Next Article in Journal
Vegetation Identification in Hyperspectral Images Using Distance/Correlation Metrics
Previous Article in Journal
Research on the Emission Reduction Effect of International Technology Import in China’s Key Industries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantifying Raindrop Evaporation Deficit in General Circulation Models from Observed and Model Rain Isotope Ratios on the West Coast of India

by
Saikat Sengupta
1,*,
Sourendra Kumar Bhattacharya
2,
Nimya Sheena Sunil
1,3 and
Sumit Sonar
1
1
Indian Institute of Tropical Meteorology, Ministry of Earth Sciences, Pune 411008, India
2
Institute of Earth Sciences, Academia Sinica, Taipei 11529, Taiwan
3
Department of Atmospheric and Space Sciences, Savitribai Phule Pune University, Pune 411007, India
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(7), 1147; https://doi.org/10.3390/atmos14071147
Submission received: 9 April 2023 / Revised: 10 July 2023 / Accepted: 11 July 2023 / Published: 14 July 2023
(This article belongs to the Section Meteorology)

Abstract

:
Raindrop evaporation is an important sub-cloud process that modifies rainfall amount and rainwater isotope values. Earlier studies have shown that various general circulation models (GCMs) do not incorporate this process properly during the simulation of water isotope ratios (oxygen and hydrogen). Our recent study has demonstrated that an inadequate estimation of this process for the Indian Summer Monsoon (ISM) results in significant biases (model-observed values) in the simulation of various GCMs on a monthly scale. However, a quantitative estimation was lacking. The magnitude of raindrop evaporation depends upon ambient humidity and temperature, which vary considerably during the ISM. Consequently, the isotope biases would also vary over various time scales. The present study aims to investigate the magnitude of the monthly scale variation in raindrop evaporation in the simulations and its causal connection with the corresponding variation in isotope biases. Towards this, we compare an 11-year-long (1997–2007) dataset of rain isotope ratios (both oxygen and hydrogen) from an Indian station, Kozhikode (Kerala), obtained under the Global Network of Isotopes in Precipitation (GNIP) programme of the International Atomic Energy Agency (IAEA) with the corresponding outputs of two isotope-enabled nudged GCMs—ISOGSM and LMDZ4. The raindrop evaporation fractions are estimated for 44 ISM months (June–September) of the study period using the Stewart (1975) formalism. Using a simple condensation–accretion model based on equilibrium fractionation from vapour, obtained from two adopted vapour isotope profiles, we estimate the liquid water isotope ratios at the cloud base. Considering this water as the initial rain, the raindrop evaporation fractions are estimated using the observed oxygen and hydrogen isotope ratios of Kozhikode surface rain samples. The estimated fractions show strong positive correlations with the isotope biases (R2 = 0.60 and 0.66). This suggests that lower estimates of raindrop evaporation could be responsible for the rain isotope biases in these two GCMs.

Graphical Abstract

1. Introduction

The oxygen and hydrogen isotope ratios in water have long been considered important tracers of the hydrologic cycle [1,2,3,4]. These ratios are primarily affected by various physical processes involving phase changes during liquid–vapor transition (evaporation or condensation). Since different water sources in the hydrologic cycle have different evaporation and condensation histories, they usually have different isotopic ratios. Isotope-enabled general circulation models (GCMs) constitute a useful tool to investigate the relationship between water isotopes and climate, since many climatic processes are considered in these models [5,6,7]. These processes include evaporation, advection, condensation, evaporative isotope exchange of raindrops, and the recycling of moisture through evaporation and evapotranspiration. The models have been used to simulate water isotope tracers in the atmosphere [8,9,10,11], the ocean/ice system [12,13,14] and the coupled ocean atmosphere system [15,16,17]. Historical simulation (in multi-decadal scales) of these models is often compared with proxy isotope values for constraining past atmospheric circulation [18,19]. These GCMs reproduce the direction and large-scale patterns of past climate changes successfully. However, they cannot accurately simulate the magnitude of regional changes archived in the local climate proxies [20,21]. The accurate simulation of water isotopes in these models is required not only to correctly estimate the extent of these changes in the modern climate but also to reconstruct/project the past/future climate. However, the accuracy in the simulation can be achieved only by the correct representation of the controlling physical processes.
The Indian sub-continent experiences rainfall governed primarily by the seasonal variation in atmospheric circulation [22]. Most of the rainfall occurs during the Indian Summer Monsoon (ISM; June–September) season. The sub-continent and the adjacent Tibetan Plateau host a rich repository of natural archives from which decadal to centennial-scale climates have been reconstructed [23,24,25]. However, due to the paucity of long-term rain isotope data, studies evaluating the performances of GCMs in this region are limited [26,27]. A recent study by Shi et al. [28] documented a strong seasonality in the rain isotope ratios over the Tibetan Plateau (TP). Their analyses indicate that most GCMs underestimate seasonality of the isotopic composition in the precipitation and vapour. In the southern TP, the upwind precipitation seasonality controls the inter-model spread in isotopic seasonality. This observation is consistent with the findings of our recent study [27], which demonstrated that the rain δ18O biases in North India are inversely proportional to the rainfall biases over the monsoon trough. In addition, this study suggested that the efficacy of these models in Western India depends on how accurately they simulate (a) vapour isotope values in the mid-troposphere and (b) the raindrop evaporation. Raindrops, while falling through the sub-cloud layer of unsaturated atmosphere, exchange isotopes with the ambient vapour [29,30]. The isotopic composition of the falling raindrops changes due to evaporation and exchange. It is difficult to quantify raindrop evaporation directly from the satellite-based meteorological data. However, using water isotope analysis and various parameterization equations, the magnitude of isotope exchange and drop evaporation can be estimated [30,31]. Normally, the isotope exchange model proposed by Stewart [32] is used in most of the GCMs. In that model, the fall of a raindrop in the atmosphere is associated with equilibrium exchange and diffusive transport of vapour isotopologues between the drop surface and the environment. Under certain simplifying assumptions, the Stewart model estimates the final isotopic composition of a raindrop as a function of temperature, normalized relative humidity, isotope ratios of the initial raindrop and ambient vapour, and ratio of the final mass to the initial mass of the drop (see later in Section 3.2).
While most of the GCMs use this evaporation-exchange scheme to compute final isotope ratios of the raindrops at the ground level, there is no way we can separate out the sub-cloud evaporation contribution, which plays a major role [30]. The present study is an attempt to evaluate this contribution. Using a model vapour isotope profile, we can estimate the composition of the cloud droplets at various levels inside a cloud, assuming equilibrium condensation and zero evaporation (saturated humidity assumption). These cloud droplets can grow and combine to make rain droplets that descend through the cloud to reach the cloud base. The raindrop composition at the cloud base is estimated by the isotope mass balance of these droplets. This composition serves as the initial isotope ratio to be used for obtaining the sub-cloud evaporation.
In an earlier study dealing with the improper estimation of drop evaporation in the GCMs, we adjusted their simulated vapour isotope profiles to match the available surface and satellite based data [27]. Since the objective of that study was to carry out an inter-comparison of seven different GCMs, a long-term mean isotope profile of a common time window (1997–2007) was considered. However, the magnitude of the sub-cloud raindrop evaporation depends on the ambient air temperature and humidity and should vary in various summer monsoonal months. Especially in the latter part of the summer monsoon (August–September), less evaporation is expected due to high humidity in the atmosphere [33]. These aspects were not addressed in that study. In the present work, we explore the monthly variation in drop evaporation during the ISM months and corresponding rain isotope biases of the GCMs.

2. Data and Methods

2.1. Study Area and Choice of GCMs

The limited available studies dealing with the performance of the GCMs over the Asian monsoon region show that two nudged models—LMDZ4 and IsoGSM Nudged [26,27,28]—perform better over these regions. The other models, in contrast, show large biases in various physical fields like temperature, humidity, wind etc., both in monthly and seasonal scales. Therefore, we decided to consider only IsoGSM and LMDZ4 models for the estimation of raindrop evaporation effects. The model outputs are available at the Stable Water Isotope Intercomparison Group, Phase 2 (SWING2; https://data.giss.nasa.gov/swing2/swing2_mirror/; [34], accessed on 8 April 2023). Monthly outputs of simulated rain isotope values (δ18O and δD) for summer monsoon (June–September) months are extracted for the period (1997–2007) over the location Kozhikode (11.25° N; 75.77° E), where corresponding observed isotope data are available for the same time window. Kozhikode, situated on the west coast of India at an altitude of 1 masl, falls under the Global Network of Isotopes in Precipitation (GNIP) program [35]. The station experiences a tropical monsoon climate with brief spells of pre-monsoon (April–May) shower. It receives a substantial rainfall (average monsoonal rainfall = 2237 mm) during the summer monsoon (June–September) and a sizable amount (on average 444 mm) of rain during the winter monsoon (October–December). During the summer monsoon months, the daytime air temperature remains nearly constant at around 26 °C (https://imdpune.gov.in/library/public/1981-2010%20CLIM%20NORMALS%20%28STATWISE%29.pdf, accessed on 8 April 2023). Considering the availability and continuity of the observed and model rain/vapour isotope datasets, Kozhikode serves as an appropriate location for this analysis. Furthermore, being situated on the coast, the contribution of inland moisture recycling through evaporation and evapotranspiration is insignificant compared to the advected moisture flux during the ISM (Figure 1).

2.2. Gridded Meteorological Data

For estimating the effect of evaporation on the monthly rain isotope ratios, monthly gridded data (relative humidity, air temperature, specific humidity, wind (both zonal and meridional) and specific cloud liquid water content (CLWC) for the pressure levels 1000 to 100 mb from the European Centre for Medium-Range Weather Forecasts Reanalysis (ERA 5) reanalysis dataset with a resolution of 0.25° × 0.25° [36] are used. Vertically integrated moisture transport (VIMT) is calculated from the specific humidity, zonal and meridional wind components, integrated from 300 to 1000 mb using the standard equation [37]. The mean monsoon VIMT and moisture fluxes are shown in Figure 1. The moisture fluxes are very large and a comparison with precipitation indicates that only about 15% of this moisture is precipitated [38]. Moreover, the contribution from evapotranspiration is negligible compared to this flux.

2.3. Satellite-Based Vapour Isotope Data

To investigate how the models simulate vapour isotope values at various pressure levels, we compare them with the Tropospheric Emission Spectrometer (TES) Level 2 (Nadir-Lite-Version 6) retrievals of HDO and H2O profiles. The data are available for the period (2005–2007) [39,40]. Vapour δD data are obtained from 17 pressure levels with a 5.3 km × 8.4 km footprint. These observations have a precision of 1–2‰ when the zonal average of this data is considered [34,41]. All valid TES observations of HDO and H2O concentrations over a box (6°–14° N and 69°–78° E) are extracted (Figure 1). This box includes Kozhikode and part of the Arabian Sea.

3. Model for Estimation of Raindrop Evaporation Effect

We carry out the calculations in two steps for estimating the raindrop evaporation effects: (1) calculations that are involved in vapour to cloud droplets and cloud droplets to raindrop formation processes inside the convective cloud system and (2) sub-cloud evaporative exchange below the cloud base until the raindrops reach the ground.

3.1. Rain Forming Processes and Raindrop Isotopic Composition at the Cloud Base Height

The rainfall over Kozhikode during the ISM season (JJAS) occurs mainly due to orography-induced convective processes [42,43]. In these cases, rains form due to rapid uplift of moisture-laden vapour that originates by evaporation from the Arabian Sea. The vapour is carried inland by strong westerly winds impinging on the coast in a perpendicular direction (Figure 1). The available ceilometer observations suggest that majority of the clouds occur at a low level (LLC; average Cloud Base height (CBH) = 800 m) for rains over the study region during the ISM [44]. The ceilometer study also shows that clouds hover around the LLC region about 90% of the time during the active phases of the ISM. In addition, the occurrence of clouds is rare above 6 km, a region where the ice phase forms [45]. For simplicity, we assume that the contribution from higher-level clouds or ice phase is small and can be neglected. This implies that most of the rainfall over the Western Ghat region is induced by shallow convection.
The cloud droplets form from the ambient vapour and an isotopic fractionation occurs during condensation. Therefore, to estimate isotope ratios of the cloud droplets, the composition of the ambient vapour needs to be known/estimated at various heights. It is known that certain suitable adjustments and zonal integrations provide observational validation for the profiles of the two chosen models used here [6,7]. The height-specific monthly average vapour δD and δ18O values from the IsoGSM/LMDZ4 outputs (see later in Section 4) are obtained for the JJAS monsoon months over the period 1997–2007. All gridded meteorological data and model vapour isotope outputs are averaged over the box surrounding Kozhikode (6–14° N; 74–78° E; Figure 1). The parameters used in the calculations (for Section 3.1 and Section 3.2) are the following: air temperature, relative humidity, vertical profiles of CLWC, model vapour isotope ratios, and observed ground-level rain isotope ratios at Kozhikode.
Available CLWC data show that cloud droplets exist at various pressure levels (from 1000 mb to 100 mb), with major liquid water peaks at 550 and 900 mb levels. Raindrops form by the collision and coalescence of the small cloud water droplets. The bigger drops come down without suffering any exchange or evaporation effect within the clouds. This situation prevails because the presence of significant amount of cloud water at these levels is indicative of super-saturated humidity, where evaporative exchange is minimal [29,30]. The ground-level rain (collected over a period) is derived from a uniform mix of these raindrops that arrive at the cloud base without alteration. It is important to mention here that the presence of CLWC does not explicitly confirm rain formation. The rain liquid water content (RLWC) is a better indicator for that. The connection between the CLWC and RLWC is not always clear. However, in a study based on a micro-rain radar, Zhang et al. [46] showed that the Cloud Liquid Water Path (CLWP) is larger than the Rain Liquid Water Path (RLWP; integral for the rain water concentrations up to the top of the atmosphere), but they both show a similar variation. They also found that both CLWP and RLWP are positively correlated with the rain rate, and the precipitation efficiency (100% × RLWP/CLWP) is high (about 50%) when precipitation finally happens. From a case study in central China, they showed that during the period of increasing precipitation, both CLWC and RLWC increase with height (up to about 3 km) in a similar fashion (see Figure 2). The RLWC data require micro rain radar measurements, which are not available in and around our study area for the said period. Therefore, for the present purpose, we assume that in case it rains, the raindrop formation rate would be in direct proportion to the CLWC values in the average profile.
We use the available CLWC average monthly profile to estimate the altitude variation in the rain formation rate by taking the CLWC values as statistical weights in the rain liquid water content as a function of height. Following earlier studies [29,32,47], we also consider that condensation takes place in equilibrium, and there is no fractionation during the coalescence of cloud drops in the course of raindrop formation. The fractionation factors for the condensation of vapour to liquid at various temperatures corresponding to the altitudes are determined using the formula of Horita and Wesolowski, 1994 [48], for both δD and δ18O. The isotope ratios (δi) of the rains formed at different altitudes (ith level) are determined using these fractionation factors. In this approach, we cannot determine the amount of total rain but only the average isotope ratios of the total rain at the cloud base height (CBH) for different months using the following equation:
δ r ( C B H ) = i = 1000   m b i = 100   m b δ i   C i h i i = 1000   m b i = 100   m b C i h i
The subscript “i” refers to the altitude/pressure level (1000 mbar to 100 mbar in 25 mbar intervals up to 750 and then in 50 mbar intervals, denoted by Δh). Ci denotes the CLWC values at the ith level.
The above procedure is a simplified version of the condensation and coalescence process (hereafter referred to as the condensation-accretion process) that allows us to calculate the isotope ratio of the liquid rain at the cloud base. We made the following assumptions: (1) no in-cloud exchange or evaporation takes place, (2) no updraft or downdraft occurs, (3) condensation occurs in equilibrium between the vapour and liquid at various altitudes within the cloud, and (4) liquid rain amount is proportional to the CLWC. In actual practice, rain formation and its downward migration is a complex process. Admittedly, since no data for the isotope ratio of liquid rain leaving the cloud base are available, the procedure that we follow cannot be verified. However, we note that the IsoGSM-Nudged model simulates a δD value of −77‰ (Kei Yoshimura, Personal Communication) for the liquid at the average CBH over Kozhikode, which is only 10‰ lower compared to our average accretion model estimate of −67‰ for the same period [27]. The accumulated raindrops at CBH (varying from 820 mbar to 925 mbar) then fall into the sub-cloud region where the air is under-saturated and warmer. The falling raindrops undergo evaporative exchange with the environmental vapour, and their isotope ratios are modified. We use the Stewart model [32] to estimate the magnitude of this evaporative enrichment.

3.2. Sub-Cloud Evaporative Exchange

The Stewart (1975) model is essentially a Craig–Gordon [2,49]-type model, which deals with the evaporation of water drops falling through the atmosphere with varying relative humidity and temperature levels and exchanging isotopes with the ambient moisture of different isotopic compositions. The isotope ratio of falling raindrops increases due to evaporation and exchange with ambient vapour (which has higher isotope ratios relative to the higher-level vapour inside the cloud) in the sub-cloud layer. The evaporation/exchange process simulated by Stewart [32] is usually adopted by all GCMs, where the isotopic composition of falling raindrops is given by,
[Rr(g) − γRv(amb)] = [Rr(cb) − γRv(amb)] × fβ
where Rr(g) and Rr(cb) represent the ground level (g subscript) and cloud base (cb subscript) isotope ratios of the drop, and Rv(amb) is the ratio of the ambient (amb subscript) vapour. The parameter f denotes the fraction of the raindrop mass remaining. The constants γ and β involve three parameters: fractionation factor, relative humidity, and diffusivity ratios, as given below:
β = 1 α p ( D / D ) n ( 1 h ) α p ( D / D ) n ( 1 h )
γ = α p h 1 α p ( D / D ) n ( 1 h )
where αp is the equilibrium fractionation factor (liquid to vapour), D and D′ are the diffusivities for H216O and HD16O (or H218O), n = 0.58, and h is the relative humidity. We assume that the isotope ratio of the sub-cloud vapour, the RH and temperature do not change significantly between the cloud base and ground, and we can use the average values of the temperature, RH, Rv(amb), and the fractionation factor for water to vapour conversion for the calculations. Using the three R-values and the two constants in equation 2 corresponding to the D/H and 18O/16O ratios, we finally estimate the remaining fraction of the rain (f) for the 44 ISM months. The raindrop evaporation fraction is then given by (1-f). The simulated rain δD and δ18O values of two GCMs, our estimated liquid isotope values at CBH, the observed rain isotopes at Kozhikode, and estimated raindrop evaporation fractions are given in Table 1 (δD) and Table 218O).

3.3. Uncertainty Calculation by Numerical Quadrature Method

The uncertainty in the evaporation estimates is obtained by varying the values of the variables in the Stewart equation. The evaporation depends on five parameters: temperature, relative humidity, initial rain composition at the CBH, final rain composition at the ground, and the sub-cloud vapour isotope composition. Each of these parameters can vary and affect the final estimated value of evaporation. The uncertainty in the evaporation is calculated by the following formula:
Assume   x = f u , v , .
σ x       2   σ u   2 δ x δ u 2 +   σ v 2 δ x δ v 2 +
We assume that the errors in the parameters are as follows: σ(T) = 2 °C, σ(RH) = 2%, σ(δDini) = ±2‰, σ(δDfin) = ±2‰, and σ(δDamb) = ±2‰. For example, Table 3 shows the calculated final error in the evaporation (1-f) for the case of June 1997 (IsoGSM), corresponding to the first entry in Table 1. The sigma column indicates individual error terms, and the final column gives the error (calculated by quadrature) in the evaporation (1-f), which is 0.27 ± 0.06 (for the chosen month), or about 27%. The sensitivity is highest for the temperature, followed by the ambient vapour and the final rain isotope ratios. Considering all the other entries, typical error values are ±20%.

4. Results

As mentioned earlier (Section 2.3), the TES observations of δD values are available for a smaller period compared to their model counterparts, while to check the performances of the models, we need to use their vapour isotope profiles for the 44 months during 1997–2009. Therefore, for validation, we compared the results obtained by using the TES and the two models’ vapour isotope profiles over a common interval. Mean averaging kernels of all valid observations for the summer monsoon months of 2005–2007 suggest that the TES observations over the study region between the pressure levels 750 mb and 400 mb are of acceptable quality. Therefore, all such observations at different pressure levels for the summer monsoon months of 2005–2007 are averaged over the box (Figure 1). Notably, the TES observations are not very sensitive in the atmospheric boundary layer, especially below 750 mb in our study area. To circumvent this issue, we used the available surface vapour isotope observations (mean monsoon δD ~78‰; [33]) at the location Wayanad (around 70 km away) as the surface value and extrapolated the curve from the surface to 750 mb and above 400 mb to construct the adopted vapour δD profiles for TES. The mean monsoon vapour δD profiles of TES and the two models (IsoGSM and LMDZ4) over the box for the synchronous period are shown in Figure 3. The average TES profile is considered as representing the ‘true’ atmospheric situation (called ‘observed’). The figure shows that IsoGSM values reasonably agree with the TES values from the surface to 400 mb and deviate only slightly above that level. However, notable differences (mean model—TES) of about −30‰ exist for LMDZ4. To explore whether these differences have any impact on the drop evaporation estimates, we determined the rain δD composition at the CBH by using the three vapour δD profiles (TES and the two models) following our condensation-accretion model discussed in Section 3.1. The drop evaporation fraction is estimated from the difference between the liquid δD at CBH (acting as the initial rain) calculated for each model and the observed GNIP average rain value at the ground (acting as the final rain). The values given in Table 4 show that the calculated evaporation differs only by 2% for IsoGSM and 7% for LMDZ4; these two differences are small (within and slightly above the uncertainty of the estimates, respectively). This provides adequate validation of our conceived condensation-accretion model along with its assumptions.

5. Discussion

5.1. Comparison between Observed and Model Rain Isotope Values

Table 1 and Table 2 show that barring 5 cases out of 44, all IsoGSM simulations are lower compared to the observed values. An overall negative rain isotope bias is observed for both models (on average −8‰ in δD and −1‰ in δ18O for IsoGSM; −12‰ in δD and −1.4‰ in δ18O for LMDZ4; Table 1 and Table 2). The rain δ18O and δD biases are strongly correlated for both models (Figure 4), suggesting that same processes control both isotope biases. The data points fall reasonably well with a mean regression line: δD = 7.3 (±0.27) × δ18O −1.4 (±0.47) for IsoGSM and δD = 7.9 (±0.17) × δ18O − 0.96 (±0.52) for LMDZ4. The model, IsoGSM (LMDZ4), under or (over) estimates the range of variation (δD varying −8‰ to −29‰ for IsoGSM and +9‰ to −53‰ for LMDZ4) as compared to observed values (5‰ to −40‰; Table 1). The difference in the range of variations (between observed and model values) is clearly evident in the frequency distribution of δD values (Figure 5). The negative isotope biases in the models during the ISM have also been reported in a few earlier studies [26,28,50]. In our earlier study [27], we identified that the underestimation of the mean isotope values can be due to (a) an inaccurate vapour isotope profile and (b) a lower estimate of the raindrop evaporation. By comparing the IsoGSM simulation with the TES estimation for the same period (2005–2007), we see that the difference (at the CBH) could be at most 4‰ (in δD) due to the isotope profile difference (Table 4), which is too small to account for the discrepancy, assuming the TES represents a realistic profile. However, the aforementioned difference is large (38‰) for LMDZ4. Therefore, for LMDZ4, we cannot rule out a significant contribution from the vapour isotope bias towards the negative rain isotope bias. The above result suggests that we should look for a lower evaporation estimate as the main factor for the models’ discrepancy.

