Next Article in Journal
Large-Scale Populus euphratica Distribution Mapping Using Time-Series Sentinel-1/2 Data in Google Earth Engine
Next Article in Special Issue
Review Analysis of Irrigation and Application of Remote Sensing in the Lower Mekong River Basin
Previous Article in Journal
Effect of Partially Melting Droplets on Polarimetric and Bi-Spectral Retrieval of Water Cloud Particle Size
Previous Article in Special Issue
Impacts of Shape Assumptions on Z–R Relationship and Satellite Remote Sensing Clouds Based on Model Simulations and GPM Observations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Multi-Model Ensemble Pattern Method to Estimate the Refractive Index Structure Parameter Profile and Integrated Astronomical Parameters in the Atmosphere

1
School of Environmental Science and Optoelectronic Technology, University of Science and Technology of China, Hefei 230026, China
2
Advanced Laser Technology Laboratory of Anhui Province, Hefei 230037, China
3
Key Laboratory of Atmospheric Optics, Anhui Institute of Optics and Fine Mechanics, Hefei Institutes of Physical Science (HFIPS), Chinese Academy of Sciences, Hefei 230031, China
4
Science Island Branch of Graduate School, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(6), 1584; https://doi.org/10.3390/rs15061584
Submission received: 15 February 2023 / Revised: 5 March 2023 / Accepted: 9 March 2023 / Published: 14 March 2023

Abstract

:
In this study, we devised a constraint method, called multi-model ensemble pattern (MEP), to estimate the refractive index structure parameter ( C n 2 ) profiles based on observational data and multiple existing models. We verified this approach against radiosonde data from field campaigns in China’s eastern and northern coastal areas. Multi-dimensional statistical evaluations for the C n 2 profiles and integrated astronomical parameters have proved MEP’s relatively reliable performance in estimating optical turbulence in the atmosphere. The correlation coefficients of MEP and measurement overall C n 2 in two areas are up to 0.65 and 0.76. A much higher correlation can be found for a single radiosonde profile. Meanwhile, the difference evaluation of integrated astronomical parameters also shows its relatively robust performance compared to a single model. The prowess of this reliable approach allows us to carry out regional investigation on optical turbulence features with routine meteorological data soon.

Graphical Abstract

1. Introduction

Optical turbulence (OT), caused by atmospheric inhomogeneities and fluctuations, is one of the most critical factors that limit the transmission and performance of imaging systems [1]. Researchers involved in light propagation in the atmosphere, especially laser physicists and astronomers, have been concerned with this issue for decades [2,3,4,5,6,7,8,9,10,11]. A turbulent atmosphere impacts light wave propagation in various aspects, such as phase changes and intensity fluctuations. These distortions lead to significant blurring, scintillations, broadening, arrival angle fluctuations, and laser beam wander [1,2,8]. Hence, parameterization and characterization of OT are essential for designing and operating photoelectric systems.
Among all the parameters assessing the influence on optoelectronic systems from the turbulent atmosphere, the refractive index structure parameter ( C n 2 ) is commonly used to characterize the optical turbulence in the atmosphere. The past decades have witnessed researchers’ efforts to measure, parameterize, and estimate C n 2 . Up to now, different techniques (direct or indirect) using optical or non-optical principles have developed to obtain C n 2 [12]. Among these techniques, a pair of micro-thermometers (MT) is the most common equipment used to obtain C n 2 by invoking several hypotheses [1]. Utilizing a balloon-borne MT (in situ measurements), usually accompanied by measurements of routine meteorological parameters, is extensively employed for getting the C n 2 profile in photoelectric applications, for example, site testing [6,13,14] and astronomical observatories routine scheduling [3,15]. Other remote sensing methods and instruments, for example, Multi-Aperture Scintillation Sensor (MASS), Slope Detection And Ranging (Slodar) and Solar Differential Image Motion Monitor+ (S-Dimm+), are also of vital importance for the development of modeling and the refinement of empirical dependencies for astronomy [16,17,18,19,20].
Meanwhile, methods parameterizing and estimating C n 2 profiles are established to meet the need of engineering practices. Empirical, physically-based, statistical, and data-driven learning methods to estimate C n 2 were subsequently developed. Simple empirical methods, such as the submarine laser communication (SLC) model [21], are only involved in a single elevation parameter. Physically-based models referring to thermodynamics or dynamics factors exist in lots of literature. Owing to their abundant physical connotations, these models are competitive in characterizing OT in terms of its physical mechanism. Hufnagel developed the Hufnagel model based on meteorology and stellar scintillation data [22]. The Hufnagel-Valley5/7 (HV5/7) model [22,23] is one of the most popular forms related to wind velocity in the free atmosphere. Ruggiero and DeBenedictis proposed the Hmnsp99 outer scale model, referring to gradients of temperature and wind shear [24]. Dewan developed a similar turbulence outer scale method utilizing wind shear [25]. Thorpe investigated the relationship between potential temperature inversion and the Thorpe scale; Basu proposed a simple approach to estimate C n 2 profiles with the coarse-resolution potential temperature profiles [21,26]. The Ellison scale was developed to quantify the scales of water body overturns. This theory was also used to calculate C n 2 [27,28]. Recently, several modified models, such as the wind shear and potential temperature (WSPT) model [29] and wind shear and temperature gradient (WSTG) model [30], were also applied to estimate C n 2 profiles under different experimental environments. Other methods were developed in a statistical view, for example, statistical models devised by Vanzandt [31] and Trinquet [32]. Along with the development of computer science, deep learning tools have shown their advantage in handling high-dimensional and nonlinear issues. Researchers also applied this useful tool in estimating C n 2 [33,34,35].
However, no one of the existing estimating approaches are superior to any of the others, to the best of our knowledge. Each existing approach has its own merits and limitations [21]. The universality and robustness of most existing approaches and models should be improved. However, the turbulent atmosphere with random, nonlinear, and infinite-element features makes it difficult to completely specify the precise mathematical expression of C n 2 from the routine macroscopic meteorological parameters—for now, at least. The existing physical-based approaches were established on several hypotheses and statistical evidence, more or less. Here, we propose a multi-model ensemble pattern (MEP) method to estimate C n 2 based on several existing physically-based methods. The purpose of this study is to take advantage of different existing approaches. The proposed model performance is not always the best. However, it can ensure that the C n 2 and integrated astronomical parameters estimated by the MEP are competitive compared to the best of the existing models if it is not.
This paper is organized as follows: Section 2 describes the experimental site, instruments, and radiosonde data. Section 3 presents the theory of several existing approaches to estimate C n 2 that we adopted and the proposed MEP method. Section 4.1 depicts the results of C n 2 using different models. Section 4.2 exhibits the evaluation of different models in calculating integrated astronomical parameters. The summary and conclusions are given in Section 5.

2. Experimental Principles and Scientific Data

2.1. Experimental Principles

According to the Gladstone law [12,36] and neglecting the water vapor concentration contribution, the refractive index structure parameter C n 2 (m 2 / 3 ) can be computed via pressure P (hPa), absolute temperature T (K), and temperature structure parameter C T 2 (K 2 m 2 / 3 ) as follows:
C n 2 = 79 × 10 6 P T 2 2 C T 2 .
C T 2 can be calculated by the temperature structure function D T 2 based on the Kolmogorov–Obukhov turbulence assumption [1]. D T 2 is defined as:
D T r = T x T x + r 2 = C T 2 r 2 / 3 l 0 r L 0 ,
where triangle brackets denote an ensemble average; x and x + r are the positions of the temperature probes; l 0 and L 0 represent the inner and outer scales, respectively; and r represents the distance of two probes that should be in the inertial sub-region. Radiosonde balloons equipped with micro-thermometers (MT) and routine meteorological sensors are used worldwide to obtain optical turbulence and meteorology parameters profiles.
In our case, the temperature probes (red rectangular boxes) used are shown in Figure 1b. The two platinum probes were isolated 1 m (r = 1 m) horizontally. T and P data necessary for calculating C n 2 were measured by onboard temperature and pressure sensors. A Global Positioning System (GPS) was used to obtain the position information, and wind velocity was calculated from GPS data with a precision of 0.3 m/s. The Anhui Institute of Optics and Fine Mechanics (AIOFM) designed the whole system. The instruments’ performance was summarized in Ref. [37]. The platinum wire probe resistance was 10 Ω with 10 μm diameter. The minimum detectable value of C T 2 was 4.0 × 10 6   K 2 m 2 / 3 . The sampling frequency of the processor was up to 100 Hz, and the data were averaged with a time interval of 1 s. The precision of the temperature and pressure sensors were 0.2 K and 1.5 hPa. The balloons ascend with a vertical velocity of approximately 5 m/s. The data were re-processed with a space interval of 10 m.

2.2. Scientific Data

The field observations were carried on two areas (Figure 1a) during April 2018. One observation was undertaken in the eastern coastal area of China (hereinafter ECACN), and the other in the northern coastal area of China (hereinafter NCACN). After removing several incomplete datasets with low termination altitude or missing data, we chose 16 and 20 profiles of ECACN and NCACN, respectively. The data collections of two areas are summarized in Table 1. More details are documented in Appendix B Table A1 and Table A2.

3. Methodology of MEP

3.1. Theory of the Adopted Models

Seven different approaches (HV: Hufnagel-Valley 5/7; H9: Hmnsp99; DN: Dewan; TE: Thorpe; EN: Ellison; WT: WSPT; WG: WSTG) estimating C n 2 with routine meteorological parameters were adopted in our study. We have summarized theories of these approaches in Appendix A to avoid interrupting the fluency of this article. More details can be found in the corresponding literature.
In data processing, all approaches except for HV involved gradient variables (the measured meteorological parameters or their derived parameters). Several approaches (TE, EN, and WT) calculated C n 2 related to the sizes of localized overturns of the potential temperature. It was hard to distinguish these overturns for coarse resolution data because potential temperature profiles have an increasing tendency with height most of the time. Hence, we adopted the original resolution data in these approaches. Meanwhile, the other approaches (H9, DN, and WSTG) were calculated in the vertical resolution of 60 m. All seven approach estimations were re-processed on the scale of 60 m for consistency and convenience. Meanwhile, data exceeding 1 km above the ground level (AGL) were selected. Hence, the feature of C n 2 and integrated astronomical parameters represent the free atmosphere results in our case.

3.2. MEP Method

Before introducing the principles of MEP, several theoretical basics should be elaborated first. For two variables, r and f, r n is the reference variable (MT measured C n 2 in this study), and f n is the corresponding pattern result (estimated as C n 2 in this study). The correlation coefficient (R) and their root-mean-square difference between two fields ( E , also known as the centered root-mean-square difference) are defined as:
R = 1 N n = 1 N f n f ¯ r n r ¯ σ f σ r ,
E = 1 N n = 1 N f n f ¯ r n r ¯ 2 1 / 2 ,
where σ r ( σ r = 1 / N n = 1 N r n r ¯ 2 ) and σ f ( σ f = 1 / N n = 1 N f n f ¯ 2 ) denote the reference variable standard deviation and pattern result standard deviation, respectively; r ¯ and f ¯ represent the average of two variables. Thus, we can deduce the relationship between the reference standard deviation σ r , pattern standard deviation σ f , and correlation coefficient R as:
E 2 = σ f 2 + σ r 2 2 σ f σ r R ,
Taylor devised the Taylor diagram to provide a concise statistical summary of how well the patterns match each other in terms of the above four statistics ( σ r , σ f , R and E ) [38]. In our study, we have normalized the statistics ( σ ^ f = σ f / σ r , σ ^ r = 1 , E ^ = E / σ r ) referring to σ r for convenience. According to Taylor’s work, a skill function was also developed to assess the models’ performance as follows:
S α , β = 2 α 1 + R β σ ^ f + 1 / σ ^ f α 1 + R 0 β ,
where R 0 represents the maximum of R in a set of the same model and we set R 0 = 1 ; α and β are penalty coefficients that can adjust the proportion of skill function via model variance and correlation coefficient. A more considerable value of α or β means that the corresponding statistic ( σ ^ or R) has a more significant influence on the result of S ( α , β ) .
Further, a weight function is defined as:
W j γ = S j γ i = 1 N S i γ .
Note that all skill values are in the range of 0–1. We set a penalty parameter γ to distinguish the model’s performance. Consequently, the multi-model ensemble pattern (MEP) method process is divided into three steps, as shown in Figure 2. We have summarized them as follows:
1.
Using routine meteorological parameters estimating C n 2 with multiple models;
2.
Obtaining models skills S ( α , β ) against MT results in Equation (6);
3.
Calculating weights W j ( γ ) of different models and MEP results.
Meanwhile, parameters ( α , β ) are used for modulating the weights of different statistics, and γ is used to distinguish the different models’ performance. These penalty parameters can be changed as the research focus changes in practice. For example, we can increase the α value to increase the weight of data fluctuation in the evaluation system. It is the same for the β for correlations, and we chose the latter condition in our case. Moreover, a considerable γ means a more significant influence on the evaluation of skills. In our case, we set α = 2 , β = 6 , γ = 4 .

3.3. Statistical Analysis

In addition to the correlation coefficient (R), the root mean square error ( R M S E ), bias ( B i a s ), and mean absolute error ( M A E ) were calculated to evaluate the performance of the different approaches. The definitions of these statistics are as follows:
R M S E = 1 N n = 1 N r n f n 2 ,
B i a s = 1 N n = 1 N r n f n ,
M A E = 1 N n = 1 N r n f n .

4. Results and Discussion

4.1. Measured and Estimated C n 2 Profiles

We employed 16 and 20 radiosonde datasets of two areas when evaluating the performance of the proposed and adopted methods. We used l o g 10 ( C n 2 ) instead of C n 2 to generate readable data and curves. We provide two C n 2 profiles of all approaches against the MT of each dataset in the primary text. See Appendix B for all days details in two sites (Figure A1, Figure A2, Figure A3 in ECACN; Figure A4, Figure A5, Figure A6, Figure A7 in NCACN).
Figure 3a,b displays MT and estimations C n 2 profiles from the ECACN radiosonde campaign. The overall trends of estimations are consistent with MT. The C n 2 magnitude of estimations and MT are mainly distributed in the range of 10 15 10 19 m 2 / 3 . Distinct differences can also be seen between different estimations. HV has a very high correlation with MT within the troposphere. However, it underestimates C n 2 significantly above approximately 20 km, which indicates that a more turbulent and complex atmospheric state might exist above the troposphere in this area. TE, EN, and WT have better performance in magnitude owing to the calibration of unknown proportionality constants according to MT measurements. H9, DN, and WG have a similar trend in the overall trend, while these estimations fluctuate a little bit more around the mean value against TE, EN, WT, and ME. By combining the corresponding Taylor diagrams in Figure 3c,d, we can also easily find that the values of normalized standard deviations of HV, H9, DE, and WG are much bigger than MT most of the time. Meanwhile, closer normalized standard deviation values to 1 (or MT) of TE, EN, and ME means that these approaches have similar behavior in C n 2 fluctuation magnitude. In addition, ME also shows its advantage in correlation evaluation. Among all 16 launches, correlation coefficients between ME and MT are mainly distributed around 0.6–0.8 and the best one is up to approximately 0.9.
Figure A1, Figure A2, Figure A3 in Appendix B display all the C n 2 profiles in ECACN. Scatter figures of all approaches against MT for all launches were plotted to further study the overall statistical features. Figure 4 shows all launches estimated C n 2 statistical feature in ECACN. The relevant statistics are summarized in Table 2. Although the overall B i a s of ME is slightly larger than WT, R, M A E and R M S E of ME present the best performance of all approaches.
Validation was also done to the radiosonde data from the campaign carried out in NCACN. Twenty sounding datasets were selected in this area. Figure 5 exhibits two launches C n 2 profiles and their corresponding Taylor diagrams. The characteristics of different approaches estimation profiles are similar to those in ECACN. MEP correlation coefficients of a single launch in NCACN are mainly distributed around 0.7–0.9, and the best one is more than 0.95. Figure A4, Figure A5, Figure A6, Figure A7 in Appendix B display all the C n 2 profiles in NCACN. Figure 6 shows all 20 launches estimated C n 2 against the MT statistical feature in NCACN. The relevant statistics are summarized in Table 3. The overall correlation criteria R of MEP is the best of all approaches, up to 0.7632. Meanwhile, the deviation criteria B i a s , M A E , and R M S E of MEP are the smallest. The above results of all the approaches in ECACN and NCACN have proved the potential of MEP in estimating C n 2 utilizing radiosonde data.

4.2. Integrated Astronomical Parameters from Measured and Estimated C n 2 Profiles

Evaluating the optical turbulence influence on optoelectronic facilities (ground-based observatories, laser transmission, and free atmosphere optical communication systems) in the atmosphere is one of the primary aims of researchers. Hence, we also calculated the integrated astronomical parameters (Fried parameter r 0 , seeing ϵ , isoplanatic angle θ A O and scintillation rate σ I 2 ) to evaluate the performance of the proposed method. These parameters are defined as [5,9,12]:
r 0 = 0.423 2 π λ 2 sec φ h 0 C n 2 h d h 3 / 5 ,
ε = 5.25 λ 1 / 5 h 0 C n 2 h d h 3 / 5 ,
θ A O = 0.057 λ 6 / 5 h 0 C n 2 h h 5 / 3 d h 3 / 5 ,
σ I 2 = 19.12 λ 7 / 6 h 0 C n 2 h h 5 / 6 d h .
φ is the solar zenith angle set as 0 ; λ is a given wavelength (we set λ = 550 nm); h denotes the elevation above ground level (AGL) of the sites; h 0 represents the initial elevation (we set h 0 = 1000 m). Therefore, the conclusions of the integrated astronomical parameters included in this study can only represent the influence of the free atmosphere.
Details of these integrated astronomical parameters of all launches in ECACN are listed in the Appendix B, Table A3, Table A4, Table A5, Table A6. The median values represent regional features and are of referential value for photoelectric applications. Median values of r 0 , ϵ , θ A O , and σ I 2 calculated from MT are 10.10 cm, 1.10 , 0.67 , and 0.54 , respectively. These parameters calculated from ME are 8.93 cm, 1.25 , 0.73 , and 0.56 . The relative errors of median values are rather small. All the integrated astronomical parameters are depicted in Figure 7, and the relevant statistical feature of these parameters are summarized in Table 4. HV and DN overestimated r 0 and θ A O and underestimated ϵ and σ I 2 can be easily found both from the figures and their B i a s from the table against MT. The ME correlation coefficients of R 0 , ϵ and σ I 2 are quite good. Meanwhile, the deviations are rather small compared to the other approaches.
The same computation process was done for the 20 radiosonde data from NCACN. The parameters calculated in NCACN are listed in the Appendix B, Table A7, Table A8, Table A9, Table A10. Median values of r 0 , ϵ , θ A O , and σ I 2 calculated from MT are 10.31 cm. 1.08 , 0.63 , and 0.66 , respectively. The parameters calculated from ME are 8.83 cm, 1.26 , 0.72 , and 0.58 . All the integrated astronomical parameters are portrayed in Figure 8, and the relevant statistical feature of these parameters are summarized in Table 5. The features of these parameters’ statistics are similar to the results in ECACN. The ME correlation coefficients of all parameters are even better in general compared to ECACN. Meanwhile, the deviation statistics are relatively small overall compared to the other approaches. A comprehensive comparison in two experimental areas between MT and the best estimations of all the parameter statistics were summarized in Table 6. Although MEP is not always the best estimation among these eight approaches, its gap with the optimal approach is minimal. All the above show MEP’s considerable universality in studying optical turbulence characteristics.

5. Conclusions

In this study we propose a multi-model ensemble pattern method to estimate the C n 2 based on several existing physical-based approaches. Balloon radiosonde data were collected in two areas of China to validate this method. Multiple dimensions evaluation including C n 2 and the integrated astronomical parameters ( r 0 , ϵ , θ A O and σ I 2 ) of all approaches were done against the MT measured results. Statistical analysis of these methods’ performance mainly focuses on the overall trend (R) and deviation (Bias, MAE, and RMSE). The best performance of all approaches against the MEP is summarized in Table 6. The C n 2 correlation coefficients of MEP are up to 0.65 and 0.76. The overall agreements of the C n 2 profiles in two areas are quite good. A single profile has an even higher correlation coefficient of more than 0.95. Several statistical assessments of deviations of C n 2 are relatively small. These indicate that MEP has the capacity to estimate C n 2 well. Meanwhile, the evaluations of integrated astronomical parameters also show its promising potential for calculating these parameters as C n 2 does. Although MEP was not always the best method in all parameters statistical evaluations, it showed competitive performance in these evaluations. Hence, the MEP method presented good stability and universality, and even the validation radiosonde data were collected in different areas, which meant significantly different atmospheric conditions. The MEP appreciably contains more information than a single method, including thermodynamic and dynamic factors of the optical turbulence. Moreover, the MEP method could be less sensitive to different parametric settings caused by each method, producing a more robust C n 2 estimate.
It should be noted that a single approach performed relatively well after well-designing the relevant parameters according to field radiosonde measurement in previous practice [21,28,29,30]. However, the designed parameters might be less effective for other sites. This weakness makes a single model challenging to extend without sufficient prior data. Reliable and universal methods estimating C n 2 from routine meteorological parameters are critical to evaluate the optical turbulence influence on adaptive optics systems. The most obvious example is a forecasting study in which the astronomer can not obtain optical turbulence in advance directly. Generally, researchers can forecast C n 2 via weather forecasting models, combining different estimating approaches [8,12,39,40]. In addition, it also provides us with an applicable method to study regional optical turbulence characteristics from historical meteorological data. To be certain, more validation work should be done up until that point.

Author Contributions

H.Z. ((Hanjiu Zhang): methodology, software, data analysis and writing—draft preparation; L.Z.: software, data analysis, and editing; G.S.: funding acquisition, guidance, data collection, data curation and writing—review; K.Z., Y.L., X.M. and H.Z. (Haojia Zhang): investigation and editing; Q.L.: instrument design and manufacture; S.C., T.L. and X.L.: funding acquisition, experiment guarantee, and data collection; N.W.: guidance and writing—review. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National High-tech Research and Development Program (No. E23D0HA65S2).

Data Availability Statement

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

Acknowledgments

The authors express their thanks to all AIOFM colleagues carrying out the experiments. The authors would like to thank the reviewers for their remarks and suggestions on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
MEP or MEMulti-model Ensemble Pattern
OTOptical Turbulence
MTMicro-thermometers
SLCSubmarine Laser Communication
ECACNEastern coastal area of China
NCACNNorthern coastal area of China
MASSMulti-Aperture Scintillation Sensor
SlodarSlope Detection And Ranging
S-Dimm+Solar Differential Image Motion Monitor+
GPSGlobal Positioning System
AGLAbove the ground level
HV5/7 or HVHufnagel-Valley5/7 Model
H9Hmnsp99 Model
DNDewan Model
TEThorpe Model
ENEllison Model
WTWSPT Model
WGWSTG Model
AIOFMAnhui Institute of Optics and Fine Mechanics

Appendix A. Theories of Adopted Approaches

Appendix A provides theories of adopted approaches estimating C n 2 . We place the contents of Section 3.1 (primary manuscript) in this appendix document to avoid interrupting the fluency of the original article. In Section 3 (primary manuscript), we introduce several adopted physical-based approaches to estimate C n 2 . Theories of these approaches are briefly summarized below in the appendix document. For more details, one may refer to the corresponding literature. We roughly classify these approaches into three types. The first type is the semi-physical and semi-statistical approach. The typical approach is the Hufnagel model. The second type focuses on the calculation of the turbulence outer scale. We adopted the Hmnsp99, Dewan, and wind shear and temperature gradient (WSTG) models. The third type estimates C n 2 via the temperature structure parameter. The Thorpe, Ellison, and wind shear and potential temperature (WSPT) models are examples.

Appendix A.1. Hufnagel-Valley 5/7 Model

The Hufnagel-Valley model was developed based on the statics of radio sounding and stellar scintillation [22,23]. The most commonly used form is the Hufnagel-Valley 5/7 model related to wind speed. The calculated formula is expressed as:
C n 2 h = 8.2 × 10 26 W h 10 e h + 2.7 × 10 16 e h / 1.5 + A e h / 0.1 ,
where e is the Euler number; h (unit: m) is the height above the ground. In our case, the parameters A = 1.7 × 10 4 and W = 1 / 15 5 20 V 2 h d h , where V is the wind speed (unit: m/s) between 5 and 20 km.

Appendix A.2. The Outer-Scale Method

Other approaches (Hmnsp99, Dewan, WSTG) estimating C n 2 are based on the outer scale according to the work by Tatarskii [1]:
C n 2 = 2.8 L 0 4 / 3 M 2 ,
where L 0 is the turbulence outer scale and M is the vertical gradient of the potential refractive index. The value of M 2 can be calculated from the temperature (T in K) and pressure (P in hPa) profiles as below:
M 2 = 79 × 10 6 P T 2 d θ d h 2 ,
where θ (unit: K) is the potential temperature defined as θ = T 1000 / P 0.286 . Hence, the key to the C n 2 estimation becomes the calculation of the outer scale. Most methods that parameterize the outer scale or turbulence with macroscopic quantities rely not only on the existing theoretical basis of turbulence but also on the statistics of large amounts of experiment and numerical simulation data and physical intuition and perspicacity of the founders of these models. The adopted approaches in our study calculating L 0 are listed below.

Appendix A.2.1. Hmnsp99 Model

The Hmnsp99 model defines L 0 with wind shear (S) and temperature gradients ( d T / d h ) [24]. The expressions are different in the troposphere and stratosphere as
L 0 4 / 3 = 0.1 4 / 3 × 10 0.362 + 16.728 S 192.347 d T d h , T r o p o s p h e r e 0.1 4 / 3 × 10 0.757 + 13.819 S 57.784 d T d h , S t r a t o s p h e r e
where S = d u / d h 2 + d v / d h 2 (hereinafter), u and v are the north and east horizontal wind components, respectively.

Appendix A.2.2. Dewan Model

Dewan model deduces L 0 from only one parameter (wind shear) [25]. Meanwhile, it is similar to the Hmnsp99 model, which has a different form in the troposphere and stratosphere:
L 0 4 / 3 = 0.1 4 / 3 × 10 1.64 + 42 S , T r o p o s p h e r e 0.1 4 / 3 × 10 0.506 + 50 S . S t r a t o s p h e r e

Appendix A.2.3. WSTG Model

The WSTG model is a modified model that comes from Hmnsp99. The calculation of the outer scale is related to the dynamic and thermodynamic state of the atmosphere [30]. The expression is as follows:
L 0 4 / 3 = 0.1 4 / 3 × 10 0.835 37.464 S 306.034 d T d h , S < 0.016 d T / d h < 0 0.1 4 / 3 × 10 0.825 + 66.9 S 52.783 d T d h , S < 0.016 d T / d h > 0 0.1 4 / 3 × 10 0.715 + 52.907 S 102.515 d T d h , S > 0.016 d T / d h < 0 0.1 4 / 3 × 10 2.215 9.882 S 101.666 d T d h , S > 0.016 d T / d h > 0

Appendix A.3. The Temperature Structure Parameter Method

The remaining approaches (Thorpe, Ellison, WSPT) deduce C n 2 from the Gladstone relationship [12] as follows:
C n 2 = 79 × 10 6 P T 2 2 C T 2 ,
and the temperature structure parameter ( C T 2 ) expressed as:
C T 2 = c 0 L 0 4 / 3 θ ¯ h 2 .
c 0 is a constant that should be determined by experiment. c 1 , c 2 , and c 3 are also unknown proportionality constants determined by experiment data in the following Equations (A9), (A11), and (A13).

Appendix A.3.1. Thorpe Model

The Thorpe model quantifies C T 2 with the Thorpe scale ( L T ) and sorted potential temperature gradients ( θ s / h ) [26] as follows:
C T 2 = c 1 L T 4 / 3 θ s h 2 ,
L T = h o r i g i n a l h s o r t e d o r h o r i g i n a l h s o r t e d .
θ s (hereinafter) is the sorted potential temperature rearranged in ascending order; h o r i g i n a l and h s o r t e d are the corresponding heights of the original potential temperature and sorted potential temperature, respectively. In our case, we chose the latter formula of Equation (A10) to calculate L T .

Appendix A.3.2. Ellison Model

Ellison proposed the Ellison scale, which refers to density or potential temperature, to study the overturning of fluid caused by turbulences [27,28]. The calculation formula is as follows:
C T 2 = c 2 L E 4 / 3 θ s h 2 ,
L E = Δ θ θ s / h .
L E is the Ellison scale; Δ θ (hereinafter) represents the difference value of the original and sorted (ascending) potential temperature.

Appendix A.3.3. WSPT Model

The WSPT model involves both the wind speed and potential temperature information, calculating C T 2 [29] as follows:
C T 2 = c 3 L W 4 / 3 θ s h 2 ,
L W = Δ θ θ s / h · u v S 2 1 / 2 .

Appendix B. Details of Two Areas Radiosonde, Models Estimations, and Integrated Astronomical Parameters Results

This Appendix B provides the details of two areas radiosonde, models estimations, and integrated astronomical parameters results. Variables, symbols, and abbreviations used in this document have the same meanings as the primary manuscript.

Appendix B.1. Radiosonde Details

Appendix B.1.1. ECACN

Radiosonde details of the eastern coastal area of China (ECACN) are included in Table A1.
Table A1. ECACN radiosonde details (BJT: Beijing time).
Table A1. ECACN radiosonde details (BJT: Beijing time).
SiteFlight NumberDateRelease TimeFlight DurationTermination Altitude
   (BJT)/s(AGL)/m
 15 Apirl 201819:30501029,020
 29 Apirl 201819:30502731,210
 310 Apirl 201819:30476730,410
 411 Apirl 201819:30478827,880
 512 Apirl 201819:30501429,810
 615 Apirl 201819:30408825,740
 717 Apirl 201819:30435526,410
 818 Apirl 201819:30468628,150
ECACN919 Apirl 201819:30503730,120
 1020 Apirl 201819:30537131,620
 1122 Apirl 201819:30482028,220
 1224 Apirl 201819:30517630,710
 1325 Apirl 201819:30505130,910
 1426 Apirl 201819:30508829,410
 1527 Apirl 201819:30544331,750
 1628 Apirl 201819:30514431,170

Appendix B.1.2. NCACN

Radiosonde details of the northern coastal area of China (NCACN) are included in Table A2.
Table A2. NCACN radiosonde details.
Table A2. NCACN radiosonde details.
SiteFlight NumberDateRelease TimeFlight DurationTermination Altitude
   (BJT)/s(AGL)/m
 13 Apirl 201819:30446428,660
 24 Apirl 20187:30421027,970
 34 Apirl 201819:30474729,460
 45 Apirl 20187:30480829,320
 58 Apirl 201819:30427127,370
 69 Apirl 20187:30485528,880
 79 Apirl 201819:30478029,660
 810 Apirl 201819:30527529,780
912 Apirl 20187:30459127,810
 1013 Apirl 20187:30463328,710
NCACN1114 Apirl 201819:30506929,680
 1216 Apirl 20187:30536029,380
 1316 Apirl 201819:30529230,050
 1417 Apirl 20187:30517628,850
 1520 Apirl 20187:30515529,660
 1621 Apirl 201819:30485329,400
 1725 Apirl 201819:30501230,750
 1826 Apirl 20187:30471428,790
 1926 Apirl 201819:30490130,660
 2027 Apirl 201819:30479828,530

Appendix B.2. The Refractive Index Structure Parameter of MT and Estimations

Appendix B.2.1. ECACN MT and Models Estimations

The refractive structure index parameter profiles of MT and estimations in ECACN are exhibited in Figure A1, Figure A2, Figure A3.
Figure A1. The refractive index structure parameter profiles of MT and estimations in ECACN: Figure A1 sub-figures (ad) are flight numbers 1–4 in Table A1.
Figure A1. The refractive index structure parameter profiles of MT and estimations in ECACN: Figure A1 sub-figures (ad) are flight numbers 1–4 in Table A1.
Remotesensing 15 01584 g0a1
Figure A2. The refractive index structure parameter profiles of MT and estimations in ECACN: Figure A2 sub-figures (af) are flight numbers 5–10 in Table A1.
Figure A2. The refractive index structure parameter profiles of MT and estimations in ECACN: Figure A2 sub-figures (af) are flight numbers 5–10 in Table A1.
Remotesensing 15 01584 g0a2
Figure A3. The refractive index structure parameter profiles of MT and estimations in ECACN: Figure A3 sub-figures (af) are flight numbers 11–16 in Table A1.
Figure A3. The refractive index structure parameter profiles of MT and estimations in ECACN: Figure A3 sub-figures (af) are flight numbers 11–16 in Table A1.
Remotesensing 15 01584 g0a3

Appendix B.2.2. NCACN MT and Models Estimations

The refractive structure index parameter profiles of MT and estimations in NCACN are exhibited in Figure A4, Figure A5, Figure A6, Figure A7.
Figure A4. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A4 sub-figures (ad) are flight numbers 1–4 in Table A2.
Figure A4. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A4 sub-figures (ad) are flight numbers 1–4 in Table A2.
Remotesensing 15 01584 g0a4
Figure A5. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A5 sub-figures (af) are flight numbers 5–10 in Table A2.
Figure A5. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A5 sub-figures (af) are flight numbers 5–10 in Table A2.
Remotesensing 15 01584 g0a5
Figure A6. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A6 sub-figures (af) are flight numbers 11–16 in Table A1.
Figure A6. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A6 sub-figures (af) are flight numbers 11–16 in Table A1.
Remotesensing 15 01584 g0a6
Figure A7. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A7 sub-figures (ad) are flight numbers 17–20 in Table A2.
Figure A7. The refractive index structure parameter profiles of MT and estimations in NCACN: Figure A7 sub-figures (ad) are flight numbers 17–20 in Table A2.
Remotesensing 15 01584 g0a7

Appendix B.3. The Integrated Astronomical Parameters

Appendix B.3.1. ECACN Integrated Astronomical Parameters Details

The integrated astronomical parameters details calculated from radiosonde and model results in ECACN are included in Table A3, Table A4, Table A5, Table A6.
Table A3. ECACN integrated astronomical parameters details ( r 0 @ λ = 550 nm).
Table A3. ECACN integrated astronomical parameters details ( r 0 @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 16.4515.415.9519.314.254.0810.485.925.30
 213.1514.386.5221.057.3311.3211.796.679.05
 321.2216.914.8022.648.7215.6812.406.9810.10
 410.4717.374.5120.138.1315.2412.735.647.78
 55.6318.976.4320.467.5013.7610.946.349.34
 69.7212.373.7720.697.0513.109.986.339.03
 710.6312.534.5120.727.2814.4711.046.647.56
 88.6615.114.5721.327.6414.3012.027.149.65
 915.6119.535.6821.357.6914.9316.866.929.21
r 0 /cm1023.2820.116.2323.698.2115.3914.217.2211.84
 1118.6919.127.5221.148.0114.6912.237.039.71
 126.4316.282.3120.377.2114.479.766.788.82
 139.1915.404.0822.027.3814.279.586.828.06
 148.1214.947.0618.546.056.9711.017.197.47
 1512.9013.446.3619.827.4114.8212.436.708.66
 169.7213.125.5718.497.1112.1412.346.788.27
 Median10.1015.415.6320.707.3914.3811.916.788.93
 Mean11.8715.945.3720.737.3113.1011.866.698.74
Table A4. ECACN integrated astronomical parameters details ( ϵ @ λ = 550 nm).
Table A4. ECACN integrated astronomical parameters details ( ϵ @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 11.720.721.870.582.612.731.061.882.10
 20.850.771.710.531.520.980.941.671.23
 30.520.662.320.491.280.710.901.591.10
 41.060.642.470.551.370.730.871.971.43
 51.970.591.730.541.480.811.021.751.19
 61.140.902.950.541.580.851.111.761.23
 71.050.892.470.541.530.771.011.671.47
 81.280.742.430.521.460.780.921.561.15
 90.710.571.960.521.450.740.661.611.21
ϵ / 100.480.551.790.471.350.720.781.540.94
 110.590.581.480.531.390.760.911.581.15
 121.730.684.810.551.540.771.141.641.26
 131.210.722.720.501.510.781.161.631.38
 141.370.741.570.601.841.591.011.551.49
 150.860.831.750.561.500.750.891.661.28
 161.140.851.990.601.560.920.901.641.34
 Median1.100.721.980.541.500.770.931.641.25
 Mean1.110.712.250.541.560.960.961.671.31
Table A5. ECACN integrated astronomical parameters details ( θ A O @ λ = 550 nm).
Table A5. ECACN integrated astronomical parameters details ( θ A O @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 10.331.230.691.690.570.990.840.570.69
 20.781.090.471.790.571.130.680.600.74
 32.011.470.311.890.621.200.700.580.77
 40.831.550.281.770.601.140.700.550.68
 50.521.910.421.740.591.020.640.550.73
 60.570.870.381.760.601.080.680.560.73
 70.850.880.311.670.601.180.740.550.54
 80.691.180.391.750.581.160.710.610.77
 91.362.070.461.770.581.161.000.610.75
θ A O / 101.132.270.461.900.601.160.880.600.91
 111.101.950.471.770.591.120.760.600.73
 120.471.360.321.750.611.170.680.630.80
 130.461.230.471.860.591.140.670.630.66
 140.401.160.681.580.480.530.740.590.58
 150.650.980.611.730.581.150.830.620.71
 160.600.940.491.600.570.970.740.560.66
 Median0.671.230.461.760.591.140.720.600.73
 Mean0.801.380.451.750.581.080.750.590.72
Table A6. ECACN integrated astronomical parameters details ( σ I 2 @ λ = 550 nm).
Table A6. ECACN integrated astronomical parameters details ( σ I 2 @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 11.540.240.760.151.100.750.410.990.81
 20.400.281.040.130.800.300.490.870.54
 30.100.192.020.120.660.240.470.850.49
 40.440.172.440.140.730.260.461.030.67
 51.030.131.220.140.780.310.560.970.55
 60.630.401.950.140.820.310.570.950.58
 70.370.392.100.140.790.260.490.950.86
 80.610.251.610.140.780.260.490.820.51
 90.210.121.120.130.770.260.270.820.54
σ I 2 100.180.111.080.120.720.250.340.820.38
 110.220.130.990.140.750.270.440.840.54
 121.010.213.180.140.780.260.570.820.52
 130.820.241.440.120.780.270.580.810.66
 141.020.260.720.171.181.050.490.840.85
 150.470.330.870.140.820.260.400.840.60
 160.630.351.160.160.840.370.460.920.67
 Median0.540.241.190.140.780.260.480.850.56
 Mean0.610.241.480.140.820.350.470.880.61

Appendix B.3.2. NCACN Integrated Astronomical Parameters Details

The integrated astronomical parameters details calculated from radiosonde and model results in NCACN are included in Table A7, Table A8, Table A9, Table A10.
Table A7. NCACN integrated astronomical parameters details ( r 0 @ λ = 550 nm).
Table A7. NCACN integrated astronomical parameters details ( r 0 @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 12.9517.183.7419.166.848.2112.005.217.12
 28.6314.053.5720.757.3613.0110.206.158.27
 39.0414.563.8521.197.5014.5310.756.648.16
 45.6713.402.6720.686.908.999.815.777.40
 54.9312.473.9617.867.7915.888.426.198.03
 615.1113.974.8618.127.8214.039.475.818.84
 77.1915.955.8218.417.7616.7810.386.205.66
 819.4610.793.8619.559.4117.6015.126.5710.77
 99.8610.724.1620.768.9415.7310.186.4010.73
r 0 103.3214.171.6120.217.8513.0211.186.466.27
 1113.1711.673.9919.277.6014.789.666.577.11
 1215.6715.555.1617.367.8415.6512.095.939.40
 1313.3518.656.1720.628.1616.5014.616.919.29
 147.0618.804.7720.027.4211.0213.746.248.79
 1510.7721.234.9221.438.1912.1912.686.119.73
 1612.6720.453.7424.388.5614.9714.646.848.83
 1716.2313.346.6120.658.5016.8412.936.829.93
 1813.0313.836.0021.468.3514.3212.106.3110.11
 1915.0613.406.6020.388.8616.9013.657.109.62
 206.0618.605.2221.958.4316.2413.986.7510.36
 Median10.3114.114.4620.507.8414.8812.046.368.83
 Mean10.4615.144.5620.218.0014.3611.886.358.72
Table A8. NCACN integrated astronomical parameters details ( ϵ @ λ = 550 nm).
Table A8. NCACN integrated astronomical parameters details ( ϵ @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 13.770.652.970.581.631.350.932.131.56
 21.290.793.120.541.510.851.091.811.34
 31.230.762.880.521.480.761.031.671.36
 41.960.834.170.541.611.241.131.931.50
 52.260.892.810.621.430.701.321.801.39
 60.740.802.290.611.420.791.171.911.26
 71.550.701.910.601.430.661.071.791.97
 80.571.032.880.571.180.630.741.691.03
 91.131.042.670.541.240.711.091.741.04
ϵ / 103.350.786.920.551.420.850.991.721.77
 110.840.952.790.581.460.751.151.691.56
 120.710.712.160.641.420.710.921.871.18
 130.830.601.800.541.360.670.761.611.20
 141.570.592.330.561.501.010.811.781.26
 151.030.522.260.521.360.910.881.821.14
 160.880.542.970.461.300.740.761.621.26
 170.690.831.680.541.310.660.861.631.12
 180.850.801.850.521.330.780.921.761.10
 190.740.831.680.551.250.660.811.571.16
 201.830.602.130.511.320.680.801.651.07
 Median1.080.792.500.541.420.750.921.751.26
 Mean1.390.762.710.551.400.810.961.761.31
Table A9. NCACN integrated astronomical parameters details ( θ A O @ λ = 550 nm).
Table A9. NCACN integrated astronomical parameters details ( θ A O @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 10.491.520.351.820.620.970.910.540.72
 20.641.050.621.760.631.050.730.590.72
 30.611.110.361.980.661.230.690.630.71
 40.470.970.331.840.590.800.680.540.65
 50.410.880.241.860.631.300.550.570.67
 61.021.040.311.560.621.110.560.460.67
 70.351.310.561.630.631.380.680.580.53
 81.260.720.211.630.641.300.880.560.78
 90.510.710.261.710.651.190.630.570.75
θ A O / 100.151.060.191.810.620.960.680.530.59
 110.970.800.291.860.641.160.650.610.59
 120.931.250.371.650.611.260.740.500.73
 131.021.830.391.960.611.280.780.650.77
 140.381.870.271.930.590.950.830.570.66
 150.512.750.751.800.630.920.790.580.76
 161.452.400.412.060.681.230.990.650.77
 171.290.970.451.770.631.310.660.590.77
 180.741.020.441.890.621.100.710.650.75
 191.060.970.431.700.621.280.770.590.73
 200.611.820.361.930.661.290.740.630.87
 Median0.631.060.361.820.631.210.720.580.72
 Mean0.741.300.381.810.631.150.730.580.71
Table A10. NCACN integrated astronomical parameters details ( σ I 2 @ λ = 550 nm).
Table A10. NCACN integrated astronomical parameters details ( σ I 2 @ λ = 550 nm).
 Flight    Method    
ParameterNumberMTHVH9DNTEENWTWGME
 12.280.181.940.140.800.500.361.160.69
 20.630.301.330.140.740.320.500.940.60
 30.690.271.900.120.690.250.520.820.62
 41.190.332.850.140.850.540.581.060.74
 51.660.392.960.150.720.220.790.990.68
 60.290.302.070.170.730.290.721.260.64
 71.380.220.990.160.730.200.530.971.10
 80.190.533.690.160.640.210.330.960.49
 90.790.542.830.140.650.250.630.930.50
σ I 2 105.370.298.170.140.730.350.521.010.96
 110.360.452.420.140.740.260.620.880.85
 120.310.231.620.170.750.230.471.190.56
 130.320.141.400.130.720.220.400.800.54
 141.320.142.490.130.810.410.380.990.65
 150.740.090.770.130.680.390.410.960.51
 160.230.101.580.110.610.240.290.790.54
 170.220.341.180.140.690.210.510.910.52
 180.450.311.250.130.720.290.480.860.52
 190.270.331.260.150.690.220.410.860.57
 200.850.141.660.120.640.220.410.840.44
 Median0.660.301.780.140.720.250.490.950.58
 Mean0.980.282.220.140.720.290.490.960.64

References

  1. Tatarskii, V.I. Wave Propagation in a Turbulent Medium; McGraw-Hill: New York, NY, USA, 1961. [Google Scholar]
  2. Beckers, J.M. Adaptive optics for astronomy: Principles, performance, and applications. Annu. Rev. Astron. Astrophys. 1993, 31, 13–62. [Google Scholar] [CrossRef]
  3. Coulman, C.E. Fundmental and applied aspects of astronomical seeing. Annu. Rev. Astron. Astrophys. 1985, 23, 19–57. [Google Scholar] [CrossRef]
  4. Davies, R.; Kasper, M. Adaptive Optics for Astronomy. Annu. Rev. Astron. Astrophys. 2012, 50, 305–351. [Google Scholar] [CrossRef] [Green Version]
  5. Fried, D.L. Limiting resolution looking down through the atmosphere. J. Opt. Soc. Am. 1966, 56, 1380–1385. [Google Scholar] [CrossRef]
  6. Deng, L.; Yang, F.; Chen, X.; He, F.; Liu, Q.; Zhang, B.; Zhang, C.; Wang, K.; Liu, N.; Ren, A.; et al. Lenghu on the Tibetan Plateau as an astronomical observing site. Nature 2021, 596, 353. [Google Scholar] [CrossRef] [PubMed]
  7. Fugate, R.Q.; Fried, D.L.; Ameer, G.A.; Boeke, B.R.; Browne, S.L.; Roberts, P.H.; Ruane, R.E.; Tyler, G.A.; Wopat, L.M. Measurement of atmospheric wave-front distortion using scatterd-light from a laser guide-star. Nature 1991, 353, 144–146. [Google Scholar] [CrossRef]
  8. Mahalov, A.; Moustaoui, M. Characterization of atmospheric optical turbulence for laser propagation. Laser Photonics Rev. 2010, 4, 144–159. [Google Scholar] [CrossRef]
  9. Roddier, F. V the Effects of Atmospheric Turbulence in Optical Astronomy. In Progress in Optics; Wolf, E., Ed.; Elsevier: Amsterdam, The Netherlands, 1981; Volume 19, pp. 281–376. [Google Scholar] [CrossRef]
  10. Zhang, R.Z.; Hu, N.Z.; Zhou, H.B.; Zou, K.H.; Su, X.Z.; Zhou, Y.Y.; Song, H.Q.; Pang, K.; Song, H.; Minoofar, A.; et al. Turbulence-resilient pilot-assisted self-coherent free-space optical communications using automatic optoelectronic mixing of many modes. Nat. Photonics 2021, 15, 743–750. [Google Scholar] [CrossRef]
  11. Zhu, Z.; Janasik, M.; Fyffe, A.; Hay, D.; Zhou, Y.; Kantor, B.; Winder, T.; Boyd, R.W.; Leuchs, G.; Shi, Z. Compensation-free high-dimensional free-space optical communication using turbulence-resilient vector beams. Nat. Commun. 2021, 12, 1666. [Google Scholar] [CrossRef]
  12. Masciadri, E.; Vernin, J.; Bougeault, P. 3D mapping of optical turbulence using an atmospheric numerical model I. A useful tool for the ground-based astronomy. Astron. Astrophys. Suppl. Ser. 1999, 137, 185–202. [Google Scholar] [CrossRef] [Green Version]
  13. Marks, R.D.; Vernin, J.; Azouit, M.; Briggs, J.W.; Burton, M.G.; Ashley, M.C.B.; Manigault, J.F. Antarctic site testing–Microthermal measurements of surface-layer seeing at the South Pole. Astron. Astrophys. Suppl. Ser. 1996, 118, 385–390. [Google Scholar] [CrossRef]
  14. Qing, C.; Wu, X.; Li, X.; Luo, T.; Su, C.; Zhu, W. Mesoscale optical turbulence simulations above Tibetan Plateau: First attempt. Opt. Express 2020, 28, 4571–4586. [Google Scholar] [CrossRef] [PubMed]
  15. Guryanov, A.E.; Kallistratova, M.A.; Kutyrev, A.S.; Petenko, I.V.; Shcheglov, P.V.; Tokovinin, A.A. The contribution of the lower atmospheric layers to the seeing at some mountain observatories. Astron. Astrophys. 1992, 262, 373–381. [Google Scholar]
  16. Tokovinin, A.; Kornilov, V. Measuring turbulence profile from scintillations of single stars. In Astronomical Site Evaluation in the Visible and Radio Range Workshop, IAU Technical Workshop, Proceedings of a Workshop Held at Cadi Ayyad University, Marrakech, Morocco, 13–17 November 2000; Astronomical Society of the Pacific Conference Series; Astronomical Soc Pacific: San Francisco, CA, USA, 2002; Volume 266, pp. 104–112. [Google Scholar]
  17. Osborn, J.; Wilson, R.; Butterley, T.; Shepherd, H.; Sarazin, M. Profiling the surface layer of optical turbulence with SLODAR. Mon. Not. R. Astron. Soc. 2010, 406, 1405–1408. [Google Scholar] [CrossRef]
  18. Wang, Z.Y.; Zhang, L.Q.; Kong, L.; Bao, H.; Guo, Y.M.; Rao, X.J.; Zhong, L.B.; Zhu, L.; Rao, C.H. A modified S-DIMM plus: Applying additional height grids for characterizing daytime seeing profiles. Mon. Not. R. Astron. Soc. 2018, 478, 1459–1467. [Google Scholar] [CrossRef]
  19. Kovadlo, P.G.; Shikhovtsev, A.Y.; Kopylov, E.A.; Kiselev, A.V.; Russkikh, I.V. Study of the Optical Atmospheric Distortions using Wavefront Sensor Data. Russ. Phys. J. 2021, 63, 1952–1958. [Google Scholar] [CrossRef]
  20. Deng, J.; Song, T.; Liu, Y. Development Overview of Daytime Atmospheric Optical Turbulence Profile Detection Technology. Prog. Astron. 2022, 40, 345–363. [Google Scholar]
  21. Basu, S. A simple approach for estimating the refractive index structure parameter ( C n 2 ) profile in the atmosphere. Opt. Lett. 2015, 40, 4130–4133. [Google Scholar] [CrossRef]
  22. Hufnagel, R.E.; Stanley, N.R. Modulation Transfer Function Associated with Image Transmission through Turbulent Media. J. Opt. Soc. Am. 1964, 54, 52–61. [Google Scholar] [CrossRef]
  23. Barletti, R.; Ceppatelli, G.; Paternò, L.; Righini, A.; Speroni, N. Mean vertical profile of atmospheric turbulence relevant for astronomical seeing. J. Opt. Soc. Am. 1976, 66, 1380–1383. [Google Scholar] [CrossRef]
  24. Ruggiero, H.; DeBenedictis, D.A. Forecasting optical turbulence from mesoscale numerical weather prediction models. In Proceedings of the DoD High Performance Modernization Program Users Group Conference, Vicksburg, MS, USA, June 2002. [Google Scholar]
  25. Dewan, E.M.; Grossbard, N. The inertial range “outer scale” and optical turbulence. Environ. Fluid Mech. 2007, 7, 383–396. [Google Scholar] [CrossRef]
  26. Thorpe, S.A. The Turbulent Ocean; Cambridge University Press: Cambridge, UK, 2005; Volume 9780521835435. [Google Scholar] [CrossRef]
  27. Ellison, T.H. Turbulent transport of heat and momentum from an infinite rough plane. J. Fluid Mech. 1957, 2, 456–466. [Google Scholar] [CrossRef]
  28. Wu, S.; Su, C.D.; Wu, X.Q.; Luo, T.; Li, X.B. A Simple Method to Estimate the Refractive Index Structure Parameter ( C n 2 ) in the Atmosphere. Publ. Astron. Soc. Pac. 2020, 132, 11. [Google Scholar] [CrossRef]
  29. Wu, S.; Wu, X.Q.; Su, C.D.; Yang, Q.K.; Xu, J.Y.; Luo, T.; Huang, C.; Qing, C. Reliable model to estimate the profile of the refractive index structure parameter ( C n 2 ) and integrated astroclimatic parameters in the atmosphere. Opt. Express 2021, 29, 12454–12470. [Google Scholar] [CrossRef] [PubMed]
  30. Xu, M.M.; Shao, S.Y.; Weng, N.Q.; Liu, Q. Analysis of the Optical Turbulence Model Using Meteorological Data. Remote Sens. 2022, 14, 3085. [Google Scholar] [CrossRef]
  31. Vanzandt, T.E.; Gage, K.S.; Warnock, J.M. Statistical-model for probability of turbulence and calculation of vertical profiles of turbulence parameters. Bull. Am. Meteorol. Soc. 1978, 59, 1230. [Google Scholar]
  32. Trinquet, H.; Vernin, J. A statistical model to forecast the profile of the index structure constant C N 2 . Environ. Fluid Mech. 2007, 7, 397–407. [Google Scholar] [CrossRef]
  33. Bi, C.C.; Qing, C.; Wu, P.F.; Jin, X.M.; Liu, Q.; Qian, X.M.; Zhu, W.Y.; Weng, N.Q. Optical Turbulence Profile in Marine Environment with Artificial Neural Network Model. Remote Sens. 2022, 14, 2267. [Google Scholar] [CrossRef]
  34. Bolbasova, L.A.; Andrakhanov, A.A.; Shikhovtsev, A.Y. The application of machine learning to predictions of optical turbulence in the surface layer at Baikal Astrophysical Observatory. Mon. Not. R. Astron. Soc. 2021, 504, 6008–6017. [Google Scholar] [CrossRef]
  35. Wang, Y.; Basu, S. Using an artificial neural network approach to estimate surface-layer optical turbulence at Mauna Loa, Hawaii. Opt. Lett. 2016, 41, 2334–2337. [Google Scholar] [CrossRef]
  36. Gladstone, J.H.; Dale, T.P. Researches on the Refraction, Dispersion, and Sensitiveness of Liquids. Philos. Trans. R. Soc. Lond. 1863, 153, 317–343. [Google Scholar]
  37. Chen, X.W.; Li, X.B.; Sun, G.; Liu, Q.; Zhu, W.Y.; Weng, N.Q. Effects of intermittency and stratification on the evaluation of optical propagation. Chin. Opt. Lett. 2017, 15, 5. [Google Scholar] [CrossRef]
  38. Taylor, K.E. Summarizing multiple aspects of model performance in a single diagram. J. Geophys. Res.-Atmos. 2001, 106, 7183–7192. [Google Scholar] [CrossRef]
  39. Haslebacher, C.; Demory, M.E.; Demory, B.O.; Sarazin, M.; Vidale, P.L. Impact of climate change on site characteristics of eight major astronomical observatories using high-resolution global climate projections until 2050 Projected increase in temperature and humidity leads to poorer astronomical observing conditions. Astron. Astrophys. 2022, 665, 53. [Google Scholar] [CrossRef]
  40. Yang, Q.K.; Wu, X.Q.; Wang, Z.Y.; Hu, X.D.; Guo, Y.M.; Qing, C. Simulating the night-time astronomical seeing at Dome A using Polar WRF. Mon. Not. R. Astron. Soc. 2022, 515, 1788–1794. [Google Scholar] [CrossRef]
Figure 1. Field observation areas and instrument. (a) Sites locations: the eastern coastal area of China (ECACN) and the northern coastal area of China (NCACN). (b) Instrument: platinum wire probes (red rectangular boxes).
Figure 1. Field observation areas and instrument. (a) Sites locations: the eastern coastal area of China (ECACN) and the northern coastal area of China (NCACN). (b) Instrument: platinum wire probes (red rectangular boxes).
Remotesensing 15 01584 g001
Figure 2. Process of MEP. Abbreviation meanings are as follows after this. MT: micro-thermal; HV: Hufnagel-Valley 5/7; H9: Hmnsp99; DN: Dewan; TE: Thorpe; EN: Ellison; WT: WSPT; WG: WSTG; MEP or ME: multi-model ensemble pattern method.
Figure 2. Process of MEP. Abbreviation meanings are as follows after this. MT: micro-thermal; HV: Hufnagel-Valley 5/7; H9: Hmnsp99; DN: Dewan; TE: Thorpe; EN: Ellison; WT: WSPT; WG: WSTG; MEP or ME: multi-model ensemble pattern method.
Remotesensing 15 01584 g002
Figure 3. Measured and estimated C n 2 profiles in ECACN. (a,b): single day C n 2 profiles; (c,d): corresponding Taylor diagram of a single day C n 2 statistics. The black solid curves in circles mark MT and models normalized standard deviation σ ^ ; the blue dashed lines mark the correlation coefficient R; the red dashed lines mark the root-mean-square difference E between the models and MT. Intuitively, the closer the model is to the reference (MT, the black circle solid point) in the diagram, the better the estimation is.
Figure 3. Measured and estimated C n 2 profiles in ECACN. (a,b): single day C n 2 profiles; (c,d): corresponding Taylor diagram of a single day C n 2 statistics. The black solid curves in circles mark MT and models normalized standard deviation σ ^ ; the blue dashed lines mark the correlation coefficient R; the red dashed lines mark the root-mean-square difference E between the models and MT. Intuitively, the closer the model is to the reference (MT, the black circle solid point) in the diagram, the better the estimation is.
Remotesensing 15 01584 g003
Figure 4. ECACN all launches MT vs. estimations C n 2 scatter diagrams. (a): MT vs. Hafnagel-Valley 5/7; (b) MT vs. Hmnsp99; (c): MT vs. Dewan; (d): MT vs. Thorpe; (e): MT vs. Ellison; (f): MT vs. WSPT; (g): MT vs. WSTG; (h): MT vs. MEP. The color indicates the frequency distribution of C n 2 .
Figure 4. ECACN all launches MT vs. estimations C n 2 scatter diagrams. (a): MT vs. Hafnagel-Valley 5/7; (b) MT vs. Hmnsp99; (c): MT vs. Dewan; (d): MT vs. Thorpe; (e): MT vs. Ellison; (f): MT vs. WSPT; (g): MT vs. WSTG; (h): MT vs. MEP. The color indicates the frequency distribution of C n 2 .
Remotesensing 15 01584 g004
Figure 5. Measured and estimated C n 2 profiles in NCACN. (a,b) A single day C n 2 profiles; (c,d) corresponding Taylor diagram of a single day C n 2 statistics.
Figure 5. Measured and estimated C n 2 profiles in NCACN. (a,b) A single day C n 2 profiles; (c,d) corresponding Taylor diagram of a single day C n 2 statistics.
Remotesensing 15 01584 g005
Figure 6. NCACN all launches MT vs. estimations C n 2 scatter diagram. (ah) The same as Figure 4 but for NCACN.
Figure 6. NCACN all launches MT vs. estimations C n 2 scatter diagram. (ah) The same as Figure 4 but for NCACN.
Remotesensing 15 01584 g006aRemotesensing 15 01584 g006b
Figure 7. ECACN integrated astronomical parameters (@ λ = 550 nm). (a) Fried parameter r 0 (in cm); (b) seeing ϵ (arcsec in ); (c) isoplanatic angle θ A O (arcsec in ); (d) scintillation rate σ I 2 ; (eh): corresponding scatter diagram.
Figure 7. ECACN integrated astronomical parameters (@ λ = 550 nm). (a) Fried parameter r 0 (in cm); (b) seeing ϵ (arcsec in ); (c) isoplanatic angle θ A O (arcsec in ); (d) scintillation rate σ I 2 ; (eh): corresponding scatter diagram.
Remotesensing 15 01584 g007
Figure 8. NCACN integrated astronomical parameters (@ λ = 550 nm). (a) Fried parameter r 0 ; (b) seeing ϵ ; (c) isoplanatic angle θ A O ; (d) scintillation rate σ I 2 ; (eh): corresponding scatter diagram.
Figure 8. NCACN integrated astronomical parameters (@ λ = 550 nm). (a) Fried parameter r 0 ; (b) seeing ϵ ; (c) isoplanatic angle θ A O ; (d) scintillation rate σ I 2 ; (eh): corresponding scatter diagram.
Remotesensing 15 01584 g008
Table 1. Radiosonde data collection of two areas.
Table 1. Radiosonde data collection of two areas.
AreasMorning LaunchesEvening LaunchesTotal Launches
ECACN01616
NCACN91120
Table 2. ECACN 16 C n 2 profiles statistics.
Table 2. ECACN 16 C n 2 profiles statistics.
StatisticsHVH9DNTEENWTWGME
R0.640.610.530.600.630.530.600.65
B i a s 0.600.570.08−0.300.18−0.005−0.15−0.11
M A E 0.810.730.760.560.530.540.640.51
R M S E 1.130.920.950.700.670.680.810.64
Table 3. NCACN 20 C n 2 profiles statistics.
Table 3. NCACN 20 C n 2 profiles statistics.
StatisticsHVH9DNTEENWTWGME
R0.690.720.620.740.750.640.730.76
B i a s 0.520.610.06−0.220.260.02−0.13−0.09
M A E 0.740.710.730.490.490.520.560.45
R M S E 1.010.890.900.620.640.660.690.58
Table 4. ECACN integrated astronomical parameters statistics (@ λ = 550 nm).
Table 4. ECACN integrated astronomical parameters statistics (@ λ = 550 nm).
 StatisticsHVH9DNTEENWTWGME
 R0.460.320.700.610.460.620.460.69
r 0 B i a s −4.076.50−8.874.56−1.230.0055.173.13
  M A E 5.016.608.874.894.173.275.314.01
  R M S E 6.118.139.866.484.794.287.175.32
 R0.140.340.530.530.440.650.400.51
ϵ B i a s 0.39−1.140.57−0.450.150.15−0.56−0.20
  M A E 0.421.170.570.540.390.290.600.38
  R M S E 0.581.370.700.580.520.390.690.42
 R0.45−0.420.560.490.450.290.030.38
θ A O B i a s −0.590.35−0.950.21−0.280.050.210.08
  M A E 0.660.430.970.300.410.280.310.28
  R M S E 0.740.591.030.450.470.400.470.40
 R0.14−0.080.510.720.620.400.310.50
σ I 2 B i a s 0.37−0.870.47−0.210.250.14−0.28−0.004
  M A E 0.381.010.470.330.290.270.400.28
  R M S E 0.531.180.600.370.390.380.460.34
Table 5. NCACN integrated astronomical parameters statistics (@ λ = 550 nm).
Table 5. NCACN integrated astronomical parameters statistics (@ λ = 550 nm).
 StatisticsHVH9DNTEENWTWGME
 R−0.220.51−0.0060.640.590.400.420.60
r 0 B i a s −4.685.90−9.752.46−3.90−1.424.111.74
  M A E 6.285.989.754.044.193.694.863.63
  R M S E 7.617.1710.894.875.374.446.034.27
 R−0.160.620.120.580.640.240.530.58
ϵ B i a s 0.63−1.320.84−0.0070.580.43−0.370.08
  M A E 0.721.400.840.600.600.570.760.55
  R M S E 1.091.601.200.800.950.940.880.75
 R0.010.03−0.030.380.470.320.200.45
θ A O B i a s −0.560.36−1.070.11−0.410.010.160.03
  M A E 0.670.411.070.290.430.290.290.29
  R M S E 0.850.521.130.360.510.330.380.32
 R−0.070.81−0.050.310.390.110.240.59
σ I 2 B i a s 0.69−1.240.840.260.690.480.020.34
  M A E 0.761.310.840.640.690.640.690.60
  R M S E 1.361.541.421.161.311.241.131.11
Table 6. Performance of MEP/ME against the best one (within parentheses) (the integrated astronomical parameters were calculated for the wavelength of light at λ = 550 nm. The values retain two decimal places).
Table 6. Performance of MEP/ME against the best one (within parentheses) (the integrated astronomical parameters were calculated for the wavelength of light at λ = 550 nm. The values retain two decimal places).
AreasParametersR Bias MAE RMSE
  C n 2 0.65(0.65:ME)−0.11(−0.005:WT)0.51(0.51:ME)0.64(0.64:ME)
  r 0 0.69(0.70:DN)3.13(0.005:WT)4.01(3.27:WT)5.32(4.28:WT)
ECACN ϵ 0.51(0.65:WT)−0.20(0.15:EN)0.38(0.29:WT)0.42(0.39:WT)
  θ A O 0.38(0.56:DN)0.08(0.05:WT)0.28(0.28:WT)0.40(0.40:ME)
  σ I 2 0.50(0.72:TE)−0.004(−0.004:ME)0.28(0.27:WT)0.34(0.34:ME)
  C n 2 0.76(0.76:ME)−0.09(0.02:WT)0.45(0.45:ME)0.58(0.58:ME)
  r 0 0.60(0.64:TE)1.74(−1.42:WT)3.63(3.63:ME)4.27(4.27:ME)
NCACN ϵ 0.58(0.64:EN)0.08(−0.007:TE)0.55(0.55:ME)0.75(0.75:ME)
  θ A O 0.45(0.47:EN)0.03(0.01:WT)0.29(0.29:ME)0.32(0.32:ME)
  σ I 2 0.59(0.59:ME)0.34(0.02:WG)0.60(0.60:ME)1.11(1.11:ME)
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

Zhang, H.; Zhu, L.; Sun, G.; Zhang, K.; Liu, Y.; Ma, X.; Zhang, H.; Liu, Q.; Cui, S.; Luo, T.; et al. A Multi-Model Ensemble Pattern Method to Estimate the Refractive Index Structure Parameter Profile and Integrated Astronomical Parameters in the Atmosphere. Remote Sens. 2023, 15, 1584. https://doi.org/10.3390/rs15061584

AMA Style

Zhang H, Zhu L, Sun G, Zhang K, Liu Y, Ma X, Zhang H, Liu Q, Cui S, Luo T, et al. A Multi-Model Ensemble Pattern Method to Estimate the Refractive Index Structure Parameter Profile and Integrated Astronomical Parameters in the Atmosphere. Remote Sensing. 2023; 15(6):1584. https://doi.org/10.3390/rs15061584

Chicago/Turabian Style

Zhang, Hanjiu, Liming Zhu, Gang Sun, Kun Zhang, Ying Liu, Xuebin Ma, Haojia Zhang, Qing Liu, Shengcheng Cui, Tao Luo, and et al. 2023. "A Multi-Model Ensemble Pattern Method to Estimate the Refractive Index Structure Parameter Profile and Integrated Astronomical Parameters in the Atmosphere" Remote Sensing 15, no. 6: 1584. https://doi.org/10.3390/rs15061584

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