Next Article in Journal
Different-Classification-Scheme-Based Machine Learning Model of Building Seismic Resilience Assessment in a Mountainous Region
Previous Article in Journal
Study of the Preparation Phase of Turkey’s Powerful Earthquake (6 February 2023) by a Geophysical Multi-Parametric Fuzzy Inference System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Global Spatio-Temporal Characteristics of Astronomical Seeing

1
Key Laboratory of Atmospheric Optics, Anhui Institute of Optics and Fine Mechanics, HFIPS, Chinese Academy of Sciences, Hefei 230031, China
2
Science Island Branch of Graduate School, University of Science and Technology of China, Hefei 230026, China
3
Advanced Laser Technology Laboratory of Anhui Province, Hefei 230037, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(9), 2225; https://doi.org/10.3390/rs15092225
Submission received: 3 March 2023 / Revised: 12 April 2023 / Accepted: 20 April 2023 / Published: 22 April 2023
(This article belongs to the Section Atmospheric Remote Sensing)

Abstract

:
Astronomical seeing is important for monitoring the atmospheric environment, observation scheduling and management, and selecting astronomical sites. This study first attempts to establish a near-global astronomical seeing map by employing the fifth European Centre for Medium-Range Weather Forecasts Reanalysis (ERA5) data combined with the estimated model. Then, some example sites’ results from ERA5 were compared against the astronomical seeing results from the balloon-borne microthermal measurements and the differential image motion monitor (DIMM) instrument. The global astronomical seeing variations exhibit large spatial dependence. The best seeing areas are generally discerned over the mid-latitude regions, consistent with the regions of the European Southern Observatory and Tibetan Plateau. In addition, the astronomical seeing values of the Tibetan Plateau in spring are better than in the other seasons. The site results from ERA5 show that the astronomical seeing values of some example sites are generally consistent with the measurements. Overall, the global astronomical seeing map presented in this study can provide a reference and basis to further understand the astronomy site selection and optoelectronics equipment observation path selection.

1. Introduction

The Earth’s atmospheric turbulence is a key parameter in determining the performance of ground-based telescopes and interferometers. Atmospheric turbulence is caused by random fluctuations of the atmosphere, mainly due to wind shear caused by the drag of airflow on the Earth’s surface, differences in the heating of different locations on the Earth’s surface by solar radiation, or thermal convection [1,2]. Additionally, knowledge of the atmospheric optical turbulence characteristics is equivalent to knowing the variability of astronomical seeing. Astronomical seeing (called ’seeing’ in the following) represents the angular size of stellar images blurred by atmospheric turbulence. It is defined in terms of the full width at half-maximum of a star image on the focal surface of a large aperture telescope measured in arcseconds. Seeing is an essential parameter for assessing the quality of astronomical sites at optical/infrared wavelengths and selecting astronomical observations [3,4,5]. Therefore, it is necessary to know global seeing data to evaluate astronomical seeing conditions. Weaker turbulence will correspond to a small seeing.
Atmospheric optical turbulence represents the characteristics of the refractive index structure constant C n 2 . Several techniques have been implemented to measure the atmospheric turbulence profile [ C n 2 ( h ) ] [6,7,8,9,10,11,12]. The most widely used is balloon-borne microthermometry [13], which has the characteristics of high vertical resolution. In addition, seeing is a function of C n 2 integrated over the light’s propagation path, which can be acquired from C n 2 vertical profiles [14,15]. The measurement instruments such as the differential image motion monitor (DIMM) [16,17,18,19,20,21] and the multi-aperture scintillation sensor (MASS) [22,23] are widely used to measure seeing data. Although many site-testing programs have been carried out worldwide, global long-term seeing data are unavailable and impractical by employing instruments due to observation difficulties and financial issues. Alternatively, adopting reanalysis data combined with turbulence estimate models is convenient, available, and economical. There are many typical accessible global-scale atmospheric reanalysis data, including ERA5 (the fifth European Centre for Medium-Range Weather Forecasts Reanalysis) data [24], the MERRA-2 (Modern-Era Retrospective analysis for Research and Applications, version 2) reanalysis data [25], JRA-55 (Japanese 55-yr) reanalysis data [26], and CFSv2 (Climate Forecast System Reanalysis, version2) data [27]. Compared to these global reanalysis products, the ERA5 dataset provides atmosphere parameters every hour with a horizontal resolution of 31 km and 137 vertical pressure levels, which exhibits higher temporal and spatial resolution [28,29].
Based on measurement experiments, scholars have summarized different atmospheric optical turbulence estimation models to model realistic atmospheric turbulence conditions. The simplest models of atmospheric optical turbulence are empirical models [30,31,32], which only require the altitude parameter. However, this kind of model lacks spatial and temporal characteristics. Another model is the optical turbulence parameterization model, such as the Tatarskii [1], which incorporates more meteorological parameters. According to Tatarskii’s theory, scholars have conducted vast research analysis [33,34]. In addition, considering the complexity and variability of atmospheric turbulence, more meteorological parameters need to be accounted for in the atmospheric estimation models. Subsequently, several outer-scale models have been developed to calculate the C n 2 profile based on the Tatarskii theory, including the Dewan model [35], the HMNSP99 model [36], and the Thorpe model [37]. However, the Thorpe model only contains temperature variables, while the Dewan model includes wind shear. It is worth noting that the HMNSP99 model contains temperature gradients in addition to wind shear. Meanwhile, early efforts show that the HMNSP99 model has been widely used in different conditions [14,24,38,39], such as marine environments and plateau areas. Although the results of turbulence detail changes that deviate from the measurements at some altitudes, the HMNSP99 model can be used as a practical method to convert the global meteorological reanalysis data into global three-dimensional quantitative astronomical seeing characteristics.
Considering the complexity of turbulence instrumentation measurements and the spatial-temporal measurement limitations for the large-scale experimental area, this study quantified and visualized the global astronomical seeing data based on the REA5 reanalysis data, providing a reference for ground-based astronomy to select the best candidate sites and schedule the best observation time. Subsequently, worldwide seasonal seeing variation characteristics are presented. Additionally, some example site seeing values from the ERA5 data, such as Da Qaidam, Haikou, Rongcheng, Ali, Daocheng, Muztagh-ata, and Lenghu, are validated by comparison with corresponding radiosonde measurements and DIMM data.
This paper is organized as follows: Section 2 describes the measurement of in situ balloon-borne microthermal and DIMM, the ERA5 data and the methodology in this study. In Section 3, the results based on the global astronomical seeing map are presented. In addition, some statistical site results from REA5 data are compared to the measurements. Finally, the conclusions are provided in Section 4.

2. Data and Methodology

2.1. Balloon-Borne Microthermal and DIMM Measurements

High-resolution micro-thermometer radiosonde measurements were performed in Da Qaidam, Haikou, and Rongcheng in August 2020, from March to April 2018, and November 2018, respectively. The detailed distribution of geographical locations and terrain heights are displayed in Figure 1. Da Qaidam, Haikou, and Rongcheng are marked by black spots. The Da Qaidam site (37.74 ° N, 95.35 ° E, Qinghai) is located on the northern edge of the Qaidam Basin above the Tibetan Plateau and has a typical inland highland desert climate, with an altitude of about 3200 m above sea level (ASL). The Haikou site (20.05 ° N, 110.19 ° E, Hainan) is located in the northern region of Hainan Island, with an altitude of about 15 m ASL. It is located at the northern edge of the low tropics and has a tropical monsoon climate. The Rongcheng site (36.46 ° N, 122.11 ° E, Shandong) is located in the eastern-most region of the Shandong Peninsula, surrounded by the Yellow Sea to the north, east, and south, with an altitude of about 80 m ASL. In addition, the site has a warm temperate continental monsoonal humid climate.
The micro-thermometer are balloon-borne payloads that measure the in situ temperature structure coefficient ( C T 2 ) using a pair of fine wire probes separated horizontally by 1 m. For visible and near-infrared wavelengths, the random fluctuations in the refractive index is mainly caused by the temperature effect. C T 2 is a proportional constant in the inertial sub-range form of the temperature structure function D T ( r , h ) [40], defined by
C T 2 ( r , h ) = D T ( r , h ) r 2 / 3 , l 0 r L 0 ,
where C T 2 is the temperature structure constant in K 2 m 2 / 3 . D T ( r , h ) is defined as one of the statistics describing temperature disturbance from two micro-temperature probes with distance r (in m), and is given by
D T ( r , h ) = [ T ( x ) T ( x + r ) ] 2 , l 0 r L 0 ,
where D T ( r , h ) denotes the temperature structure function, with units K 2 . The angle bracket represents the ensemble average, T ( x ) and T ( x + r ) in K denote the temperature at the two points separated by r, while l 0 and L 0 are the inner and outer scales of the atmospheric turbulence, in metres.
Considering the Gladstone formula [41], the relationship between the index of refraction structure constant ( C n 2 ) and the temperature structure constant ( C T 2 ) are expressed as
C n 2 ( h ) = 79 × 10 6 P ( h ) T 2 ( h ) 2 C T 2 ( r , h ) ,
where C n 2 ( h ) is the index of the refraction structure constant, with units m 2 / 3 , T is the absolute atmospheric temperature in K, and P is the air pressure in hPa.
In this study, the micro-thermometer is attached to the meteorological micro-thermometer measurement system and is carried into the atmosphere by a balloon to carry out meteorological radiosonde measurements at Da Qaidam, Haikou, and Rongcheng. Figure 2 displays the balloon-borne micro-thermometer measurement system. The device comprising two platinum wire sensors of 10 μ m in diameter employed for micro-thermometer radiosonde measurements is shown in the left panel in Figure 2. Two resistance wire probes are employed to measure the temperature difference across a 1 m horizontal distance as the instrument ascends through the atmosphere at around 5 m s 1 . In addition, two probe wires have a linear resistance-temperature coefficient and can be seen as two legs of a Wheatstone bridge. The bridge output voltage is amplitude-modulated based on the temperature difference between the probes. Subsequently, this voltage is amplified, bandpass filtered and then averaged. The frequency response range of the micro-thermometer is 0.05–30 Hz, and the noise level of the sensors and electronic processing of the signals corresponds to a temperature difference of 0.002 K [14,15,38,42].
In the sounding experiment, the altitude range of the micro-thermometer measurements is from the surface to about 30 km above sea level. Since the balloon-borne payload is attached to a meteorological radiosonde, air temperature, wind speed, air pressure, and relative humidity information are also relayed to the ground station. Then, the signal is transmitted to the ground station along with the standard meteorological data. In addition, all of the flights were released during the early morning or late evening at local time. During the sounding measurements, some data were abnormal with insufficient height because of the strong winds. After eliminating the abnormal data, 29, 30, and 14 profiles from measurements were reserved at the Da Qaidam, Haikou, and Rongcheng field campaign sites, respectively.
The DIMM is now a widely accepted instrument and has become a standard assessment for integrated atmospheric optical turbulence and a good integrated seeing measurement. DIMM measures the integrated optical turbulence through the entire atmosphere, providing an estimate of the total atmospheric seeing [16]. Note that the DIMM instruments were installed on a roof or tower several metres above the ground to avoid near-ground turbulence at the different testing sites.
The DIMM instrument at the Ali observatory is displayed in Figure 3. The DIMM at the Ali observatory is a small transportable instrument that uses a differential technique to accurately measure astroclimatic seeing conditions. Seeing is measured via the differential image motion in two sub-aperture images of a star. In particular, this measurement technique has the advantage of being insensitive to telescope and ground vibrations [43,44,45]. DIMM measures the integrated seeing from the telescope pupil to the top of the atmosphere. In addition, DIMM instruments have been widely used for many other different site-selection campaigns, such as Daocheng, Muztagh-ata and Lenghu [5,44,45,46]. The locations and topographic distribution of Ali, Daocheng, Muztagh-ata and Lenghu are shown in Figure 1. The DIMM measurements were carried out from March 2017 to February 2019 at Ali, Daocheng, and Muztagh-ata. In addition, the site testing campaign of Lenghu was conducted between October 2018 and 2020.

2.2. Methodology

Atmospheric optical turbulence is detrimental to ground-based astronomy. The estimated models of optical turbulence have been widely used to calculate vertical profiles of C n 2 , the refractive index structure constant from standard radiosonde meteorological parameters [35]. Based on the locally uniform isotropic theory, C n 2 can be estimated using the Tatarskii equation as below [1]
C n 2 = 2.8 79 × 10 6 P T 2 d T d h + γ 2 L 0 4 / 3 ,
where T is the absolute atmospheric temperature in K, P indicates the pressure in hPa, γ denotes the dry adiabatic lapse rate ( 9.8 × 10 3 K/m), and h denotes the height above ground. L 0 is the outer scale of atmospheric turbulence.
Ruggiero [36] summarized the HMNSP99 outer-scale model based on the sounding data. The HMNSP99 model is the function of wind shear (S) and the temperature gradient (dT/ d h ). Its specific expression is
L 0 4 / 3 = 0.1 4 / 3 × 10 0.362 + 16.728 S 192.347 d T d h , Troposphere 0.1 4 / 3 × 10 0.757 + 13.819 S 57.784 d T d h , Stratosphere
where S represents the vertical shear of horizontal velocity, defined as
S = u h 2 + v h 2 ,
where u and v are the north–south and east–west wind components, respectively.
Adaptive optics (AO) and diffraction-limited infrared instruments are planned as future upgrades to the telescope. To maximize and optimize the performance of the telescope and its instrument suite, it is crucial to find more candidate sites that are excellent both in terms of their astronomical performance and operation as well as maintenance. Astroclimatic seeing plays an important role in the design and optimization of AO systems. Seeing stands for the angular size of stellar images influenced by atmospheric turbulence and is usually computed by integrating the optical turbulence profiles with respect to the zenith [3]. Seeing is referred to as the zenith direction, and its specific parameter is given as follows:
ε = 5.25 λ 1 / 5 0 C n 2 ( h ) d h 3 / 5 .
where ε denotes astronomical seeing in arcseconds. Small values better represent astronomical seeing. λ is the optical wavelength, and λ = 550 nm for this study.

2.3. ERA5 Data

ERA5 is the fifth generation ECMWF reanalysis data for the global climate and weather, replacing the ERA-Interim reanalysis data publicly available from 1950, with climate data store entries for 1950–1978 (preliminary back extension) and from 1959 onwards. ERA5 adopts Cycle 41r2 of the integrated forecasting system (IFS) and is based on the 4D-Var data assimilation method. ERA5 presents global data with flexible regional options for extended periods, assimilated by combining model data with observations from across the world into a globally complete and consistent dataset using the laws of physics. ERA5 data contain a variety of atmospheric parameters with a horizontal resolution of 31 km, 137 vertical levels from the surface to 0.01 hPa, and an hourly output, encompassing absolute temperature, relative humidity, and horizontal wind speed components (u and v). The specific information on the ERA5 data is listed in Table 1.
In addition, ERA5 data in the free atmosphere have quite good reliability and resolution, both horizontally and vertically (137 levels) [28], and capture more detailed atmospheric parameters than previous lower-resolution global reanalysis data. In this study, ERA5 data from some example sites have a resolution of 0.25 ° latitude and 0.25 ° longitude. The specific information of ERA5 data at different example sites is tabulated in Table 2.

3. Results and Discussion

3.1. Global Distribution Characteristics of Seeing

Figure 4 depicts the spatio-temporal distribution characteristics of astronomical seeing worldwide with zonal mean seeing. The first row corresponds to the seeing results in DJF and MAM; while the second row shows the JJA and SON seeing results. Note that the blue colour indicates the good-seeing areas in each sub-graph. It is observed from Figure 4 that the worldwide seeing map exhibits large spatial dependence. The western part of China, such as the Tibetan Plateau (latitude of ∼30 ° N), shows significantly smaller seeing values. Additionally, there is a blue zonal band surrounded by the Andes (at a longitude of ∼70 ° W) in South America. The regions of low seeing values are broadly located in the mid-latitude regions within Asia, North America, and South America, such as the Tibetan Plateau and Cordillera Mountains, including the Andes in all seasons. These are regions where important astronomical stations are located, such as Ali, Daocheng, and Cerro Paranal, and are consistent with the locations of the important astronomical observatories from the European Southern Observatory (ESO) and Tibetan Plateau. In addition, it is evident that the seeing values over the Tibetan Plateau are smaller than those of other areas during all seasons. Particularly, the astronomical seeing values over the Tibetan Plateau are better in spring than in other seasons, seen in Figure 4. Furthermore, the seeing in Africa, North America, and South America notably depicts good-seeing areas, especially in the Cordillera Mountains. In comparison, seeing over the Cordillera Mountains appears to slightly change during the seasons.
Here the spatio-temporal distribution characteristics of astronomical seeing can be quantified by computing with zonal mean seeing in each sub-graph (right panel) from Figure 4. It is clear that most of the zonal mean seeing values are consistent with spatial distribution of the astronomical seeing, particularly in the mid-latitude regions. The zonal mean seeing results also show that the best-seeing areas are generally consistent with the regional scope of international observatory stations. Moreover, the Antarctic has small seeing values, which exhibited a similar distribution during all seasons.

3.2. Verification of the Example Sites

As we all know, it is difficult to validate global-scale seeing values due to the lack of seeing values from instrumental measurements. In this section, statistical analysis of the seeing values between the measurements and ERA5 data was conducted at some example sites to verify the model’s reliability. Note that the actual seeing values which were used for comparison were computed from in situ balloon-borne micro-thermometer measurements and reported from previous papers. The measured seeing values at Ali, Daocheng, Muztagh-ata, and Lenghu were acquired from the DIMM instrument [5,44,45].
Table 3 lists the statistical median and the first and third quartile values of the astronomical seeing values from the measurements and ERA5 data at several sites. The median seeing values from sounding measurements are 1.06, 1.15, and 1.09 arcseconds, and the corresponding values from ERA5 are 0.88, 1.17, and 1.23 arcseconds, at Da Qaidam, Haikou, and Rongcheng, respectively. The statistical median seeing values from ERA5 generally revealed the local seeing characteristics at Da Qaidam, Haikou, and Rongcheng. However, the scope of the first and third quartile values from the sounding measurements appear to be wider compared with the ERA5 data. This may result from the wider variety of climatic conditions during balloon measurements and the differences in maximum detection height in the calculation process. A clearer picture will emerge from the acquisition of a bigger dataset.
In addition, from Table 3, the median seeing from DIMM measurements are 1.08, 1.01, 0.82, and 0.75 arcseconds at Ali, Daocheng, Muztagh-ata, and Lenghu, respectively. Next, the distribution characteristics of the seeing results from ERA5 are introduced in detail at Ali, Daocheng, Muztagh-ata, and Lenghu. Firstly, the statistical seeing results from the ERA5 data at Ali, Daocheng, and Muztagh-ata are shown in Figure 5. The left panel of Figure 5 shows the histogram of all seeing data for each site. The statistical seeing values follow a log-normal distribution. The median seeing value from ERA5 at Ali is 0.87 arcsecond, and about 75% of the data are below 1 arcsecond.
Additionally, the median seeing values from ERA5 at Daocheng and Muztagh-ata are 0.96 and 0.85 arcseconds, respectively. The histogram distribution of seeing from ERA5 generally reveals the measured distributional features at the three sites, although the median values deviate slightly from the DIMM measurements [44]. Moreover, Figure 5 (right panel) illustrates the monthly statistical boxplot of the seeing values from ERA5 at the three sites. In contrast, the seeing values from the boxplots from ERA5 and DIMM measurements [45] are consistent with the general changing trend at the three sites. Note that the median seeing values from ERA5 do not exhibit the good characteristics of the measured seeing values in May 2017 at Ali from the boxplot. As reported in [45], the DIMM was previously installed on the 3 m rooftop, and after May 2017, DIMM measurements were carried out on a 5 m tower. Nevertheless, the cross-comparison results from the DIMM measurements indicate that the seeing values at 3 m and 5 m are not noticeable different. As for the monthly statistical results of seeing at Daocheng, the ERA5 results generally showed the seeing characteristics in August 2017. In addition, the seeing results from the ERA5 data at the Muztagh-ata site exhibited poor seeing characteristics in July and August, consistent with the actual DIMM measurement results at the Muztagh-ata site. Similarly, the scope of the first and third quartile values of the seeing values from the DIMM measurements in Table 3 seem to be slightly wider than the ERA5 data results. The reason for this may be that the DIMM data obtained in the process is affected by the wider variety of climatic conditions. Furthermore, the DIMM instruments are installed on top of a tower or observation building several metres from the ground. In addition, a few months of seeing data were missing due to reinforcing the tower.
Moreover, the statistical histogram of the seeing data from ERA5 at Lenghu is displayed in Figure 6. The seeing results from ERA5 have a similar distribution compared with the DIMM measurements [5]. From the figure, one can see that the median seeing value from ERA5 is 0.96 arcseconds. Although the median value obtained by ERA5 is higher than the median measured seeing value of 0.75 arcseconds, about 75% of the statistical seeing values from ERA5 are lower than 1.07 arcseconds, similar to the statistically measured results of the DIMM measurements. The seeing distribution from ERA5 in Figure 5 and Figure 6 visualizes the general distribution of the astronomical seeing values at Ali, Daocheng, Muztagh-ata, and Lenghu compared to the DIMM measurements, showing a certain ability to reconstruct the distribution of the seeing values at these sites, thus providing a potential reference for selecting optoelectronic equipment observation path selection. It also provides a new idea for finding promising astronomy sites, and then evaluating them by balloon-borne microthermal and DIMM measurements.

4. Conclusions

In this study, the global astronomical seeing distribution map was presented from the ERA5 reanalysis data combined with the estimated model. Subsequently, the spatio-temporal distribution characteristics of astronomical seeing worldwide were analysed with the seasonal variation results. Furthermore, the verification of the seeing results from ERA5 and the corresponding measurements, including sounding measurements at Da Qaidam, Haikou, Rongcheng sites and DIMM data at Ali, Daocheng, Muztagh-ata, and Lenghu. The results are summarized as follows.
The global median seeing areas generally display a large spatial dependence, discerned over the mid-latitude regions such as the Andes in South America and the Tibetan Plateau in the western part of China. The seeing values over the Tibetan Plateau and the Andes are smaller than the other areas in all seasons. The astronomical seeing values on the Tibetan Plateau are better in the spring than in other seasons. Additionally, several areas such as the Cordillera Mountains from North America and some areas in Africa also exhibit potential candidate sites for observatory stations. The results of the zonal mean seeing values demonstrate its capabilities by capturing the best seeing areas generally located in the mid-latitude regions, such as the Tibetan Plateau and the Cordillera Mountains, including the Andes in all seasons, remarkably coherent with international observatory stations at the ESO and Tibetan Plateau.
For the statistical seeing results at some example sites, the results from the ERA5 data are generally reasonable compared to the sounding and DIMM measurements. Although the seeing values from the ERA5 data are not completely consistent with the measured values at these sites, the results achieved here provide an attempt at acquiring seeing values worldwide from the high spatio-temporal resolution ERA5 data, which provides a reference for research and the selection of observatories.

Author Contributions

Conceptualization, C.B., C.Q. and W.Z.; methodology, C.B., C.Q. and W.Z.; validation, X.Q., W.Z., T.L. and N.W.; data analysis, C.B. and C.Q.; writing—original draft preparation, C.B.; writing—review and editing, C.B. and C.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Foundation of Advanced Laser Technology Laboratory of Anhui Province (Grant No. AHL2021QN02), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDA17010104), the Foundation of Key Laboratory of Science and Technology Innovation of Chinese Academy of Sciences (Grant No. CXJJ-21S028), and the Special Project of Nanhu Laser Laboratory (Grant No. 22-NHLL-ZZKY-005).

Data Availability Statement

ERA5 data underlying the results presented in this paper are available in the Copernicus Climate Change Service Climate Data Store (CDS) https://cds.climate.copernicus.eu/cdsapp#!/home, accessed on 2 March 2023; Radiosounding data may be restricted for privacy reasons.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tatarskii, V.I. Wave Propagation in a Turbulent Medium; McGraw-Hill: New York, NY, USA, 1961. [Google Scholar]
  2. Hutt, D.L. Modeling and measurement of atmospheric optical turbulence over land. Opt. Eng. 1999, 38, 1288–1295. [Google Scholar] [CrossRef]
  3. Roddier, F. V the effects of atmospheric turbulence in optical astronomy. In Progress in Optics; Elsevier: Amsterdam, The Netherlands, 1981; Volume 19, pp. 281–376. [Google Scholar]
  4. Yang, Q.K.; Wu, X.Q.; Han, Y.J.; Qing, C.; Wu, S.; Su, C.D.; Wu, P.F.; Luo, T.; Zhang, S.T. Estimating the astronomical seeing above Dome A using Polar WRF based on the Tatarskii equation. Opt. Express 2021, 29, 44000–44011. [Google Scholar] [CrossRef]
  5. Deng, L.; Yang, F.; Chen, X.; He, F.; Liu, Q.; Zhang, B.; Zhang, C.; Wang, K.; Liu, N.; Ren, A. Lenghu on the Tibetan Plateau as an astronomical observing site. Nature 2021, 596, 353–356. [Google Scholar] [CrossRef]
  6. Vernin, J.; Roddier, F. Experimental determination of two-dimensional spatiotemporal power spectra of stellar light scintillation Evidence for a multilayer structure of the air turbulence in the upper troposphere. J. Opt. Soc. Am. 1973, 63, 270–273. [Google Scholar] [CrossRef]
  7. Avila, R.; Vernin, J.; Masciadri, E. Whole atmospheric-turbulence profiling with generalized scidar. Appl. Opt. 1997, 36, 7898–7905. [Google Scholar] [CrossRef]
  8. Butterley, T.; Wilson, R.W.; Sarazin, M. Determination of the profile of atmospheric optical turbulence strength from SLODAR data. Mon. Not. R. Astron. Soc. 2006, 369, 835–845. [Google Scholar] [CrossRef]
  9. Kornilov, V.; Tokovinin, A.; Shatsky, N.; Voziakova, O.; Potanin, S.; Safonov, B. Combined MASS–DIMM instruments for atmospheric turbulence studies. Mon. Not. R. Astron. Soc. 2007, 382, 1268–1278. [Google Scholar] [CrossRef]
  10. Gimmestad, G.G.; Roberts, D.W.; Stewart, J.M.; Wood, J.W.; Eaton, F.D. Testing of LIDAR system for turbulence profiles. In Proceedings of the SPIE Defense and Security Symposium, Orlando, FL, USA, 18 April 2008; SPIE: Orlando, FL, USA, 2008; p. 695109. [Google Scholar]
  11. Carlisle, E.; Schmidt, D.; Marino, J.; Guesalaga, A. Use of SLODAR for daytime turbulence profiling. In Proceedings of the Adaptive Optics for Extremely Large Telescopes, Tenerife, Spain, 25–30 June 2017. [Google Scholar]
  12. Odintsov, S.L.; Gladkikh, V.A.; Kamardin, A.P.; Nevzorova, I.V. Determination of the Structural Characteristic of the Refractive Index of Optical Waves in the Atmospheric Boundary Layer with Remote Acoustic Sounding Facilities. Atmosphere 2019, 10, 711. [Google Scholar] [CrossRef]
  13. Azouit, M.; Vernin, J. Optical Turbulence Profiling with Balloons Relevant to Astronomy and Atmospheric Physics. Publ. Astron. Soc. Pac. 2005, 117, 536–543. [Google Scholar] [CrossRef]
  14. Qing, C.; Wu, X.Q.; Li, X.B.; Luo, T.; Su, C.D.; Zhu, W.Y. Mesoscale optical turbulence simulations above Tibetan Plateau: First attempt. Opt. Express 2020, 28, 4571–4586. [Google Scholar] [CrossRef]
  15. Han, Y.J.; Wu, X.Q.; Luo, T.; Qing, C.; Yang, Q.K.; Jin, X.M.; Liu, N.N.; Wu, S.; Su, C.D. New ( C n 2 ) statistical model based on first radiosonde turbulence observation over Lhasa. J. Opt. Soc. Am. A 2020, 37, 995–1001. [Google Scholar] [CrossRef] [PubMed]
  16. Sarazin, M.; Roddier, F. The ESO differential image motion monitor. Astron. Astrophys. 1990, 227, 294–300. [Google Scholar]
  17. Vernin, J.; Munoz-Tunon, C. Measuring astronomical seeing: The DA/IAC DIMM. Publ. Astron. Soc. Pac. 1995, 107, 265. [Google Scholar] [CrossRef]
  18. Bally, J.; Theil, D.; Billawala, Y.; Potter, D.; Loewenstein, R.F.; Mrozek, F.; Lloyd, J.P. A Hartmann differential image motion monitor (H-DIMM) for atmospheric turbulence characterisation. Publ. Astron. Soc. Aust. 1996, 13, 22–27. [Google Scholar] [CrossRef]
  19. Tokovinin, A. From differential image motion to seeing. Publ. Astron. Soc. Pac. 2002, 114, 1156. [Google Scholar] [CrossRef]
  20. McHugh, J.P.; Jumper, G.Y.; Chun, M. Balloon Thermosonde Measurements over Mauna Kea and Comparison with Seeing Measurements. Publ. Astron. Soc. Pac. 2008, 120, 1318–1324. [Google Scholar] [CrossRef]
  21. Wang, Z.; Zhang, L.; Kong, L.; Bao, H.; Guo, Y.; Rao, X.; Zhong, L.; Zhu, L.; Rao, C. A modified S-DIMM+: Applying additional height grids for characterizing daytime seeing profiles. Mon. Not. R. Astron. Soc. 2018, 478, 1459–1467. [Google Scholar] [CrossRef]
  22. Tokovinin, A.; Kornilov, V. Accurate seeing measurements with MASS and DIMM. Mon. Not. R. Astron. Soc. 2007, 381, 1179–1189. [Google Scholar] [CrossRef]
  23. Lyman, R.; Cherubini, T.; Businger, S. Forecasting seeing for the Maunakea Observatories. Mon. Not. R. Astron. Soc. 2020, 496, 4734–4748. [Google Scholar] [CrossRef]
  24. Han, Y.J.; Yang, Q.K.; Liu, N.N.; Zhang, K.; Qing, C.; Li, X.B.; Wu, X.Q.; Luo, T. Analysis of wind-speed profiles and optical turbulence above Gaomeigu and the Tibetan Plateau using ERA5 data. Mon. Not. R. Astron. Soc. 2021, 501, 4692–4702. [Google Scholar] [CrossRef]
  25. Kuo, C.L. Assessments of Ali, Dome A, and Summit Camp for mm-wave observations using MERRA-2 reanalysis. Astrophys. J. 2017, 848, 64. [Google Scholar] [CrossRef]
  26. Kobayashi, S.; Ota, Y.; Harada, Y.; Ebita, A.; Moriya, M.; Onoda, H.; Onogi, K.; Kamahori, H.; Kobayashi, C.; Endo, H. The JRA-55 reanalysis: General specifications and basic characteristics. J. Meteorol. Soc. Japan. Ser. II 2015, 93, 5–48. [Google Scholar] [CrossRef]
  27. Saha, S.; Moorthi, S.; Wu, X.; Wang, J.; Nadiga, S.; Tripp, P.; Behringer, D.; Hou, Y.T.; Chuang, H.Y.; Iredell, M. The NCEP climate forecast system version 2. J. Clim. 2014, 27, 2185–2208. [Google Scholar] [CrossRef]
  28. Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.; Horányi, A.; Muñoz-Sabater, J.; Nicolas, J.; Peubey, C.; Radu, R.; Schepers, D.; et al. The ERA5 global reanalysis. Q. J. R. Meteorol. Soc. 2020, 146, 1999–2049. [Google Scholar] [CrossRef]
  29. Guo, J.P.; Zhang, J.; Yang, K.; Liao, H.; Zhang, S.D.; Huang, K.M.; Lv, Y.M.; Shao, J.; Yu, T.; Tong, B. Investigation of near-global daytime boundary layer height using high-resolution radiosondes: First results and comparison with ERA5, MERRA-2, JRA-55, and NCEP-2 reanalyses. Atmos. Chem. Phys. 2021, 21, 17079–17097. [Google Scholar] [CrossRef]
  30. Miller, M.G.; Zieske, P.L. Turbulence Environment Characterization; Interim Report; Avco-Everett Research Lab.: Everett, MA, USA, 1979; p. 135. [Google Scholar]
  31. Good, R.E.; Beland, R.R.; Murphy, E.A.; Brown, J.H.; Dewan, E.M. Atmospheric models of optical turbulence. In Proceedings of the Modeling of the Atmosphere, Orlando, FL, USA, 24 August 1988; SPIE: Orlando, FL, USA, 1988; pp. 165–186. [Google Scholar]
  32. Jumper, G.; Beland, R. Progress in the understanding and modeling of atmospheric optical turbulence. In Proceedings of the 31st Plasmadynamics and Lasers Conference, Denver, CO, USA, 19 June 2000; p. 2355. [Google Scholar]
  33. Abahamid, A.; Jabiri, A.; Vernin, J.; Benkhaldoun, Z.; Azouit, M.; Agabi, A. Optical turbulence modeling in the boundary layer and free atmosphere using instrumented meteorological balloons. Astron. Astrophys. 2004, 416, 1193–1200. [Google Scholar] [CrossRef]
  34. Nath, D.; Venkat Ratnam, M.; Patra, A.; Krishna Murthy, B.; Bhaskar Rao, S.V. Turbulence characteristics over tropical station Gadanki (13.5°N, 79.2°E) estimated using high-resolution GPS radiosonde data. J. Geophys. Res. 2010, 115, D07102. [Google Scholar] [CrossRef]
  35. Dewan, E.M.; Good, R.E.; Beland, B.; Brown, J. A Model for C n 2 (Optical Turbulence) Profiles Using Radiosonde Data; Phillips Laboratory Technical Report, PL-TR-93-2043; ADA 279399; Phillips Laboratory: Springfield, VA, USA, 1993. [Google Scholar]
  36. Ruggiero, F.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, Austin, TX, USA, 10–14 June 2002; pp. 10–14. [Google Scholar]
  37. 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]
  38. Bi, C.C.; Qian, X.M.; Liu, Q.; Zhu, W.Y.; Li, X.B.; Luo, T.; Wu, X.Q.; Qing, C. Estimating and measurement of atmospheric optical turbulence according to balloon-borne radiosonde for three sites in China. J. Opt. Soc. Am. A 2020, 37, 1785–1794. [Google Scholar] [CrossRef]
  39. 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]
  40. Bufton, J.L. Correlation of microthermal turbulence data with meteorological soundings in the troposphere. J. Atmos. Sci. 1973, 30, 83–87. [Google Scholar] [CrossRef]
  41. Beland, R.R. Propagation through atmospheric optical turbulence. Atmos. Propag. Radiat. 1993, 2, 157–232. [Google Scholar]
  42. Qing, C.; Luo, T.; Bi, C.C.; Li, X.B.; Cui, S.C.; Yang, Q.K.; Su, C.D.; Wu, S.; Qian, X.M.; Wu, X.Q. Optical turbulence and wind speed distributions above the Tibetan Plateau from balloon-borne microthermal measurements. Mon. Not. R. Astron. Soc. 2021, 508, 4096–4105. [Google Scholar] [CrossRef]
  43. Liu, L.; Yao, Y.; Vernin, J.; Wang, H.; Yin, J.; Qian, X. Multi-instrument characterization of optical turbulence at the Ali observatory. J. Phys. Conf. Ser. 2015, 595, 012019. [Google Scholar] [CrossRef]
  44. Cao, Z.H.; Hao, J.X.; Feng, L.; Jones, H.R.; Li, J.; Xu, J.; Liu, L.Y.; Song, T.F.; Wang, J.F.; Chen, H.L. Data processing and data products from 2017 to 2019 campaign of astronomical site testing at Ali, Daocheng and Muztagh-ata. Res. Astron. Astrophys. 2020, 20, 082. [Google Scholar] [CrossRef]
  45. Feng, L.; Hao, J.X.; Cao, Z.H.; Bai, J.M.; Yang, J.; Zhou, X.; Yao, Y.Q.; Hou, J.L.; Zhao, Y.H.; Liu, Y. Site testing campaign for the Large Optical/infrared Telescope of China: Overview. Res. Astron. Astrophys. 2020, 20, 080. [Google Scholar] [CrossRef]
  46. Song, T.F.; Liu, Y.; Wang, J.X.; Zhang, X.F.; Liu, S.Q.; Zhao, M.Y.; Li, X.B.; Cai, Z.C.; Song, Q.W.; Cao, Z.H.; et al. Site testing campaign for the Large Optical/infrared Telescope of China: General introduction of the Daocheng site. Res. Astron. Astrophys. 2020, 20, 085. [Google Scholar] [CrossRef]
Figure 1. The black spots in the topographic distribution map mark the locations of Da Qaidam (95.35 ° E, 37.74 ° N), Haikou (110.19 ° E, 20.05 ° N), Rongcheng (122.11 ° E, 36.46 ° N), Ali (80.06 ° E, 32.31 ° N), Daocheng (100.11 ° E, 29.11 ° N), Muztagh-ata (74.90 ° E, 38.33 ° N) and Lenghu (93.89 ° E, 38.61 ° N), whereas the green to violet colour gradient indicates the terrain heights from low to high.
Figure 1. The black spots in the topographic distribution map mark the locations of Da Qaidam (95.35 ° E, 37.74 ° N), Haikou (110.19 ° E, 20.05 ° N), Rongcheng (122.11 ° E, 36.46 ° N), Ali (80.06 ° E, 32.31 ° N), Daocheng (100.11 ° E, 29.11 ° N), Muztagh-ata (74.90 ° E, 38.33 ° N) and Lenghu (93.89 ° E, 38.61 ° N), whereas the green to violet colour gradient indicates the terrain heights from low to high.
Remotesensing 15 02225 g001
Figure 2. (Left) micro-thermometer of the balloon-borne micro-thermometer measurement system, where the red circles represent the platinum wire probes. (Right) The balloon-borne micro-thermometer system in flight.
Figure 2. (Left) micro-thermometer of the balloon-borne micro-thermometer measurement system, where the red circles represent the platinum wire probes. (Right) The balloon-borne micro-thermometer system in flight.
Remotesensing 15 02225 g002
Figure 3. The DIMM instrument at the Ali observatory.
Figure 3. The DIMM instrument at the Ali observatory.
Remotesensing 15 02225 g003
Figure 4. The spatio-temporal characteristics of the astronomical seeing worldwide. Note that the zonal mean seeing values are plotted in the right panel at each sub-graph. The first and second rows correspond to the seeing results in December–February (DJF), March–May (MAM), June–August (JJA), and September–November (SON), respectively.
Figure 4. The spatio-temporal characteristics of the astronomical seeing worldwide. Note that the zonal mean seeing values are plotted in the right panel at each sub-graph. The first and second rows correspond to the seeing results in December–February (DJF), March–May (MAM), June–August (JJA), and September–November (SON), respectively.
Remotesensing 15 02225 g004
Figure 5. Statistical analysis of the seeing values at Ali, Daocheng, and Muztagh-ata during March 2017 and February 2019 from the ERA5 data. (Left panel) Histogram of all seeing data from ERA5 for each site. (Right panel) Monthly statistical boxplots of the seeing values from ERA5 at three sites. The boxplot presents the monthly statistical seeing results at each site, whereas the red boxes refer to the data from March 2017 to February 2018 UTC, and the blue boxes signify data from March 2018 to February 2019 UTC. The first, second, and third rows correspond to the seeing statistical results from ERA5 at Ali, Daocheng, and Muztagh-ata, respectively.
Figure 5. Statistical analysis of the seeing values at Ali, Daocheng, and Muztagh-ata during March 2017 and February 2019 from the ERA5 data. (Left panel) Histogram of all seeing data from ERA5 for each site. (Right panel) Monthly statistical boxplots of the seeing values from ERA5 at three sites. The boxplot presents the monthly statistical seeing results at each site, whereas the red boxes refer to the data from March 2017 to February 2018 UTC, and the blue boxes signify data from March 2018 to February 2019 UTC. The first, second, and third rows correspond to the seeing statistical results from ERA5 at Ali, Daocheng, and Muztagh-ata, respectively.
Remotesensing 15 02225 g005
Figure 6. Statistical histogram of the seeing data at the Lenghu site from ERA5.
Figure 6. Statistical histogram of the seeing data at the Lenghu site from ERA5.
Remotesensing 15 02225 g006
Table 1. The specific information of ERA5 data.
Table 1. The specific information of ERA5 data.
Horizontal CoverageHorizontal ResolutionVertical ResolutionVertical CoverageTemporal CoverageTemporal Resolution
Global0.25 ° × 0.25 ° 137 level1000 hPa to 1 hPa1959 to presentHourly
Table 2. The specific information of ERA5 data at different example sites.
Table 2. The specific information of ERA5 data at different example sites.
NameSite (lon, lat)ERA5 (lon, lat)Time (UTC)
Da Qaidam95.35 ° E, 37.74 ° N95.25 ° E, 37.75 ° NAug. 2020
Haikou110.19 ° E, 20.05 ° N110.25 ° E, 20.00 ° NApr. 2018
Rongcheng122.11 ° E, 36.46 ° N122.00 ° E, 36.50 ° NNov. 2018
Ali80.06 ° E, 32.31 ° N80.00 ° E, 32.25 ° NMar. 2017 to Feb. 2019
Daocheng100.11 ° E, 29.11 ° N100.00 ° E, 29.00 ° NMar. 2017 to Feb. 2019
Muztagh-ata74.90 ° E, 38.33 ° N75.00 ° E, 38.25 ° NMar. 2017 to Feb. 2019
Lenghu93.89 ° E, 38.61 ° N94.00 ° E, 38.50 ° NOct. 2018 to 2020
Table 3. Comparison of the astronomical seeing values between the measurements and the ERA5 data at some example sites. Note that 25% and 75% refer to the first and third quartiles of the seeing values from the statistical data. In addition, the measured seeing values at Da Qaidam, Haikou, and Rongcheng are calculated from the sounding measurements; while the seeing values at Ali, Daocheng, Muztagh-ata and Lenghu are measured from the DIMM instrument.
Table 3. Comparison of the astronomical seeing values between the measurements and the ERA5 data at some example sites. Note that 25% and 75% refer to the first and third quartiles of the seeing values from the statistical data. In addition, the measured seeing values at Da Qaidam, Haikou, and Rongcheng are calculated from the sounding measurements; while the seeing values at Ali, Daocheng, Muztagh-ata and Lenghu are measured from the DIMM instrument.
Site NameSeeing (Arcseconds) (Median [25%, 75%])
MeasurementERA5
Da Qaidam1.06 [0.72, 1.75]0.88 [0.85, 0.93]
Haikou1.15 [0.74, 2.32]1.17 [1.15, 1.20]
Rongcheng1.09 [0.98, 1.38]1.23 [1.17, 1.30]
Ali1.08 [0.88, 1.39] [44]0.87 [0.79, 1.00]
Daocheng1.01 [0.84, 1.22] [44]0.96 [0.85, 1.10]
Muztagh-ata0.82 [0.64, 1.06] [44]0.85 [0.78, 0.92]
Lenghu0.75 [0.61, 1.03] [5]0.96 [0.88, 1.07]
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

Bi, C.; Qing, C.; Qian, X.; Luo, T.; Zhu, W.; Weng, N. Investigation of the Global Spatio-Temporal Characteristics of Astronomical Seeing. Remote Sens. 2023, 15, 2225. https://doi.org/10.3390/rs15092225

AMA Style

Bi C, Qing C, Qian X, Luo T, Zhu W, Weng N. Investigation of the Global Spatio-Temporal Characteristics of Astronomical Seeing. Remote Sensing. 2023; 15(9):2225. https://doi.org/10.3390/rs15092225

Chicago/Turabian Style

Bi, Cuicui, Chun Qing, Xianmei Qian, Tao Luo, Wenyue Zhu, and Ningquan Weng. 2023. "Investigation of the Global Spatio-Temporal Characteristics of Astronomical Seeing" Remote Sensing 15, no. 9: 2225. https://doi.org/10.3390/rs15092225

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