5.2. Rain Evaporation Estimates and Their Relations with Model Rain Isotope Biases

Table 1 shows that average raindrop evaporation required to match the observed δD values at Kozhikode is about 32% (±12%). If the IsoGSM model rain isotope values were taken as the final rain isotope values (instead of the observed rain isotope values), the average evaporation required to reproduce them would be about 22%. This suggests a shortfall of about 10% in evaporation (though evaporative effects, sensu stricto, do not add linearly) in the IsoGSM. On the other hand, the mean rain evaporation required for the LMDZ4 model to match the observed values is 34% (±12%), and the corresponding shortfall is about 12%.
We also carried out similar exercises using the δ18O values simulated by IsoGSM and LMDZ4. Table 2 shows that the evaporation required to match the observed δ18O values is on average about 18% (±5%) for IsoGSM and 22% (±8%) for LMDZ4. The discrepancy between the two estimates (i.e., based on δD and δ18O vapour profiles) using the condensation-accretion process is quite large. We explore the reason for this difference below. Since the only input from the models is the vapour isotope values, it seems reasonable to assume that an inconsistency in the two isotope profiles (δD and δ18O) may be responsible for this discrepancy. As the TES observations do not provide δ18O values, the corresponding model outputs could not be validated. Therefore, it is possible that model vapour δ18O values are higher than they should be (considering what one expects from the δD profile), which results in slightly more enriched liquid δ18O simulations at the CBH, thus requiring less evaporation. This suggestion is corroborated by the δD-δ18O correlation plots (Figure 6 and Figure 7). The δD-δ18O correlations for the observed and models’ rains are shown in Figure 6. The same correlations for model vapour and estimated liquid isotopes at CBH are displayed in Figure 7. The local meteoric water line constructed on the basis of simulated rain at Kozhikode (δD = 6.8 (±0.5) × δ18O + 5.9 (±1.68); Figure 6) and the vapour profile over Kozhikode (δD = 6.63 (±0.06) × δ18O −11.92 (±3.12); Figure 7a) in IsoGSM have lower slopes and intercepts compared to the observed values: δD = 7.7 (±0.17) × δ18O + 9.6 (±0.46); (Figure 6). In contrast, the estimated liquid values at the cloud base based on our condensation-accretion process align with δD = 8 (±0.12) × δ18O + 16.8 (±1.35); (Figure 7b). The δ(δD)/δ(δ18O) differential is much less for the IsoGSM simulation (both vapours; Figure 7a and ground rain; Figure 6) and indicates a relatively higher value of δ18O vapour with respect to the δD vapour in the profiles. We cannot identify the real reason, but this could be caused by nonequilibrium fractionation like the diffusion of vapour inside the cloud. In the case of equilibrium fractionation, we obtain:
d δ 2 H d δ 18 O = α 2 H 1 α 18 O 1     8 1
However, in the case of diffusion (like raindrop evaporation in an unsaturated environment), the smaller diffusion velocities of HD16O and H218O compared to H216O lead to additional fractionation. Since H218O is heavier than HD16O, nonequilibrium fractionation is slightly stronger for H218O than for HD16O. In our condensation-accretion model, we have considered only equilibrium fractionation, leading to a slope close to 8.
The cross-plots of model isotope biases (both δD and δ18O) against the evaporation fraction (1-f) estimated from δD and δ18O are shown in Figure 8a,b, respectively. Both models show good correlation between δD bias and (1-f) (Figure 8a); this is expected if the discrepancy originates from a deficiency in raindrop evaporation. Since model values are less than the observed GNIP values, the bias is negative in sign. More negative values imply more evaporation is needed to match the observed values.
However, the correlations are relatively weak for both models in the case of δ18O bias and the (1-f) cross-plot (Figure 8b). The raindrop evaporation fractions calculated by δ18O and δD show a stronger correlation for IsoGSM (Figure 8c). The possible reason behind the different estimation of (1-f) for the two models has been discussed above. In addition, as we discussed earlier (Section 4), the vapour isotope values have been largely underestimated in the LMDZ4 model, which can affect the (1-f) estimation following our model. This could be a possible reason for obtaining very different slope, intercept, and correlation values in various correlation plots for the LMDZ4 model. However, even with this limitation, the rain δD bias in this model is correlated with (1-f), suggesting that raindrop evaporation must have notable control on the simulation. Decoupling the effect of this vapour isotope bias from the estimated (1-f) values for this model is beyond the scope of the present study. Based on the above discussion, we propose that the lower isotope values simulated by IsoGSM are due to lower estimates of evaporation (by about 10%) of the IsoGSM in Kozhikode rain samples.
Although our simple model could estimate the liquid isotopic composition at the cloud base and subsequently the drop evaporation amounts that should be adjusted in the GCMs for the correct simulation of rain isotope ratios, our study has some limitations. As mentioned in Section 3.1, our model does not consider any updraft, downdraft, or advection. Moreover, we consider that drop evaporation occurs only below the CBH (Section 3.2), whereas evaporation can take place wherever the relative humidity is <100%, even above the CBH. Therefore, further improvement of our approach is required after taking care of these shortcomings to refine the evaporation estimates reported in the study.

6. Conclusions and Implications

  • The rain and vapour isotope simulations of IsoGSM and LMDZ4 models are compared with the observations from Kozhikode (GNIP data). Both models underestimate the long-term mean monsoon rain δD and δ18O values at Kozhikode, on average by 8‰ and 1.0‰ for IsoGSM and 12‰ and 1.4‰ for LMDZ4, respectively. However, the observed variation in rain isotope values (from +5 to −40‰ for δD and −0.6 to −6.7‰ for δ18O) is underestimated in IsoGSM but overestimated in LMDZ4.
  • To estimate the magnitude of this underestimation, we first determined the isotope ratios of the accumulated liquid water at the cloud base height (CBH) using the GCMs’ vapour isotope profiles and a simple condensation-accretion process (without any in-cloud modification). Treating this composition as the initial rain isotope composition and the observed values as the final rain composition, we used the Stewart model to estimate the raindrop evaporation fractions on monthly scale.
  • The evaporation fraction (1-f) varies from 2% to 56% with a mean of 32% (±12%) for IsoGSM when the fraction is computed from δD values. The same value for the LMDZ4 varies from 10% to 55% with a mean of 34% (±12%). We propose that the lower isotope values simulated by IsoGSM and LMDZ4 are due to lower estimates of evaporation (by about 10%) in those models.
  • The evaporation is also calculated using the vapour δ18O profiles for both models. The evaporations required in this case are much less: 18% ± 5% for IsoGSM and 22% ± 8% for LMDZ4. This can be traced to the inconsistency of the adopted δ18O profile compared to that expected from the δD profile. The inconsistency can possibly happen because of the δ18O profile, which cannot be validated against the TES observation.
  • To estimate the raindrop evaporation fraction on monthly scale, we rely on the vapour δD profiles obtained from the models. Although vapour δD values for the IsoGSM model reasonably agree with the TES observation on a long-term mean seasonal scale, LMDZ4 shows a significant negative bias. For the same reason, the liquid water compositions at CBH estimated from TES profiles grossly agree with the same obtained from the IsoGSM model (TES: −73.5‰; IsoGSM: −77.7‰). However, the value is much lower—about −112‰ for the LMDZ model. Even with stronger bias in the vapour values, raindrop evaporation fractions estimated by LMDZ4 strongly correlate with rain δD biases, suggesting raindrop evaporation indeed affects the rain isotope simulation.
  • Our study region is very close to several climate proxies (speleothem and tree ring) locations in the Western Ghat Mountains. The isotope studies of these proxies provided valuable information on past rainfall variability in centennial scales. The present study shows that a monthly variation in raindrop evaporation fractions strongly affects rain isotope simulation. Therefore, the improvement of evaporation schemes in the GCMs may result in better rain isotope simulation, which would be useful in understanding past atmospheric circulation.

Author Contributions

Conceptualization, S.S. (Saikat Sengupta) and S.K.B.; methodology, S.S. (Saikat Sengupta) and S.K.B.; software, N.S.S. and S.S. (Sumit Sonar); validation, S.S. (Saikat Sengupta) and S.K.B.; formal analysis, N.S.S. and S.S. (Sumit Sonar); investigation, S.K.B.; data curation, N.S.S. and S.S. (Sumit Sonar); writing—original draft preparation, S.S. (Saikat Sengupta) and S.K.B.; writing—review and editing, All four authors; visualization, S.S. (Saikat Sengupta) and S.K.B.; supervision, S.S. (Saikat Sengupta); All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The model outputs are available at the Stable Water Isotope Intercomparison Group, Phase 2 (SWING2; https://data.giss.nasa.gov/swing2/swing2_mirror/ (accessed on 8 April 2023)). Various gridded meteorological data (temperature, humidity, wind speed, etc.) are used from ERA-5 (https://www.ecmwf.int/en/forecasts/datasets/reanalysis-datasets/era5 (accessed on 8 April 2023)). Vapour isotope data are availablefrom NASA Tropospheric Emission Spectrometer Mission https://tes.jpl.nasa.gov/tes/ (accessed on 8 April 2023).

Acknowledgments

The Indian Institute of Tropical Meteorology, Pune (IITM), is fully supported by the Earth System Science Organization (ESSO) of the Ministry of Earth Sciences, India. We thank Director IITM for his constant encouragement. We also thank SWING2 project contributors and the IAEA/WMO for the GNIP dataset.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Clark, I.D.; Fritz, P. Environmental Isotopes in Hydrogeology; CRC Press: Boca Raton, FL, USA, 1997; ISBN 1-56670-249-6. [Google Scholar]
  2. Craig, H.; Gordon, L.I. Deuterium and Oxygen 18 Variations in the Ocean and Marine Atmosphere; Stable Isotopes in Oceanographic Studies and Paleotemperatures: Pisa, Italy, 1965. [Google Scholar]
  3. Gat, J. Atmospheric Water Balance—The Isotopic Perspective. Hydrol. Process. 2000, 14, 1357–1369. [Google Scholar] [CrossRef]
  4. Kendall, C.; McDonnell, J.J. Preface. In Isotope Tracers in Catchment Hydrology; Elsevier: Amsterdam, The Netherlands, 1998; pp. vii–ix. ISBN 978-0-444-81546-0. [Google Scholar]
  5. Kurita, N.; Noone, D.; Risi, C.; Schmidt, G.A.; Yamada, H.; Yoneyama, K. Intraseasonal Isotopic Variation Associated with the Madden-Julian Oscillation. J. Geophys. Res. Atmos. 2011, 116, 1–20. [Google Scholar] [CrossRef]
  6. Risi, C.; Bony, S.; Vimeux, F.; Jouzel, J. Water-stable Isotopes in the LMDZ4 General Circulation Model: Model Evaluation for Present-day and Past Climates and Applications to Climatic Interpretations of Tropical Isotopic Records. J. Geophys. Res. Atmos. 2010, 115, D12118. [Google Scholar] [CrossRef] [Green Version]
  7. Yoshimura, K.; Kanamitsu, M.; Noone, D.; Oki, T. Historical Isotope Simulation Using Reanalysis Atmospheric Data. J. Geophys. Res. Atmos. 2008, 113, 2983–2998. [Google Scholar] [CrossRef]
  8. Hoffmann, G.; Werner, M.; Heimann, M. Water Isotope Module of the ECHAM Atmospheric General Circulation Model: A Study on Timescales from Days to Several Years. J. Geophys. Res. Atmos. 1998, 103, 16871–16896. [Google Scholar] [CrossRef] [Green Version]
  9. Joussaume, S.; Jouzel, J. Paleoclimatic Tracers: An Investigation Using an Atmospheric General Circulation Model under Ice Age Conditions: 2. Water Isotopes. J. Geophys. Res. Atmos. 1993, 98, 2807–2830. [Google Scholar] [CrossRef]
  10. Jouzel, J.; Koster, R.; Suozzo, R.; Russell, G.; White, J.; Broecker, W. Simulations of the HDO and H2 18O Atmospheric Cycles Using the NASA GISS General Circulation Model: Sensitivity Experiments for Present-day Conditions. J. Geophys. Res. Atmos. 1991, 96, 7495–7507. [Google Scholar] [CrossRef]
  11. Noone, D.; Simmonds, I. Associations between Δ18O of Water and Climate Parameters in a Simulation of Atmospheric Circulation for 1979–95. J. Clim. 2002, 15, 3150–3169. [Google Scholar] [CrossRef]
  12. Delaygue, G.; Jouzel, J.; Dutay, J.-C. Oxygen 18–Salinity Relationship Simulated by an Oceanic General Circulation Model. Earth Planet. Sci. Lett. 2000, 178, 113–123. [Google Scholar] [CrossRef]
  13. Paul, A.; Mulitza, S.; Pätzold, J.; Wolff, T. Simulation of Oxygen Isotopes in a Global Ocean Model. In Use of Proxies in Paleoceanography: Examples from the South Atlantic; Springer: Berlin/Heidelberg, Germany, 1999; pp. 655–686. ISBN 978-3-642-63681-3. [Google Scholar]
  14. Schmidt, G.A. Oxygen-18 Variations in a Global Ocean Model. Geophys. Res. Lett. 1998, 25, 1201–1204. [Google Scholar] [CrossRef]
  15. Schmidt, G.A.; LeGrande, A.N.; Hoffmann, G. Water Isotope Expressions of Intrinsic and Forced Variability in a Coupled Ocean-atmosphere Model. J. Geophys. Res. Atmos. 2007, 112, D10. [Google Scholar] [CrossRef]
  16. Tindall, J.C.; Valdes, P.J.; Sime, L.C. Stable Water Isotopes in HadCM3: Isotopic Signature of El Nino–Southern Oscillation and the Tropical Amount Effect. J. Geophys. Res. Atmos. 2009, 114, D4. [Google Scholar] [CrossRef] [Green Version]
  17. Zhou, J.; Poulsen, C.; Pollard, D.; White, T. Simulation of Modern and Middle Cretaceous Marine Δ18O with an Ocean-atmosphere General Circulation Model. Paleoceanography 2008, 23, 3. [Google Scholar] [CrossRef] [Green Version]
  18. Brázdil, R.; Dobrovolný, P.; Luterbacher, J.; Moberg, A.; Pfister, C.; Wheeler, D.; Zorita, E. European Climate of the Past 500 Years: New Challenges for Historical Climatology. Clim. Change 2010, 101, 7–40. [Google Scholar] [CrossRef] [Green Version]
  19. Vuille, M.; Burns, S.J.; Taylor, B.L.; Cruz, F.W.; Bird, B.W.; Abbott, M.B.; Kanner, L.C.; Cheng, H.; Novello, V.F. A Review of the South American Monsoon History as Recorded in Stable Isotopic Proxies over the Past Two Millennia. Clim. Past 2012, 8, 1309–1321. [Google Scholar] [CrossRef] [Green Version]
  20. Braconnot, P.; Harrison, S.P.; Kageyama, M.; Bartlein, P.J.; Masson-Delmotte, V.; Abe-Ouchi, A.; Otto-Bliesner, B.; Zhao, Y. Evaluation of Climate Models Using Palaeoclimatic Data. Nat. Clim. Change 2012, 2, 417–424. [Google Scholar] [CrossRef]
  21. Harrison, S.; Bartlein, P.; Brewer, S.; Prentice, I.; Boyd, M.; Hessler, I.; Holmgren, K.; Izumi, K.; Willis, K. Climate Model Benchmarking with Glacial and Mid-Holocene Climates. Clim. Dyn. 2014, 43, 671–688. [Google Scholar] [CrossRef] [Green Version]
  22. Krishnamurthy, V.; Kinter, J.L., III. The Indian Monsoon and Its Relation to Global Climate Variability. In Global Climate: Current Research and Uncertainties in the Climate System; Springer: Berlin/Heidelberg, Germany, 2003; pp. 186–236. ISBN 978-3-662-05285-3. [Google Scholar]
  23. Gautam, P.K.; Narayana, A.C.; Band, S.T.; Yadava, M.G.; Ramesh, R.; Wu, C.-C.; Shen, C.-C. High-Resolution Reconstruction of Indian Summer Monsoon during the Bølling-Allerød from a Central Indian Stalagmite. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2019, 514, 567–576. [Google Scholar] [CrossRef]
  24. Yao, T.; Shi, Y.; Thompson, L.G. High Resolution Record of Paleoclimate since the Little Ice Age from the Tibetan Ice Cores. Quat. Int. 1997, 37, 19–23. [Google Scholar] [CrossRef]
  25. Zhang, Q.; Cheng, G.; Yao, T.; Kang, X.; Huang, J. A 2326-year Tree-ring Record of Climate Variability on the Northeastern Qinghai-Tibetan Plateau. Geophys. Res. Lett. 2003, 30, HLS 2–1. [Google Scholar] [CrossRef] [Green Version]
  26. Midhun, M.; Ramesh, R. Validation of Δ18O as a Proxy for Past Monsoon Rain by Multi-GCM Simulations. Clim. Dyn. 2016, 46, 1371–1385. [Google Scholar] [CrossRef]
  27. Nimya, S.; Sengupta, S.; Parekh, A.; Bhattacharya, S.K.; Pradhan, R. Region-Specific Performances of Isotope Enabled General Circulation Models for Indian Summer Monsoon and the Factors Controlling Isotope Biases. Clim. Dyn. 2022, 59, 3599–3619. [Google Scholar] [CrossRef]
  28. Shi, X.; Risi, C.; Li, L.; Wang, X.; Pu, T.; Zhang, G.; Zhang, Y.; Wang, Z.; Kong, Y. What Controls the Skill of General Circulation Models to Simulate the Seasonal Cycle in Water Isotopic Composition in the Tibetan Plateau Region? J. Geophys. Res. Atmos. 2022, 127, e2022JD037048. [Google Scholar] [CrossRef]
  29. Graf, P.; Wernli, H.; Pfahl, S.; Sodemann, H. A New Interpretative Framework for Below-Cloud Effects on Stable Water Isotopes in Vapour and Rain. Atmos. Chem. Phys. 2019, 19, 747–765. [Google Scholar] [CrossRef] [Green Version]
  30. Salamalikis, V.; Argiriou, A.; Dotsika, E. Isotopic Modeling of the Sub-Cloud Evaporation Effect in Precipitation. Sci. Total Environ. 2016, 544, 1059–1072. [Google Scholar] [CrossRef] [PubMed]
  31. Worden, J.; Noone, D.; Bowman, K. Importance of Rain Evaporation and Continental Convection in the Tropical Water Cycle. Nature 2007, 445, 528–532. [Google Scholar] [CrossRef]
  32. Stewart, M.K. Stable Isotope Fractionation Due to Evaporation and Isotopic Exchange of Falling Waterdrops: Applications to Atmospheric Processes and Evaporation of Lakes. J. Geophys. Res. 1975, 80, 1133–1146. [Google Scholar] [CrossRef]
  33. Lekshmy, P.; Midhun, M.; Ramesh, R. Influence of Stratiform Clouds on ΔD and Δ18O of Monsoon Water Vapour and Rain at Two Tropical Coastal Stations. J. Hydrol. 2018, 563, 354–362. [Google Scholar] [CrossRef]
  34. Risi, C.; Noone, D.; Worden, J.; Frankenberg, C.; Stiller, G.; Kiefer, M.; Funke, B.; Walker, K.; Bernath, P.; Schneider, M. Process-evaluation of Tropospheric Humidity Simulated by General Circulation Models Using Water Vapor Isotopologues: 1. Comparison between Models and Observations. J. Geophys. Res. Atmos. 2012, 117, D5. [Google Scholar] [CrossRef] [Green Version]
  35. IAEA/WMO Global Network of Isotopes in Precipitation. The GNIP Database. Available online: http://www.iaea.org/water (accessed on 12 June 2018).
  36. Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.; Horányi, A.; Muñoz-Sabater, J.; Nicolas, J.; Peubey, C.; Radu, R.; Schepers, D. The ERA5 Global Reanalysis. Q. J. R. Meteorol. Soc. 2020, 146, 1999–2049. [Google Scholar] [CrossRef]
  37. Konwar, M.; Parekh, A.; Goswami, B. Dynamics of East-west Asymmetry of Indian Summer Monsoon Rainfall Trends in Recent Decades. Geophys. Res. Lett. 2012, 39, 10. [Google Scholar] [CrossRef] [Green Version]
  38. Bavadekar, S.; Mooley, D. Computation of the Average Precipitation over the Western Part of Peninsular India during the Summer Monsoon from the Continuity Equation for Atmospheric Water Vapour. Tellus 1978, 30, 537–541. [Google Scholar] [CrossRef] [Green Version]
  39. Herman, R.; Cherry, J.; Young, J.; Welker, J.; Noone, D.; Kulawik, S.; Worden, J. Aircraft Validation of Aura Tropospheric Emission Spectrometer Retrievals of HDO/H 2 O. Atmos. Meas. Tech. 2014, 7, 3127–3138. [Google Scholar] [CrossRef] [Green Version]
  40. Worden, J.; Noone, D.; Galewsky, J.; Bailey, A.; Bowman, K.; Brown, D.; Hurley, J.; Kulawik, S.; Lee, J.; Strong, M. Estimate of Bias in Aura TES HDO/H2O Profiles from Comparison of TES and in Situ HDO/H 2 O Measurements at the Mauna Loa Observatory. Atmos. Chem. Phys. 2011, 11, 4491–4503. [Google Scholar] [CrossRef] [Green Version]
  41. Lee, J.; Worden, J.; Noone, D.; Bowman, K.; Eldering, A.; LeGrande, A.; Li, J.-L.; Schmidt, G.; Sodemann, H. Relating Tropical Ocean Clouds to Moist Processes Using Water Vapor Isotope Measurements. Atmos. Chem. Phys. 2011, 11, 741–752. [Google Scholar] [CrossRef] [Green Version]
  42. Konwar, M.; Das, S.; Deshpande, S.; Chakravarty, K.; Goswami, B. Microphysics of Clouds and Rain over the Western Ghat. J. Geophys. Res. Atmos. 2014, 119, 6140–6159. [Google Scholar] [CrossRef]
  43. Kumar, S.; Bhat, G. Vertical Structure of Orographic Precipitating Clouds Observed over South Asia during Summer Monsoon Season. J. Earth Syst. Sci. 2017, 126, 114. [Google Scholar] [CrossRef] [Green Version]
  44. Varikoden, H.; Harikumar, R.; Kumar, V.S.; Das, S.M.; Sampath, S.; Kumar, G.M. Properties of Cloud Base Height during Southwest Monsoon Period over a Tropical Station, Thiruvananthapuram. Curr. Sci. 2009, 96, 562–568. [Google Scholar] [CrossRef]
  45. Pan, L.; Munchak, L. Relationship of Cloud Top to the Tropopause and Jet Structure from CALIPSO Data. J. Geophys. Res. Atmos. 2011, 116, D12. [Google Scholar] [CrossRef]
  46. Zhang, W.; Xu, G.; Xi, B.; Ren, J.; Wan, X.; Zhou, L.; Cui, C.; Wu, D. Comparative Study of Cloud Liquid Water and Rain Liquid Water Obtained from Microwave Radiometer and Micro Rain Radar Observations over Central China during the Monsoon. J. Geophys. Res. Atmos. 2020, 125, e2020JD032456. [Google Scholar] [CrossRef]
  47. Srivastava, R.; Ramesh, R.; Rao, T.N. Stable Isotopic Differences between Summer and Winter Monsoon Rains over Southern India. J. Atmos. Chem. 2014, 71, 321–331. [Google Scholar] [CrossRef]
  48. Horita, J.; Wesolowski, D.J. Liquid-Vapor Fractionation of Oxygen and Hydrogen Isotopes of Water from the Freezing to the Critical Temperature. Geochim. Et Cosmochim. Acta 1994, 58, 3425–3437. [Google Scholar] [CrossRef]
  49. Craig, H. Isotopic Variations in Meteoric Waters. Science 1961, 133, 1702–1703. [Google Scholar] [CrossRef] [PubMed]
  50. Rahul, P.; Ghosh, P.; Bhattacharya, S. Rainouts over the Arabian Sea and Western Ghats during Moisture Advection and Recycling Explain the Isotopic Composition of Bangalore Summer Rains. J. Geophys. Res. Atmos. 2016, 121, 6148–6163. [Google Scholar] [CrossRef]
Figure 1. Long-term mean (over summer monsoon periods of 1979–2010) vertically integrated moisture flux (in kg/m.s; shaded) and moisture transport (in kg/m.s; vector) over a box (6–14° N; 74–78° E) surrounding Kozhikode (right diagram). Stable isotope observations from the GNIP station, Kozhikode (black-filled circle), are considered in this study. The location of the box in India is shown in the left diagram.
Figure 1. Long-term mean (over summer monsoon periods of 1979–2010) vertically integrated moisture flux (in kg/m.s; shaded) and moisture transport (in kg/m.s; vector) over a box (6–14° N; 74–78° E) surrounding Kozhikode (right diagram). Stable isotope observations from the GNIP station, Kozhikode (black-filled circle), are considered in this study. The location of the box in India is shown in the left diagram.
Atmosphere 14 01147 g001
Figure 2. Vertical variation in Rain Liquid Water Content (RLWC; marked by burgundy curve) and Cloud Liquid Water Content (CLWC; marked by blue curve) as measured by Zhang et al., 2020 [44].
Figure 2. Vertical variation in Rain Liquid Water Content (RLWC; marked by burgundy curve) and Cloud Liquid Water Content (CLWC; marked by blue curve) as measured by Zhang et al., 2020 [44].
Atmosphere 14 01147 g002
Figure 3. The average (June–September; 2005–2007) vertical δD (vapour) profiles obtained from TES, IsoGSM and LMDZ4 models. The methodology for constructing the TES profile is discussed in Section 4. The horizontal bars indicate combined temporal (summer monsoon of 2005–2007) and spatial variation (1σ standard deviation).
Figure 3. The average (June–September; 2005–2007) vertical δD (vapour) profiles obtained from TES, IsoGSM and LMDZ4 models. The methodology for constructing the TES profile is discussed in Section 4. The horizontal bars indicate combined temporal (summer monsoon of 2005–2007) and spatial variation (1σ standard deviation).
Atmosphere 14 01147 g003
Figure 4. A scatter plot showing δD bias against δ18O bias of rains simulated in IsoGSM and LMDZ4 models. The bias is estimated as model value-observed value; the rain isotope data at Kozhikode under Global Network of Isotope in Precipitation (GNIP) are considered as the observed values. The equations of mean regression lines (‘L’ subscript for LMDZ and ‘I’ subscript for IsoGSM) are also shown in the figure.
Figure 4. A scatter plot showing δD bias against δ18O bias of rains simulated in IsoGSM and LMDZ4 models. The bias is estimated as model value-observed value; the rain isotope data at Kozhikode under Global Network of Isotope in Precipitation (GNIP) are considered as the observed values. The equations of mean regression lines (‘L’ subscript for LMDZ and ‘I’ subscript for IsoGSM) are also shown in the figure.
Atmosphere 14 01147 g004
Figure 5. Frequency distribution of (a) observed rain δD values (b) model rain δD values (c) estimated liquid δD values at the cloud base height (CBH) and (d) estimated raindrop evaporation fraction (1-f). The green (orange) bars indicate IsoGSM (LMDZ4). The procedures for estimation of liquid δD at CBH and (1f) are discussed in Section 3.1 and Section 3.2, respectively.
Figure 5. Frequency distribution of (a) observed rain δD values (b) model rain δD values (c) estimated liquid δD values at the cloud base height (CBH) and (d) estimated raindrop evaporation fraction (1-f). The green (orange) bars indicate IsoGSM (LMDZ4). The procedures for estimation of liquid δD at CBH and (1f) are discussed in Section 3.1 and Section 3.2, respectively.
Atmosphere 14 01147 g005
Figure 6. A scatter plot showing δD against δ18O values of observed data and two model outputs. Isotope data at Kozhikode under Global Network of Isotope in Precipitation (GNIP) are considered as observed data. The equations of mean regression lines (‘L’ subscript for LMDZ, ‘I’ subscript for IsoGSM and ‘O’ subscript for observed data) are also shown in the figure.
Figure 6. A scatter plot showing δD against δ18O values of observed data and two model outputs. Isotope data at Kozhikode under Global Network of Isotope in Precipitation (GNIP) are considered as observed data. The equations of mean regression lines (‘L’ subscript for LMDZ, ‘I’ subscript for IsoGSM and ‘O’ subscript for observed data) are also shown in the figure.
Atmosphere 14 01147 g006
Figure 7. A scatter plot showing δD against δ18O values of vapour in IsoGSM and LMDZ models (a). The same scatter for liquid (accreted) at CBH for the two models is shown in (b). Each point in (a) corresponds to average value at a given level for all JJAS months during the study period. The equations of mean regression lines (‘L’ subscript for LMDZ and ‘I’ subscript for IsoGSM) are also shown in the figure.
Figure 7. A scatter plot showing δD against δ18O values of vapour in IsoGSM and LMDZ models (a). The same scatter for liquid (accreted) at CBH for the two models is shown in (b). Each point in (a) corresponds to average value at a given level for all JJAS months during the study period. The equations of mean regression lines (‘L’ subscript for LMDZ and ‘I’ subscript for IsoGSM) are also shown in the figure.
Atmosphere 14 01147 g007
Figure 8. Scatter plots showing rain δD bias vs. raindrop evaporation fraction (1-f) estimated from IsoGSM and LMDZ4 (a). The same plots for rain δ18O bias are shown in (b). The (1-f) values estimated from δ18O and δD values are cross-plotted for both models in (c). Red dashed lines indicate 1:1 ratio of the two fractions. The equations of mean regression lines (‘L’ subscript for LMDZ and ‘I’ subscript for IsoGSM) are also shown in the figure. Since model values are less than the observed GNIP values, the biases are negative in sign. More negative values imply more evaporation is needed to match the observed values.
Figure 8. Scatter plots showing rain δD bias vs. raindrop evaporation fraction (1-f) estimated from IsoGSM and LMDZ4 (a). The same plots for rain δ18O bias are shown in (b). The (1-f) values estimated from δ18O and δD values are cross-plotted for both models in (c). Red dashed lines indicate 1:1 ratio of the two fractions. The equations of mean regression lines (‘L’ subscript for LMDZ and ‘I’ subscript for IsoGSM) are also shown in the figure. Since model values are less than the observed GNIP values, the biases are negative in sign. More negative values imply more evaporation is needed to match the observed values.
Atmosphere 14 01147 g008
Table 1. The model and observed δD values at Kozhikode along with the estimates of evaporation.
Table 1. The model and observed δD values at Kozhikode along with the estimates of evaporation.
Months 1 (YYYY-MM)GNIPISOGSMLMDZ4
2 Observed Rain δD (‰)3 Model Rain δD (‰) 4 Estimated Rain δD at CBH (‰)5 Estimated Rain Evaporation Fraction 3 Model Rain δD (‰) 4 Estimated Rain δD at CBH (‰)5 Estimated Rain Evaporation Fraction
1997-06−7−23−470.27−11−1080.33
1997-07−14−16−610.18−33−1140.29
1997-08−3−13−450.32−27−1060.36
1997-09−40−19−920.123−900.10
1998-06−6−21−620.47−51−1550.31
1998-07−1−11−860.40−25−1030.47
1998-08−17−19−980.21−9−980.20
1998-09−8−19−970.29−13−1130.30
1999-06−5−14−340.28−20−1040.40
1999-07−5−19−700.35−28−1130.34
1999-084−12−540.49−16−890.52
1999-09−8NA−770.32NA−930.30
2000-060−16−470.48−22−1090.54
2000-071−25−570.44−23−1190.50
2000-08−9−15−910.31−16−970.24
2000-09−26−29−860.17−1−830.14
2001-064−16−570.52−26−1290.57
2001-07−2−8−640.41−50−1330.20
2001-08−1−17−650.46−24−830.27
2001-09−12−15−670.199−900.23
2002-06−39−19−440.02−16−1010.14
2002-071−11−200.43−19−960.43
2002-081−13−440.56−3−930.32
2002-092−16−620.416−590.30
2003-065−19−740.47−34−1120.51
2003-07−11−16−90 −37−1270.46
2003-083−17−640.44−8−910.47
2003-09−8−12−710.263−840.25
2004-06−10−14−420.21−15−1120.40
2004-07−2−15−690.50−34−1040.53
2004-083−15−460.49−25−1120.55
2004-09−26−24−880.17−19−1190.19
2005-06−9−18−540.22−30−980.32
2005-07−11−14−720.28−52−720.28
2005-08−4−15−610.33−7−840.27
2005-09−7−23−780.255−950.25
2006-06−7−15−750.35−20−1140.40
2006-07−3−11−340.26−33−930.38
2006-08−7−17−720.34−33−950.30
2006-09−30−15−910.16
2007-06−33−16−1060.18
2007-07−4−19−770.32
2007-08−10−13−860.27
2007−09−8 −900.34
Average−8−16−670.32−20−1020.34
1 Monthly composite/average data of isotopes (both observed and model) for the following months are calculated; 2 Monthly average rain δD data (1997–2007) of Kozhikode under the Global Network of Isotopes in Precipitation (GNIP) are given here; 3 Monthly average rain δD data extracted from the nudged models IsoGSM (1997–2007) and LMDZ4 (1997–2006) over the grid covering Kozhikode are given here. The data are available at SWING2 repository. 4 Liquid δD value at cloud base height is estimated in Section 3.1 using vapour δD values of the respective model; 5 Raindrop evaporation fraction (1-f) is estimated from liquid δD values at CBH and observed rain δD values (Section 3.2).
Table 2. The model and observed δ18O values at Kozhikode along with the estimates of evaporation.
Table 2. The model and observed δ18O values at Kozhikode along with the estimates of evaporation.
Months (YYYY-MM) 1GNIP ISOGSMLMDZ4
Observed Rain δ18O (‰) 2Model Rain δ18O (‰) 3Estimated Rain δ18O at CBH (‰) 4Estimated Rain Evaporation Fraction 5Model Rain δ18O (‰) 3Estimated Rain δ18O at CBH (‰) 4Estimated Rain Evaporation Fraction 5
1997-06−1.6−4.2−8.20.21−2.4−13.20.24
1997-07−3.3−3.4−9.60.12−5.2−14.60.19
1997-08−1.5−3.1−7.60.17−4.7−14.00.11
1997-09−6.7−3.6−13.30.09−0.9−12.60.08
1998-06−2.1−3.8−9.50.20−6.9−9.60.20
1998-07−1.1−2.5−12.70.24−4.3−13.40.30
1998-08−3.1−3.3−14.10.17−2.1−13.20.16
1998-09−2.2−3.6−14.30.20−2.6−15.20.22
1999-06−1.8−3.0−6.40.15−4.3−12.60.25
1999-07−2.0−3.4−10.80.23−4.8−14.00.22
1999-08−0.7−2.7−8.70.24−3.1−11.80.24
1999-09−2.3 −11.60.20 −12.70.21
2000-06−1.3−3.4−7.90.22−4.1−13.20.25
2000-07−1.2−4.3−9.10.22−3.9−9.70.27
2000-08−2.3−3.3−13.20.20−3.8−13.10.18
2000-09−4.4−5.0−12.80.14−0.8−11.50.11
2001-06−1.0−3.8−9.10.24−4.9−9.10.25
2001-07−1.5−2.6−9.90.20−7.2−17.30.27
2001-08−1.6−3.2−10.20.19−4.5−16.10.18
2001-09−2.8−2.7−10.10.140.0−11.80.16
2002-06−6.3−3.7−7.90.03−3.5−12.00.08
2002-07−0.9−2.6−4.60.16−3.8−11.50.24
2002-08−1.0−2.5−7.50.20−1.2−13.30.30
2002-09−1.0−3.1−9.90.23−0.3−10.10.22
2003-06−0.6−3.6−12.00.28−5.6−14.50.30
2003-07−2.7−2.5−13.80.28−6.0−12.60.15
2003-08−0.8−3.2−9.90.25−2.1−12.40.41
2003-09−2.6−2.6−11.20.17−0.7−12.10.17
2004-06−2.6−3.3−7.60.13−3.5−13.10.20
2004-07−1.6−3.2−10.30.24−5.6−12.00.19
2004-08−1.2−2.8−7.80.19−4.3−14.20.33
2004-09−4.4−4.2−13.40.14−3.1−16.00.17
2005-06−2.6−3.6−9.10.15−5.1−10.30.16
2005-07−2.7−3.3−11.00.17−7.6−15.20.18
2005-08−1.3−2.6−9.30.21−2.0−13.20.47
2005-09−2.1−3.9−11.80.18−0.7−12.80.18
2006-06−2.6−3.1−11.90.20−3.5−13.40.20
2006-07−1.6−2.6−6.50.14−5.5−12.80.21
2006-08−2.1−3.4−10.80.20−5.0−11.30.20
2006-09−4.8−3.2−13.20.13
2007-06−5.6−3.3−15.40.14
2007-07−1.8−3.5−11.80.19
2007-08−2.9−3.0−12.60.17
2007-09−2.3 −13.30.20
Average−2.3−3.3−10.40.18−3.7−12.90.22
1. Monthly composite/average data of isotopes (both observed and model) for the following months are studied, 2 Monthly average rain δ18O data (1997–2007) of Kozhikode under the Global Network of Isotopes in Precipitation (GNIP) are given here, 3 Monthly average rain δ18O data extracted from the nudged models IsoGSM (1997–2007) and LMDZ4 (1997–2006) over the grid covering Kozhikode are given here. The data are available at SWING2 repository, 4 Liquid δ18O values at cloud base height estimated in Section 3.1 using vapour δ18O values of the respective model, 5 Raindrop evaporation fraction (1-f) is estimated from liquid δ18O composition at CBH and observed rain δ18O values (Section 3.2).
Table 3. An example of the uncertainty calculation in the evaporation parameter (1-f) corresponding to the IsoGSM model for June 1997 (see Table 1, first row, fifth column).
Table 3. An example of the uncertainty calculation in the evaporation parameter (1-f) corresponding to the IsoGSM model for June 1997 (see Table 1, first row, fifth column).
VariableRangeValuesdiffRatio = Diff/RangeSigma TermsTotal 1σ Error
Temp (C)2960.24770.07860.01970.00154450.003690.060709
2980.2802
3000.3263 0.27 ± 0.06
RH (%)850.2738−0.0067−0.00171.122 × 10−5 (Table 1)
870.2691
890.2671
Roo)−440.26230.01050.00262.756 × 10−5
−460.2677
−480.2728
Ramb−840.23160.05840.01460.0008526
amb)−860.2566
−880.2900
Rff)−50.3087−0.0707−0.01770.0012496
−70.2691
−90.2380
Table 4. Comparison of the estimated rain isotope ratios at cloud base height (CBH) and the magnitude of the average evaporation fraction (2005–2007). For computation of the reference (1-f) fraction, vapor δD values at multiple levels obtained from TES, and temperature, specific humidity data from ERA-5 have been used. Similarly, the δD, temperature, and specific humidity values from the model simulations are used in corresponding calculations for the models. Data from GNIP are used as the observed value for the rain δD at the ground level.
Table 4. Comparison of the estimated rain isotope ratios at cloud base height (CBH) and the magnitude of the average evaporation fraction (2005–2007). For computation of the reference (1-f) fraction, vapor δD values at multiple levels obtained from TES, and temperature, specific humidity data from ERA-5 have been used. Similarly, the δD, temperature, and specific humidity values from the model simulations are used in corresponding calculations for the models. Data from GNIP are used as the observed value for the rain δD at the ground level.
ModelRain δD at CBH (‰)Observed Value (‰)Rain Evaporation Fraction
TES (as Reference)−74−390.08
IsoGSM−78−390.11
LMDZ4−112−390.15
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sengupta, S.; Bhattacharya, S.K.; Sunil, N.S.; Sonar, S. Quantifying Raindrop Evaporation Deficit in General Circulation Models from Observed and Model Rain Isotope Ratios on the West Coast of India. Atmosphere 2023, 14, 1147. https://doi.org/10.3390/atmos14071147

AMA Style

Sengupta S, Bhattacharya SK, Sunil NS, Sonar S. Quantifying Raindrop Evaporation Deficit in General Circulation Models from Observed and Model Rain Isotope Ratios on the West Coast of India. Atmosphere. 2023; 14(7):1147. https://doi.org/10.3390/atmos14071147

Chicago/Turabian Style

Sengupta, Saikat, Sourendra Kumar Bhattacharya, Nimya Sheena Sunil, and Sumit Sonar. 2023. "Quantifying Raindrop Evaporation Deficit in General Circulation Models from Observed and Model Rain Isotope Ratios on the West Coast of India" Atmosphere 14, no. 7: 1147. https://doi.org/10.3390/atmos14071147

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop