.e., cloud fraction = 0). This criterion was also used in Grabowski (1998).Fig. 1. Comparison of total cloud fraction (Ctot) between (a) CloudSat observations and (b) the NICAM simulation during the 7-day period starting from 0000 UTC 25 December 2006.2) Average the original CRM fields to the 2.8° × 2.8°grid division. The cloud fraction and the vertical velocity at 500 hPa (w500) are averaged within each GCM grid. These yield vertical distributions of cloud fractions andw500for each GCM grid. In addition, the vertically projected cloud fraction (i.e.,Ctot) for each GCM grid is derived by dividing the number of cloudy CRM columns by the total number of CRM columns within the grid.3) ObtainLcf. The vertical cloud fraction profile andCtotin each GCM grid are supplied to the stochastic cloud generator with GenO incorporated to defined as the value that gives the sameCtotas the original CRM cloud field when used in Eqs. (1) and (2). To demonstrate the effectiveness of using this procedure to capture the cloud overlap characteristics, a GCM grid with a typical cloud profile often seen in the tropical deep convective region is chosen as an example and the generated and original cloud structures are properties examined include the vertical cloud fraction profile, the downward cumulative cloud fraction,and the cloud fraction exposed to space at different heights (Fig. 3). It is suggested that, by applying the achievedLcfin GenO, the generated cloud structures(dotted lines) resemble those of the original CRM field(solid lines). The characteristics of clouds shown in Fig. 3 are important to both solar reflectance and upward longwave emissivity. Therefore, the cloud structures generated by GenO with an accurate value ofLcfpotentially facilitate the computation of radiation representation of LcfThe vertical velocity in the mid-troposphere (w500) has been shown to be a representative indicator of tropical convection and cloud radiative forcing (Ichikawa et al.,2012). The relationship betweenLcfandw500is assessed in this subsection. As the overlap of clouds with very large or very small cloud fractions is of secondary importance forCtotand radiation calculations, grids with maximum layer cloud fractions > 0.9 or < 0.1 are 2. Comparison of zonal mean vertical cloud fraction profile between (a) CloudSat observations and (b) the NICAM simulation during the 7-day period starting from 0000 UTC 25 December 3. Comparison of cloud features in a selected domain for the original CRM cloud field and that generated based on GenO with the true value ofLcf. (a) Cloud fraction at different altitudes, (b) downward cumulative cloud fraction, and (c) cloud fraction exposed to space at different 4 shows the distributions ofLcfandw500for each model day. Domains with a large and positivew500(e.g., the western Pacific and South America) mostly have a large value ofLcf(typically 4-7 km and up to 10 km in extreme cases). By contrast, in domains with a small or negativew500,Lcfis mostly about or below 2 km. Figure 5 shows the pattern correlation between the geographical distributions ofLcfandw500for each snapshot. The pattern correlation stays at a moderate, but notable and constant level (from 0.61 to 0.66), implying a physically close association on pattern similarity and the clear difference between areas of ascent (w500> 0) and descent (w500< 0),the relationship betweenLcfandw500will be explored separately for areas of ascent and 4. Distributions ofw500(left-hand panels) andLcf(right-hand panels) in the tropics for each model day. Masked grids are those with maximum cloud fraction in the vertical direction of > 0.9 or < 0.1 and were thus eliminated from the 6 shows the statistics forLcfandw500and their relationships in the ascending areas. A total of 4926 samples were used to derive these statistics. The gray lines in Fig. 6a show the median and the first and third quartiles ofLcffor each bin ofLcf. These lines clearly show a positive relationship betweenLcfandw500for areas of ascent: whenw500< 0.02 m s-1, the most frequent occurrence ofLcfis around 2 km; asw500increases to 0.08 m s-1, the corresponding value ofLcfincreases to as much as 6-8 km, although the occurrence probability ofw500> 0.08 m s-1is very small (Fig. 6b). Linear regression is conducted (as shown by the black solid line and the regression equation in Fig. 6a) forLcfas a function ofw500. The regression line captures very well the relationship illustrated by the yellow shaded area in Fig. 6a. The 95% confidence interval (blue dotted lines)and 95% prediction interval (red dashed lines) for this regression are also shown in Fig. 6a. The small 95% confidence interval (the mean ofLcfis 95% likely to fall into this interval for a given value ofw500) suggests that the linear regression is an excellent representation of the average relationship betweenLcfandw500; however, it should also be noted that the dispersion of theLcfandw500relationship is relatively large, as indicated by the 95% prediction interval (an individualLcffor a givenw500falls into this interval with probability of 95%). The large dispersion of theLcfandw500relationship may stem primarily from the effect of other meteorological conditions on cloud overlap, which cannot be captured by a simple linear regression. Nevertheless, these results demonstrate that, in the statistical sense,Lcfandw500are linearly related in the ascending regions of the analysis in the following sections shows that addressing this statistical relationship will benefit the calculation of the cloud fraction and radiation same analysis for areas of descent shows thatLcfchanges very little withw500(data not shown) and that a value of 2 km is a good representation for these areas regardless of the specificw500. Consequently,Lcfcan be approximated in the region 30°S-30°N depending onw500as:Fig. 5. Pattern correlation betweenLcfandw500for each snapshot shown in Fig. 4.It should be noted that although GenO is independent of the vertical resolution, Eq. (3) may be affected by the vertical resolution of the NICAM data. Thus we should be cautious when applying Eq. (3) in a model with a notably different vertical resolution from that of NICAM,especially in the in terms of cloud fractionFig. 6. Statistics ofLcfandw500constructed for the whole 7-day period. (a) The median (gray dashed line) and first and third quartiles (gray solid lines) ofLcffor eachw500bin (bin width 0.01 m s-1) and the linear regression ofLcfas a function ofw500(black solid line) (the linear regression equation and correlation coefficient are also shown). The yellow shaded area is the interquartile range ofLcf; the blue dotted line and red dashed line are the 95% confidence interval and 95% prediction interval of the regression, respectively. (b) Probability distribution function better representation ofLcfis primarily expected to be capable of generating a realisticCtotfor given cloud fraction profiles. We evaluate here the generatedCtotby using theLcfparameterized from Eq. (3) (denoted as PARA). Two other cloud overlap treatments are also included to compare with PARA: the traditional MRO assumption, which has been widely used not only in GCMs(R?is?nen, 1998; Collins, 2001), but also in model assessment tools such as the CFMIP Observation Simulator Package (Bodas-Salcedo et al., 2011); and the simplified GenO using a constantLcf(2 km) (denoted as G2KM), which applies the framework of GenO, but with a constantLcfsuggested as a global mean value (Barker,2008). These cloud overlap treatments are used to generate sub-grid cloud fields from the cloud fraction profiles of the CRM and the values ofCtotof the generated fields are then compared with those of the original CRM 7 shows the 7-day mean biases ofCtotgenerated from the three cloud overlap representations relative to the reference value. It can be seen that the widely used MRO assumption remarkably underestimatesCtotin most regions. This is because there are few clear layers to separate the cloud layers in the vertical columns and therefore there is a maximum overlap of clouds in most demonstrates one drawback of the MRO technique:it depends greatly on the vertical resolution of the host model and is therefore usually non-equivalent among use of GenO significantly reduces the negative biases compared with the MRO assumption, even when a constantLcfof 2 km is used (Fig. 7b). However, G2KM leads to notableCtotbiases in the ITCZ, especially in the western Pacific and Amazon regions, where clouds are generally more vertically organized because of systematic large-scale upward motion. When the dynamic representation ofLcffor regions of ascent is used, the cloud fraction errors in the ITCZ are remarkably reduced (Fig. 7c). in terms of radiation effectsRadiation calculations are performed for the generated cloud fields and the original CRM cloud fields in the evaluation using the BCC-RAD correlated-k distribution radiation model (Zhang et al., 2003, 2006a, b ),which has been implemented in the GCM of the Beijing Climate Center (BCC_) (Zhang et al., 2014).For both the generated and original cloud fields, the cloud water/ice content in each GCM grid is obtained as a grid mean value to eliminate the effect of the horizontal distributions of cloud water/ice. Because only the cloud and dynamic output from NICAM are stored (due to the large amount of data), we apply in the radiation calculation the atmospheric pressures, temperatures, and gas concentrations from the tropical atmosphere of the US Air Force Geophysics Laboratory atmospheric models(Anderson et al., 1986) and the broadband surface albedos for the ocean (0.08, Jin et al., 2004) and solid surfaces (0.28, Liang, 2001), 8 shows the biases in the generated net longwave (LWTOA) and shortwave (SWTOA) radiative fluxes at the top of the atmosphere relative to the CRM results. The largest radiation biases occur in the ITCZ,where the largestCtotbiases are also seen. The MRO assumption shows significant negative biases for LWTOA and positive biases for SWTOA (mostly > 10 and > 25 W m-2, respectively) (Figs. 8a, b) due to the underestimation ofCtot. Figures 8c and 8d show that the use of GenO with a constantLcfof 2 km reduces the biases in subtropical regions, but also introduces notable biases in the ITCZ with opposite signs to the biases of the MRO assumption. The dynamic representation ofLcflargely reduces the negative biases in the ITCZ (Figs. 8e, f), with considerably fewer regions having absolute biases > 10 and > 25 W m-2for LWTOA and SWTOA, 9 shows the biases in the net longwave (LWSFC) and shortwave (SWSFC) radiative fluxes at the surface. By comparing Fig. 9 with Fig. 8, it is seen that the main features of the biases at the surface are similar to those at the TOA, except that G2KM and PARA have much smaller (larger) biases in LWSFC over (outside)the ITCZ. PARA has the smallest errors among the three treatments of cloud overlap over the ITCZ for both LWSFC and SWSFC. It is therefore evident that the dynamic representation ofLcfyields the best spatial patterns of the radiation 7. Biases of generatedCtotusing (a) the MRO assumption, (b)GenO with universalLcf= 2 km (G2KM), and (c) GenO with dynamic representation ofLcf(PARA) relative to the trueCtotfrom the CRM(REF). Contour lines are shown for . 8. Biases of generated LWTOA (left-hand panels) and SWTOA (right-hand panels) using (a, b) the MRO assumption, (c, d) GenO withLcf= 2 km universally, and (e, f) GenO with dynamic representation ofLcf, relative to those calculated directly from the CRM fields. The downward direction is defined as positive. The contour lines for ±10 and ±25 W m-2are shown for LWTOA and SWTOA, 9. As in Fig. 8, but for (a, c, e) LWSFC and (b, d, f) 10 and 11 are scatter diagrams comparing the generated and reference radiation fields at the TOA and surface, respectively. Systematically negative biases in LWTOA and positive biases in SWTOA are shown for the MRO assumption (Figs. 10a, d) and the opposite for G2KM (Figs. 10b, e). PARA shows little systematic bias in LWTOA and SWTOA—that is, the points are distributed more symmetrically around the reference lines(Figs. 10c, f). Similar features are also shown at the surface (Fig. 11), except that G2KM and PARA resemble each other for the LWSFC (Figs. 11b, c), consistent with their similarity in Fig. 9.Figure 12 shows the tropical-averaged radiation biases for all areas (solid fill) and areas of ascent only(hatched fill). As expected, the MRO assumption shows the largest errors relative to the reference, especially for the shortwave fluxes (about 16 W m-2at both the TOA and the surface). G2KM performs much better than the MRO assumption, with absolute errors of approximately 2 W m-2for LWTOA and SWTOA. PARA reduces the errors more significantly, especially for shortwave fluxes. The error in LWSFC of PARA is more negative than that of G2KM (Fig. 12c); this is because PARA reduces the positive errors in the ITCZ (as shown in Fig. 9)that compensate for negative errors in other regions. It should be stressed that, for most of these variables,PARA reduces the all-area mean errors as effectively as it does in the areas of ascent only, implying that unrealistic cloud overlap treatment in areas of ascent is a major source of radiation error in tropical areas. Considering that these areas typically have large radiation biases in GCM simulations (Lauer and Hamilton, 2013), the introduction of PARA-like overlap treatment could possibly reduce the uncertainty in tropical radiation treatment of vertical cloud overlap also influences the radiative heating rate in the atmosphere, a property important for atmospheric stability and 13 compares the effects of different overlap treatments on the ascending area mean longwave (LWHTR)and shortwave (SWHTR) heating rates. The LWHTRs of different overlap treatments are similar to each other in the middle to higher troposphere, but near the surface the MRO assumption shows remarkable longwave cooling(Fig. 13b) and G2KM and PARA show overestimated longwave heating. This is probably due to the underestimated (overestimated) cloud fraction of the MRO assumption (G2KM and PARA) over these regions (as shown in Fig. 7). For the SWHTR, the MRO assumption underestimates heating around altitudes of 2 and 10 km,whereas G2KM overestimates heating around 10 km height and underestimates heating in the lower troposphere. PARA has a similar SWHTR to G2KM, but the bias in the upper troposphere is remarkably reduced and the bias in the lower troposphere is also reduced to some extent. These changes in the radiative heating rates caused by applying PARA, when coupled with a circulation model, will exert an influence on atmospheric stability in the vertical direction and consequently change the dynamic circulation. These aspects will be explored in future 10. Comparisons between LWTOA and SWTOA calculated from generated cloud fields with different cloud overlap treatments (MRO,G2KM, and PARA) and the values calculated directly from the CRM fields (REF) for the tropical areas.4. Discussion and conclusionsThe treatment of cloud overlap plays a crucial part in radiation calculations in GCMs. However, it is difficult to achieve a unified description of this highly variable(both spatially and temporally) property of clouds. By using the simulation of a CRM isolated to the tropical region, a statistical relationship between cloud overlap and convective strength was found and a dynamic representation ofLcf(which determines the extent of cloud overlap)was established. A simple linear regression ofLcfas a function of vertical velocity in the mid-troposphere is capable of partly capturing the cloud overlap dynamic connection and thus remarkably reduces biases in the cloud fraction and radiation fields in tropical convective regions. These regions are where major cloud fraction and radiation biases exist and therefore a reduction in these biases would make a significant contribution to the reduction of the overall bias in the 11. As in Fig. 10, but for (a, b, c) LWSFC and (d, e, f) 12. Tropical mean biases in (a) LWTOA, (b) SWTOA, (c) LWSFC, and (d) SWSFC for generated clouds fields with different cloud overlap treatments (MRO, G2KM, and PARA) compared with those for the CRM fields (REF) over all areas (solid fill) and areas of ascent only(hatched fill).In spite of the improvement in cloud fraction and radiation fields by implementing the convection-dependent overlap treatment, the results rely on the fidelity of the intrinsic physical and dynamic properties of the overlap characteristics also depend on other meteorological conditions in addition to convection, such as wind shear and atmospheric instability (Naud et al.,2008; Di Giuseppe and Tompkins, 2015), even in the tropical deep convection region. Thus the method derived here should not be regarded as a perfect deterministic relationship between cloud overlap and convection,but an empirical, statistical approximation of the contribution of convection to cloud overlap. Other meteorological conditions are virtually ignored in the overlap treatment in this study. The treatment of cloud overlap in extratropical areas was not addressed in this work because the overlap of clouds in such areas relates to large-scale meteorological conditions in a more complex manner than in the tropics. Another aspect of sub-grid cloud structures, i.e., the horizontally inhomogeneous distribution of cloud water, is also highly important in radiation calculations (Barker et al., 1996; Wu and Liang, 2005a);this aspect was excluded in the overlap treatment of this results of this study do, however, provide modelers with a simple and effective approach to more realistically address cloud overlap in tropical regions. This approach acts as another, although not perfect, constraint of cloud-radiation interactions, an aspect of GCMs that contributes greatly to uncertainties in simulations. The form of theLcf-convection relationship is open to further improvement and the application of this method to extended CRM and observational datasets (e.g., datasets from the multi-sensor A-Train constellation) will provide essential evidence of theLcf-convection relationship from the points of view of both simulation and observation,and produce more robust treatments ofLcfin future 13. (a) Longwave and (c) shortwave heating rates for the three cloud overlap treatments (MRO, G2KM, and PARA) and REF over the tropical ascending area and the differences between the three overlap treatments and REF (b and d, respectively).REFERENCESAnderson, G. P., S. A. Clough, F. X. Kneizys, et al., 1986: AFGL atmospheric constituent profiles (0.120 km). AFGL , AFGL-TR-86-0110, Bedford, MA, Air Force , , H. W., 2008: Representing cloud overlap with an effective decorrelation length: An assessment using CloudSat and CALIPSO data.J. Geophys. Res., 113, D, doi: 10.1029/, H. W., and P. R?is?nen, 2005: Radiative sensitivities for cloud structural properties that are unresolved by conventional J. Roy. Meteor. Soc., 131, 3103-3122, doi:10.1256/, H. W., B. A. Wiellicki, and L. Parker, 1996: A parameterization for computing grid-averaged solar fluxes for inhomogeneous marine boundary layer clouds. Part II: Validation using satellite data.J. Atmos. Sci., 53, 2304-2316, doi: 10.1175/1520-0469(1996)053<2304:APFCGA>2.0.CO;2.Barker, H. W., G. L. Stephens, P. Partain, et al., 2003: Assessing 1D atmospheric solar radiative transfer models: Interpretation and handling of unresolved Climate, 16,2676-2699, doi: 10.1175/1520-0442(2003)016<2676:ADASR T>2.0.CO;, J. W., and P. J. Rasch, 2002: Parameterizing vertically coherent cloud Atmos. Sci., 59, 2165-2182,doi: 10.1175/1520-0469(2002)059<2165:PVCCD>2.0.CO;, A., M. J. Webb, S. Bony, et al., 2011: COSP:Satellite simulation software for model Soc., 92, 1023-1043, doi: 10.1175/, S., K.-M. Lau, and Y. C. Sud, 1997: Sea surface temperature and large-scale circulation influences on tropical greenhouse effect and cloud radiative Climate, 10,2055-2077, doi: 10.1175/1520-0442(1997)010<2055:SSTAL S>2.0.CO;2.Bony, S., B. Stevens, D. M. W. Frierson, et al., 2015: Clouds, circulation and climate Geosci., 8, 261-268,doi: 10.1038/, W. D., 2001: Parameterization of generalized cloud overlap for radiative calculations in general circulation Sci., 58, 3224-3242, doi: 10.1175/1520-0469(2001)058<3224:POGCOF>2.0.CO;2.Di Giuseppe, F., 2005: Sensitivity of one-dimensional radiative biases to vertical cloud-structure assumptions: Validation with aircraft J. Roy. Meteor. Soc., 131, 1655-1676,doi: 10.1256/ Giuseppe, F., and A. M. Tompkins, 2015: Generalizing cloud overlap treatment to include the effect of wind shear.J. Atmos. Sci., 72, 2865-2876, doi: 10.1175/ Cloud System Science Team, 1993: The GEWEX cloud system study (GCSS).Bull. Amer. Meteor. Soc., 74, 387-400,doi: 10.1175/1520-0477(1993)074<0387:TGCSS>2.0.CO;, W. W., 1998: Toward cloud resolving modeling of large-scale tropical circulations: A simple cloud microphysics Atmos. Sci., 55, 3283-3298, doi: 10.11 75/1520-0469(1998)055<3283:TCRMOL>2.0.CO;2.Hogan, R. J., and A. J. Illingworth, 2000: Deriving cloud overlap statistics from J. Roy. Meteor. Soc., 126,2903-2909, doi: 10.1002/, H., H. Masunaga, Y. Tsushima, et al., 2012: Reproducibility by climate models of cloud radiative forcing associated with tropical Climate, 25, 1247-1262, doi:10.1175/, T., M. Satoh, H. Miura, et al., 2008: Characteristics of cloud size of deep convection simulated by a global cloud resolving model over the western tropical Meteor. Soc. Japan, 86A, 1-15, doi: 10.2151/, T., M. Satoh, Y. Hagihara, et al., 2010: Comparison of high-level clouds represented in a global cloud system-resolving model with CALIPSO/CloudSat and geostationary satellite Geophys. Res., 115, D00H22, doi:10.1029/, Z. H., T. P. Charlock, W. L. Jr. Smith, et al., 2004: A parameterization of ocean surface Res. Lett., 31,L, doi: 10.1029/, X. W., H. Zhang, J. Peng, et al., 2016: Cloud overlapping parameter obtained from CloudSat/CALIPSO dataset and its application in AGCM with McICA Res., 170,52-65, doi: 10.1016/, S., S. Sun-Mack, M. F. Miller, et al., 2010: Relationships among cloud occurrence frequency, overlap, and effective thickness derived from CALIPSO and CloudSat merged cloud vertical Geophys. Res., 115, D00H28, doi:10.1029/, A., and K. Hamilton, 2013: Simulating clouds with global climate models: A comparison of CMIP5 results with CMIP3 and satellite data.J. Climate, 26, 3823-3845, doi: 10.1175/, J., J. Huang, K. Stamnes, et al., 2015: A global survey of cloud overlap based on CALIPSO and CloudSat Chem. Phys., 15, 519-536, doi: 10.5194/, J. D., Y. M. Liu, and G. X. Wu, 2009: Cloud radiative forcing in Asian monsoon region simulated by IPCC AR4 AMIP Atmos. Sci., 26, 923-939, doi: 10.1007/, S. L., 2001: Narrowband to broadband conversions of land surface albedo. I: Sens. Environ., 76,213-238, doi: 10.1016/S0034-4257(00), X. Z., and W. C. Wang, 1997: Cloud overlap effects on general circulation model climate , 102, -, doi: 10.1029/, X. Z., and X. Q. Wu, 2005: Evaluation of a GCM subgrid cloud-radiation interaction parameterization using cloudresolving model Res. Lett., 32, L06801,doi: 10.1029/, G. G., and S. Benson-Troth, 2002: Cloud-layer overlap characteristics derived from long-term cloud radar , 15, 2505-2515, doi: 10.1175/1520-0442(2002)015<2505:CLOCDF>2.0.CO;, R., G. G. Mace, T. Ackerman, et al., 2008: Hydrometeor detection usingCloudSat—An earth-orbiting 94-GHz cloud radar.J. Atmos. Oceanic Technol., 25, 519-533, doi:10.1175/, H., M. Satoh, and H. Miura, 2008: A joint satellite and global cloud-resolving model analysis of a Madden-Julian Oscillation event: Model diagnosis..J. Geophys. Res., 113,D, doi: 10.1029/, H., M. Satoh, T. Nasuno, et al., 2007: A Madden-Julian oscillation event realistically simulated by a global cloudresolving , 318, 1763-1765, doi: 10.1126/science..Morcrette, J. J., and Y. Fouquart, 1986: The overlapping of cloud layers in shortwave radiation , 43, 321-328, doi: 10.1175/1520-0469(1986)043<0321:TOOCLI>2.0.CO;, M., and H. Niino, 2006: An improved Mellor-Yamada level-3 model: Its numerical stability and application to a regional prediction of advection -Layer Meteor.,119, 397-407, doi: 10.1007/, C. M., A. Del Genio, G. G. Mace, et al., 2008: Impact of dynamics and atmospheric state on cloud vertical , 21, 1758-1770, doi: 10.1175/, L., and M. Khairoutdinov, 2003: Overlap properties of clouds generated by a cloud-resolving model.J. , 108, 4479, doi: 10.1029/, L., D. Lee, Y. C. Sud, et al., 2012: Radiative impacts of cloud heterogeneity and overlap in an atmospheric general circulation Chem. Phys., 12, 9097-9111, doi:10.5194/, J., H. Zhang, and X. Y. Shen, 2013: Analysis of vertical structure of clouds in East Asia with CloudSat J. Atmos. Sci., 37, 91-100, doi: 10.3878/. (in Chinese)R?is?nen, P., 1998: Effective longwave cloud fraction and maximumrandom overlap of clouds: A problem and a Rev., 126, 3336-3340, doi: 10.1175/1520-0493(1998)126<3336:ELCFAM>2.0.CO;2.R?is?nen, P., H. W. Barker, M. F. Khairoutdinov, et al., 2004:Stochastic generation of subgrid-scale cloudy columns for large-scale J. Roy. Meteor. Soc., 130,2047-2067, doi: 10.1256/, D., M. Khairoutdinov, A. Arakawa, et al., 2003: Breaking the cloud parameterization Amer. , 84, 1547-1564, doi: 10.1175/, T., H. Miura, M. Satoh, et al., 2009: Diurnal cycle of precipitation in the tropics simulated in a global cloud-resolving model.J. Climate, 22, 4809-4826, doi: 10.1175/2009JCL , M., T. Matsuno, H. Tomita, et al., 2008: Nonhydrostatic icosahedral atmospheric model (NICAM) for global cloud resolving Comput. Phys., 227, 3486-3514,doi: 10.1016/, M., T. Inoue, and H. Miura, 2010: Evaluations of cloud properties of global and local cloud system resolving models using CALIPSO and CloudSat Geophys. Res.,115, D00H14, doi: 10.1029/, M., H. Tomita, H. Yashiro, et al., 2014: The non-hydrostatic icosahedral atmospheric model: Description and in Earth and Planetary Science, 1, 18, doi:10.1186/, J. K. P., R. J. Hogan, J. M. Edwards, et al., 2010: Effect of improving representation of horizontal and vertical cloud structure on the Earth’s global radiation budget. Part I: Review and J. Roy. Meteor. Soc., 136,1191-1204, doi: 10.1002/, G. L., 2005: Cloud feedbacks in the climate system: A critical Climate, 18, 237-273, doi: 10.1175/, G. L., D. G. Vane, S. Tanelli, et al., 2008: CloudSat mission: Performance and early science after the first year of Geophys. Res., 113, D00A18, doi: 10.1029/2008 , L., and J. A. Curry, 1989: Cloud overlap Geophys. Res., 94, 9925-9935, doi: 10.1029/, H., and M. Satoh, 2004: A new dynamical framework of nonhydrostatic global model using the icosahedral Dyn. Res., 34, 357-400, doi: 10.1016/, A. M., and F. Di Giuseppe, 2015: An interpretation of cloud overlap Atmos. Sci., 72, 2877-2889, doi:10.1175/, X. C., Y. M. Liu, and Q. Bao, 2016: Impacts of cloud overlap assumptions on radiative budgets and heating fields in convective Res., 167, 89-99, doi: 10.1016/, X. Q., and X.-Z. Liang, 2005a: Radiative effects of cloud horizontal inhomogeneity and vertical overlap identified from a monthlong cloud-resolving model Atmos. Sci.,62, 4105-4112, doi: 10.1175/, X. Q., and X.-Z. Liang, 2005b: Effect of subgrid cloud-radiation interaction on climate Res. Lett.,32, L, doi: 10.1029/, X. Q., and X. F. Li, 2008: A review of cloud-resolving model studies of convective Atmos. Sci., 25,202-212, doi: 10.1007/, F., X.-Z. Liang, J. N. Li, et al., 2013: Dominant roles of subgrid-scale cloud structures in model diversity of cloud radiative Geophys. Res. Atmos., 118, 7733-7749, doi:10.1002/jgrd..Zhang, H., and X. W. Jing, 2010: Effect of cloud overlap assumptions in climate models on modeled earth-atmosphere radiative J. Atmos. Sci., 34, 520-532, doi: 10.3878/ (in Chinese)Zhang, H., and X. W. Jing, 2016: Advances in studies of cloud overlap and its radiative transfer in climate , 30, 156-168, doi: 10.1007/, H., T. Nakajima, G. Y. Shi, et al., 2003: An optimal approach to overlapping bands with correlatedkdistribution method and its application to radiative Geophys. Res., 108, 4641, doi: 10.1029/, H., G. Y. Shi, T. Nakajima, et al., 2006a: The effects of the choice of thek-interval number on radiative Quant. Spectro. Rad. Trans., 98, 31-43, doi: 10.1016/, H., T. Suzuki, T. Nakajima, et al., 2006b: Effects of band division on radiative Eng., 45, 0, doi:10.1117/1..Zhang, H., J. Peng, X. W. Jing, et al., 2013: The features of cloud overlapping in eastern Asia and their effect on cloud radiative China Earth Sci., 56, 737-747, doi: 10.1007/, H., X. Jing, and J. Li, 2014: Application and evaluation of a new radiation code under McICA scheme in BCC_AG Model Dev., 7, 737-754, doi: 10.5194/gmd-7-737-2014.

文章来源:《热带地理》 网址: http://www.rddlzz.cn/qikandaodu/2020/1224/453.html



上一篇:中国:我的梦·我的爱
下一篇:小学教育专业教育见习与实习课程建设与改革研

热带地理投稿 | 热带地理编辑部| 热带地理版面费 | 热带地理论文发表 | 热带地理最新目录
Copyright © 2018 《热带地理》杂志社 版权所有
投稿电话: 投稿邮箱: