Home Statistical comparison analysis of different real-time kinematic methods for the development of photogrammetric products: CORS-RTK, CORS-RTK + PPK, RTK-DRTK2, and RTK + DRTK2 + GCP
Article Open Access

Statistical comparison analysis of different real-time kinematic methods for the development of photogrammetric products: CORS-RTK, CORS-RTK + PPK, RTK-DRTK2, and RTK + DRTK2 + GCP

  • Sergio A. Monjardín-Armenta , Jesús Gabriel Rangel-Peraza , Antonio J. Sanhouse-García , Wenseslao Plata-Rocha , Sergio Arturo Rentería-Guevara and Zuriel Dathan Mora-Félix EMAIL logo
Published/Copyright: September 23, 2024
Become an author with De Gruyter Brill

Abstract

Traditional photogrammetry techniques require the use of Ground Control Points (GCPs) to accurately georeference aerial images captured by unmanned aerial vehicles (UAVs). However, the process of collecting GCPs can be time-consuming, labor-intensive, and costly. Real-time kinematic (RTK) georeferencing systems eliminate the need for GCPs without deteriorating the accuracy of photogrammetric products. In this study, a statistical comparison of four RTK georeferencing systems (continuously operating reference station (CORS)-RTK, CORS-RTK + post-processed kinematic (PPK), RTK + dynamic RTK 2 (DRTK2), and RTK + DRTK2 + GCP) is presented. The aerial photo was acquired using a Dà-Jiāng Innovation Phantom 4 RTK. The digital photogrammetric processing was performed in Agisoft Metashape Professional software. A pair of global navigation satellite systems (GNSSs) receiving antennas model CHC x900 were used for the establishment of check points (CPs). The accuracy of photogrammetric products was based on a comparison between the modeled and CP coordinates. The four methods showed acceptable planimetric accuracies, with a root mean square error (RMSE) X,Y ranging from 0.0164 to 0.0529 m, making the RTK-CORS + PPK method the most accurate (RMSE X,Y = 0.0164 m). RTK-CORS + PPK, RTK-DRTK2, and RTK-DRTK2 + GCP methods showed high altimetric accuracies, with RMSEZ values ranging from 0.0201 to 0.0334 m. In general, RTK methods showed a high planimetric and altimetric accuracy, similar to the accuracy of the photogrammetric products obtained using a large number of GCPs.

Graphical abstract

Abbreviations

ANOVA

analysis of variance

CORS

continuously operating reference station

CPs

check points

DEM

digital elevation models

DJI

Dà-Jiāng Innovations

DRTK2

dynamic real-time kinematic 2

DTMs

digital terrain models

GCPs

ground control points

GGM10

Gravimetric Geoid of Mexico 2010

GNSS

global navigation satellite system

GS

ground station

GSD

ground sample distance

INEGI

National Institute of Statistics and Geography

LAN

local area network

LIDAR

laser imaging detection and ranging

LSD

least significant difference

NTRIP

networked transport of RTCM via internet protocol

PPK

post-processed kinematic

PPP

precise point positioning

RGNA

active national geodetic network

RMSE

root mean square error

RTCM

radio technical commission for maritime services

RTK

real-time kinematic

UAV

unmanned aerial vehicle

UTM

Universal Transverse Mercator

1 Introduction

In the last decade, digital photogrammetry technology has progressed rapidly with the use of unmanned aerial vehicles (UAVs) because of the large number of sensors and devices that can be adapted to them to obtain data with higher spatial, spectral, temporal, angular, and radiometric resolution, reducing cost, time, and human resources. Photogrammetry with UAVs has been used for the development of several studies such as topographic surveys [1,2], characterization of vegetation structure [3,4], implementation of precision agriculture to improve farmland management, increasing crop yields [5,6], 3D modeling and visualization of archaeological structures [7,8], inspection of buildings and dams [9,10,11], identification of changes in urban and rural landholdings in cadastral studies [12], spatiotemporal monitoring of changes derived from natural phenomena and anthropogenic activities [13,14], detection of geomorphological changes in rivers [15], and the characterization of ice, glacial, and peri-glacial morphology [16,17].

In topography, digital elevation models (DEM) and ortho-photographs are the main results of photogrammetric processing of images obtained with UAVs, where the altimetric and planimetric accuracy represent the most relevant concerns of these products [17,18,19,20] since they are directly related to particular scientific and professional purposes. Normally, the georeferencing of photogrammetric models has been performed indirectly, establishing ground control points (GCPs) on the study surface, but this task is labor-intensive, time-consuming, and costly. However, the direct georeferencing of these models has been in constant evolution due to the technology integrated into UAVs.

Global navigation satellite systems (GNSSs) are used for UAV positioning. However, a GNSS signal blockage or a malfunction could damage the navigation system, especially if it relies exclusively on GNSS. To achieve redundancy, integrated navigation systems have been developed to ensure reliability, availability, and reliable autonomous navigation. Real-time kinematic (RTK) and post processed kinematic (PPK) technologies are significant steps in the evolution of digital aerial photogrammetry with UAVs. RTK is a crucial tool for real-time positioning in surveying and other applications that demand high levels of accuracy [21,22]. This technique uses a GNSS receiver as a reference station at a known coordinate, which then sends correction information to rover receiver(s) via an ultra-high frequency radio modem. PPK is another type of correction technology like RTK but corrects the location data after they are collected and uploaded. Both RTK and PPK technologies can achieve positioning accuracy ranging from a few centimeters to 0.1 m near the base station and are used to obtain precise location data as a substitute for conventional methods, such as GCPs [23].

Nowadays, the use of networked transport of RTCM via internet protocol (NTRIP) and continuously operating reference stations (CORS) for RTK correction is of great significance for the development of photogrammetric products. NTRIP enables the transmission of RTK correction data via the internet, supporting real-time positioning with high accuracy. Therefore, the need for a direct communication link between the base and rover receivers is eliminated, reducing cost, and simplifying field operations [24]. On the other hand, CORS is a network of reference stations that continuously collects GNSS data at known locations. These reference stations provide accurate and up-to-date correction data for RTK positioning. By using CORS data, the accuracy of the RTK measurements can be improved, where signal blockages or malfunction limitations can be overcome [25]. The combination of NTRIP and CORS enhances the effectiveness of RTK correction by enabling wider coverage, reducing reliance on local base stations, and improving the availability and accuracy of correction data [26].

The studies evaluating the accuracy of data collected by UAVs are normally conducted using GCPs or check points (CPs) [27]. Few recent studies have been developed comparing different methods using the RTK approach. Taddia et al. [21] reported the use of the PPK method to generate photogrammetric models and digital terrain models (DTMs) of coastal areas by employing GNSS data from CORS with a Dà-Jiāng Innovation (DJI) Phantom 4 RTK drone. The results show that very similar accuracies were found with and without using GCPs. Zeybek [25] evaluated the performance of planimetric and altimetric accuracy in 24 flights by direct georeferencing (without GCP) varying flight heights between 75 and 100 m, using the CORS and the Dynamic RTK 2 (DRTK2) methods. The DRTK2 method consisted of using the DRTK2 GNSS receiver manufactured by the DJI-P4 RTK company (Phantom 4 RTK UAV) to provide centimeter-level positioning accuracy in RTK. According to Zeybek [25], both methods provided acceptable positional accuracy levels, with root mean squared error (RMSE) values of 0.01–0.03 m for the horizontal axis and 0.04–0.06 cm for the vertical axis.

Štroner et al. [28] reported the accuracy of horizontal and vertical coordinate components obtained using an RTK UAV system and GCPs with a DJI Phantom 4 RTK. The results demonstrated that the errors of the RTK UAV system were similar to those obtained with GCPs in horizontal coordinates, but the errors in vertical coordinates were 2–3 times greater. Eker et al. [29] compared UAV-based RTK and PPK methods for six different surface types (roads, shadows, shrubs, boulders, trees, and ground). The lowest RMSE was observed on a solid textured surface since the surface types of ground, roads, and shrubs showed an RMSE 0.10 m lower than that of other surface types.

Other studies have focused on determining the influence of external factors on the UAV RTK technology. Cho and Lee [30] observed an accuracy reduction caused by the lack of RTK signal and proposed the use of GCP and PPK to deal with RTK signal interruption in UAV photogrammetry. Tomaštík et al. [31] evaluated the effect of vegetation seasonal variation on the accuracy of photogrammetric products (point cloud, orthoimages, and DEMs). The results evidenced that the UAV RTK/PPK technology was not influenced by vegetation seasonal variation. Kefauver et al. [32] assessed the effect of different flight patterns and altitudes on positional accuracy. They demonstrated that it is possible to enhance positioning accuracy when real-time positional access is combined with post-processing techniques. Martínez-Carricondo et al. [33] demonstrated that using multiple fixed GNSS base stations simultaneously can reduce the systematic error of elevation data for RTK photogrammetric processes, improving altimetric accuracy without the need to use GCPs or oblique photographs.

Despite several studies that have been carried out to assess the accuracy of UAV georeferencing, most of these studies only evaluate the georeferencing methods separately and no fair comparison between these methods is given [32,34,35,36,37,38]. While some studies refuse the use of real-time positioning methods without the use of GCPs [39,40], other studies acknowledge the significance, potential, and effectiveness of the PPK method for UAS surveys [22,31,41,42,43]. Also, the precise point positioning (PPP) method provides real-time positioning using corrections from multiple reference stations to compute precise satellite orbits and clock errors, achieving high-precision positioning without relying on nearby base stations [44,45,46]. However, the PPP method was not analyzed in this study.

The aim of the present work is to evaluate the altimetric and planimetric accuracy by comparing four different methods that are based on the RTK approach (CORS-RTK, CORS-RTK + PPK, RTK + DRTK2, and RTK + DRTK2 + GCP) using the same morphological conditions. 28 CPs obtained from GNSS antennas were used to assess the altimetric and planimetric accuracy for each method, but CPs 12 and 25 were eliminated because they were damaged before the photogrammetric flights were made. Once the performance of the altimetric and planimetric accuracy was obtained, a rigorous statistical analysis was carried out through ANOVA and the least square difference (LSD) multiple range test. These statistical analyses were used to demonstrate a statistically significant difference between different methods and to determine the method with the least error (greater accuracy). This statistical comparison has not been implemented in any previous work, which represents a novel scheme for the improvement of photogrammetric products using UAVs, such as DEM and ortho-photographs.

2 Methods

2.1 Study area

The study area was an urban area of the city of Culiacan, Mexico, which covers approximately 16 ha between coordinates 24.751554°, 24.746705° latitude, and −107.395484°, −107.392541° longitude (Figure 1). Part of the study area is urbanized, characterized by paved roads, buildings, and private households, but the presence of unpaved roads and vegetation is also distinguished. This area was selected because an elevation difference of 37 m is observed in the terrain. The presence of anthropogenic buildings and vegetation, and the variation in terrain elevation are external factors that could have an important effect on the accuracy of the RTK techniques used in this study.

Figure 1 
                  Study area.
Figure 1

Study area.

2.2 Equipment

A DJI Phantom 4 RTK quadcopter was used for the photogrammetric flights. This UAV was used to obtain real-time georeferenced images in RTK mode with a high positioning accuracy (cm). In addition, a pair of GNSS receiving antennas model CHC x900 were used for the establishment of the CP, also called GCPs.

2.3 Methodology

The methodology used in the present study was carried out in five stages. Stage 1 consisted of establishing a network of CPs distributed strategically throughout the study area. These CPs cover the entire area and were located in a network of irregular triangles with angles of approximately 60° as suggested by Zhang et al. [47]. These points were not obstructed by any vertical structure or vegetation shadow (Figure 2). The maximum distance between the CP points was approximately 150 m since the positioning of these points was based on topography and surface configuration. CPs were only used to verify the accuracy of the photogrammetric products obtained and did not intervene in the digital photogrammetric processing. In turn, GCPs are used to adjust the orientation of the images during photogrammetric processing. However, both (CP and GCP) are measured and established with the same methodological rigor.

Figure 2 
                  Distribution of the CP network.
Figure 2

Distribution of the CP network.

Some points of this CP network were established on concrete sidewalks or other types of concrete structures. Other points were on the natural terrain. In these cases, a flat concreted structure was constructed with dimensions of 0.40 m × 0.40 m (Figure 3a). A red mark was rendered in these structures throughout the CP network (Figure 3b and c). This rendering was carried out to facilitate the identification of CPs in the images acquired by the drone.

Figure 3 
                  Establishment and measurement of CP marks. (a) Flat concreted structure was constructed with dimensions of 0.40 m × 0.40 m; (b and c) red mark was rendered in these structures throughout the CP network; (d) Measuring CPs with the mobile antenna in RTK mode.
Figure 3

Establishment and measurement of CP marks. (a) Flat concreted structure was constructed with dimensions of 0.40 m × 0.40 m; (b and c) red mark was rendered in these structures throughout the CP network; (d) Measuring CPs with the mobile antenna in RTK mode.

In stage 2, a GNSS base point was established. This point was measured with a GNSS receiver model CHC x900 at intervals of 5 s for 2.5 h. The information acquired was processed and adjusted with the active national geodetic network (RGNA) of Mexico using the monitoring station called “CULC” located in the center of the city of Culiacan and at an approximate distance of 5.3 km from the present study area. The RGNA is operated by the National Institute of Statistics and Geography. After, the GNSS base receiver was installed and kept on a known point (GNSS base point), the mobile antenna in RTK mode was used to obtain the coordinates of each CP throughout the CP network. A bipod was set up at each point of interest to ensure the receiver was positioned correctly (centered and leveled) (Figure 3d). This procedure was carried out to ensure accurate and reliable measurements minimizing errors caused by uneven ground.

Stage 3 consisted of defining the photogrammetric flight parameters and the type of photogrammetric survey [48,49]. Two flights were performed at an altitude of 70 m. Under this condition, the ground sample distance (GSD) is comparable with the accuracies obtained with GNSS antennas in RTK mode. The DJI GS RTK App was used to plan the RTK flight. To produce accurate aerial maps, this software automatically determined the photogrammetric parameters as a function of the flight height and the longitudinal and transverse overlaps. In this study, a flight speed of 5 m/s was used, with a camera angle of −90°, a forward overlap of 80% and a side overlap of 70% according to the suggestion by Mora-Felix et al. [20].

The photogrammetric flight 1 was processed using two different approaches: RTK-CORS and RTK-CORS + PPK. The RTK-CORS approach was based on the NTRIP protocol, where a rover receiver was connected to a reference station (CORS) via the internet. The CORS station collected the receiver’s data from multiple GNSS satellites, calculated the positioning information, and generated the correction data based on the difference between the measured and known positions. These correction data were sent to the rover receiver through the NTRIP protocol. This enabled to achieve centimeter-level accuracy in positioning [25]. The approximate distance between the CORS station and the study area is 4.66 km.

In the RTK-CORS + PPK approach, the rover receiver collects real-time corrections from the nearest CORS station, similar to the RTK-CORS approach. Then, the raw GNSS data are compared with the precise data from the reference station in post-processing software called Redtoolbox [41] to calculate highly accurate positions. The post-processing consisted of adjusting the coordinates of each of the images, taking into account factors such as atmospheric conditions and satellite orbits that could affect real-time positioning [50].

The photogrammetric flight 2 was also processed using two different approaches: RTK + DRTK2 and RTK + DRTK2 + GCP. RTK + DRTK2 combines RTK positioning with advanced algorithms and technologies to enhance the accuracy and reliability of real-time positioning for UAVs. The DRTK2 technology was developed by DJI enterprise and designed specifically for UAV applications [51]. This system consists of a mobile ground station called DRTK2, which is an enhanced GNSS receiver antenna that provides real-time differential corrections to the moving UAV that can communicate via 4G, OcuSync, WiFi, and LAN, ensuring stable and continuous data transmission. Therefore, high accuracy and precision in positioning while in flight can be achieved even in difficult environments with obstacles or interference [52].

The RTK + DRTK2 + GCP approach was based on the RTK + DRTK2 method, where the RTK base station, rover receiver, and DRTK 2 mobile station work together to achieve highly accurate real-time positioning for the drone. By using this technology, the drone captures aerial imagery, and the collected data were processed using Agisoft Metashape Professional software. During the photogrammetric process, some of the CPs were used as reference points to accurately align and georeference the captured data. The alignment and georeferencing process was carried out using six CPs as GCPs. These GCPs were CP1, CP9, CP15, CP18, CP22, and CP27 (Figure 2).

In stage 4, all photogrammetric processes were performed in Agisoft Metashape Professional software using a standard workflow. First, the UAV images were imported from the files to the software. Then, the projection of Geodetic coordinates was changed to Universal Transverse Mercator Zone 13 North (UTM 13N) and the vertical coordinate reference system GGM10 (Gravimetric Geoid of Mexico 2010), because the CORS antenna is referenced to this vertical coordinate reference system. A high-precision image orientation process was carried out using the reference values (40,000 key points per image and 4,000 link points per image). The adaptive model of camera adjustment was selected since the UAV camera already has a lens correction to reduce distortion errors. This procedure was performed for all photogrammetric processing modes, except for the RTK + DRTK2 + GCP method where six CPs were used as GCPs, as mentioned before. The dense point cloud was performed for each of the photogrammetric processes, using high-quality depth filtering, which eliminated areas with low stereo vision, and calculated the disparity for each pixel using the structure from motion multi-stereo vision algorithms [53]. The DEM was created using the dense point cloud and the orthophotography was generated from the DEM for each of the processes.

Stage 5 consisted of evaluating the planimetric and altimetric accuracy of the orthophotography and DEM results from the different photogrammetric processes. This evaluation refers to the proximity of the observation of a coordinate point to a true value. In this study, the true coordinates were obtained directly with the GNSS antennas (CPs). The true coordinates were compared with the coordinates obtained from the orthophotography and DEM according to the marks observed. The root mean square error (RMSE) was used to measure the difference between the values obtained from the orthophotography and DEM with the true coordinates (GNSS points) [20,54].

Equation (1) was used to validate the planimetric error (RMSE X , Y ), which was calculated as the distance between the GCPs and the marked points (CPs) on the ground in both the x and y coordinates [20].

(1) RMSE X , Y = i = 1 n ( ( ( X GNSS X ORT ) 2 + ( Y GNSS Y ORT ) 2 ) ) n ,

where X GNSS and Y GNSS are the X and Y coordinates measured with GNSS antennas; X ort and Y ort are the coordinates obtained from the orthophotography, respectively; and n is the number of points used for the accuracy assessment.

By using equation (2), the altimetric accuracy (RMSE z ) of the DEM was assessed. RMSE z compared the elevation points acquired through GNSS with those that were marked on the ground [20].

(2) RMSE Z = i = 1 n ( ( Z GNSS Z DEM ) 2 ) n ,

where z GNSS is the Z coordinate of the GNSS antenna measurements; and Z DEM is the elevation coordinates obtained from the DEM.

The global error (RMSE x,y,z ) was measured by using the RMSE in the three-dimensional coordinates (X, Y, and Z coordinates). This statistic represents the overall accuracy between the estimated values and the true values in all three spatial dimensions (equation (3)) [55].

(3) RMSE X , Y , Z = i = 1 n ( ( ( X GNSS X ORT ) 2 + ( Y GNSS Y ORT ) 2 + ( Z GNSS Z DEM ) 2 ) ) n ,

where X GNSS , Y GNSS , and Z GNSS are the X, Y, and Z coordinates measured with GNSS antennas; and X ort , Y ort y Z DEM are the coordinates obtained from orthophotography and DEM.

An analysis of variance (ANOVA) was performed to evaluate whether any of the photogrammetric processing approaches influenced the accuracy of the photogrammetric products obtained (orthophotography and DEM). The planimetric and altimetric errors obtained were used as the response variable for this statistical analysis. A multiple-range test was used to determine significant differences between the different real-time positioning methods. This test was performed to identify the methods that are significantly different from each other. In this study, the LSD multiple range test was carried out once the ANOVA test indicated that significant differences were found between the methods. A significance level (α) of 0.05 was set for hypothesis testing in both ANOVA and LST statistical analyses since this is the most common value used for statistical decision-making.

3 Results

Since the photogrammetric products observed no visual difference between the different methods, this study only shows the orthophotographs and DEM obtained with the RTK-CORS + PPK method as an example of the photogrammetric processing (Figures 4 and 5). The pixel size was 2.2 cm × 2.2 cm (GSD size) in all orthophotographs and DEMs.

Figure 4 
               Orthophotograph obtained with the RTK-CORS + PPK method.
Figure 4

Orthophotograph obtained with the RTK-CORS + PPK method.

Figure 5 
               DEM obtained using the RTK-CORS + PPK approach.
Figure 5

DEM obtained using the RTK-CORS + PPK approach.

Based on these photogrammetric products, the CP coordinates were extracted at the sub-pixel level from each of the different photogrammetric processes. The difference between the estimated and the measured coordinates was determined. These differences (errors) were obtained at each of the CPs for the different methods.

The RTK-CORS method showed the highest absolute errors, mainly in the elevation, where the maximum difference was 0.377 m at CP17 (Figure 6), and a mean elevation error of 0.156 m was observed, being this CP, the one with the greatest topographic unevenness and therefore the greatest distance between the drone at the moment of the image capture and the terrestrial surface. The planimetry error showed a slighter error variation, where a maximum difference of 0.071 m was observed in the X axis, while a maximum error of 0.024 m was observed in the Y axis. Figure 6 confirms that the RTK-CORS method presented the greatest error variation. In particular, the greatest error difference in the Z coordinate was found from CP9 to CP27, because these points present a significant topographic unevenness with respect to the drone takeoff point. Therefore, the RTK-CORS method was the most sensitive approach since slight changes in photo scale (terrain elevation variation) produced the greatest errors. In this method, the greater the distance between the sensor and the photographed surface, the greater the error was observed.

Figure 6 
               Errors in the Z axis for each method evaluated.
Figure 6

Errors in the Z axis for each method evaluated.

The RTK-CORS + PPK method presented the lowest error difference in planimetry. This method showed maximum differences of 0.026 m on the X-axis and 0.027 m on the Y-axis. These values are lower than those observed for the other real-time positioning methods used in this study. The RTK-DRTK2 method showed maximum errors of 0.067 and 0.054 m on the X-axis and Y-axis, respectively, while the RTK-DRTK2 + GCP method presented similar errors, with maximum values of 0.068 and 0.053 m on the X-axis and Y-axis. The RTK-CORS method presented the highest error on the X-axis with a maximum error of 0.071 m. However, the Y-axis error in the RTK-CORS method was similar to the one observed in the RTK-CORS+PPK method with a maximum error of 0.024 m.

Figures 7 and 8 show that all real-time position methods presented a slight variation in planimetry, with errors ranging from −0.071 m to 0.030 m for the X-axis and −0.025 to 0.055 m for the Y-axis. These planimetric errors were an order of magnitude less than the errors observed for altimetry. According to these results, a high planimetric accuracy was achieved in this study.

Figure 7 
               Errors in the X axis for each method evaluated.
Figure 7

Errors in the X axis for each method evaluated.

Figure 8 
               Errors in the Y axis for each method evaluated.
Figure 8

Errors in the Y axis for each method evaluated.

The RMSE was calculated using equations (1)–(3). The RMSE X,Y corroborates that the RTK-CORS+PPK method obtained the lowest planimetric error (RMSE X,Y = 0.0164 m) (Table 1). In addition, it was shown that the RTK-CORS + PPK, RTK-D RTK2, and RTK-DRTK2 GCP methods presented very similar errors with a mean RMSE X,Y value close to 0.05 m. Table 1 also shows that the RTK-DRTK2 + GCP method presented the lowest RMSEz value with 0.0201 m. However, similar RMSEz values of 0.0209 and 0.033 m were obtained for the RTK-DRTK2 and RTK-CORS + PPK methods, respectively. Based on the RMSEz values obtained, a high altimetric accuracy was obtained for the photogrammetric products.

Table 1

RMSE results of the photogrammetric methods

Method RMSE x,y,z (m) RMSE xy (m) RMSE z (m)
RTK-CORS 0.3367 0.0529 0.2015
RTK-CORS + PPK 0.0372 0.0164 0.0334
RTK-DRTK2 0.0561 0.0520 0.0209
RTK-DRTK2 + GCP 0.0546 0.0507 0.0201

The overall error value obtained (RMSE X,Y,Z ) suggests that the best method to obtain photogrammetric products from orthophotos and DEMs using UAV is the RTK-CORS + PPK method. This method obtained an RMSE X,Y,Z less than 4 cm. The method with the largest RMSE X,Y,Z was RTK-CORS with 0.336 m. This high overall error was mostly influenced by the altimetric error since a low accuracy was obtained (RMSE Z = 0.201 m). However, the RTK-CORS method presented an RMSEx,y comparable with the other real-time positioning methods used in this study (RMSE X,Y = 0.0529 m).

To demonstrate if the accuracy observed is different for the different real-time positioning methods, a statistical analysis was carried out using an ANOVA (Table 2). This statistical analysis was carried out using the errors (difference between the estimated and the observed values) at each of the 26 CPs. This table shows the p-values obtained by using ANOVA, where a comparison of the four different methods was carried out using the X, Y, and Z axes errors as response variables.

Table 2

Comparison of real-time methods for the development of photogrammetric products using ANOVA

Response variable F-ratio p-value
Y-axis error 23.44 0.0000
X-axis error 78.83 0.0000
Z-axis error 33.86 0.0000

Note: A p-value less than 0.05 indicates that there is evidence of a difference between the methods used.

For the Y axis, the ANOVA showed a p-value less than 0.05 indicating that there is a significant difference between the different methods tested. Hence, a multiple-range test was carried out to recognize the method with the lowest error (the most accurate method). Using the LSD method, this study demonstrated the existence of two homogeneous groups (Figure 9). This statistical analysis validated that the RTK-CORS and RTK-CORS + PPK methods showed the lowest error for the Y-axis. The mean error observed for these last two methods was close to zero. The ANOVA for the X-axis also demonstrated a significant difference between the different methods used (Table 2). Based on the LSD multiple range test, the RTK-CORS + PPK method showed higher accuracy in the X-axis and the error values obtained with this method are closer to zero. According to this statistical analysis, higher mean errors were observed for the other three real-time methods used.

Figure 9 
               Confidence interval plot for the errors obtained in the photogrammetric products when using four different real-time positioning methods.
Figure 9

Confidence interval plot for the errors obtained in the photogrammetric products when using four different real-time positioning methods.

Regarding altimetry (Z axis), the ANOVA showed that there is a significant difference between the different methods used (Table 2). Figure 7c shows 2 homogeneous groups: one homogeneous group comprised three methods, RTK-CORS + PPK, RTK-DRTK2, and RTK-DRTK2 + GCP, which are statistically equal (p-value > 0.05). The mean altimetric error obtained using these three methods was close to zero. According to these results, a high altimetric accuracy was obtained when using these three methods. The RTK CORS method was statistically different from the other three methods and showed a mean altimetric error difference of 0.15 m.

4 Discussion

Different techniques and methodologies are emerging to provide highly accurate photogrammetric products. The accuracy of these photogrammetric products using UAVs depends directly on a set of parameters, which include the geometric resolution of the sensor system, the longitudinal and transverse overlaps of the images acquired, the angle of the images, the UAV speed flight, the radiometric resolution of the images, among other parameters [20].

Accuracy analysis of photogrammetric products obtained using UAVs is generally performed using GCPs [25,56]. However, using GCPs has drawbacks, such as the time-consuming process of GNSS measurement and relatively higher costs. To overcome these challenges, certain techniques are used. One such technique is direct georeferencing on UAV, which is an aerial triangulation method that eliminates the need for GCP. By employing direct georeferencing techniques, field surveys can be reduced, resulting in the production of topographic maps with high accuracy [23]. This study provides a detailed comparison of four direct georeferencing methods used for digital photogrammetry: RTK-CORS, RTK-CORS + PPK, RTK-DRTK2, and RTK-DRTK2 + GCP. Based on the statistical analysis, the RTK-CORS + PPK method showed higher accuracy in planimetry, where the differences between the estimated and measured values were close to zero. When the RTK-CORS + PPK method was used, an RMSExy value of 0.0164 m was obtained. The planimetric accuracy using the RTK-CORS + PPK method suggests that the quality of photogrammetric products obtained can be comparable to the accuracy of LiDAR products [57]. According to Famiglietti et al. [41], the error range of RTK drones can be accurately measured up to 4.0 cm when used in conjunction with GCP. In this study, the enhanced accuracy of the RTK-CORS + PPK method was obtained, which measured up to 1.64 cm with very satisfactory results. These results can be attributed to the combination of RTK and PPK techniques. When RTK and PPK are used together, both real-time corrections and post-processing refinement produce a more thorough error correction and optimization of the data, resulting in improved accuracy levels.

Regarding the altimetric accuracy, no statistical differences were found between the RTK-CORS + PPK, RTK-DRTK2, and RTK-DRTK2 + GCP methods (p-value > 0.05). In the literature, the altimetric accuracy is reported as a function of GSD. Some studies based on traditional photogrammetry methodologies using GCP have reported an accuracy between 1 and 3 GSD, while planar topography projects range from 1 to 4.5 GSD [58,59,60]. For complex topography, RMSE Z between 1.5 and 5 GSD are reported [39,61]. In the present study, an RMSE Z between 1 and 1.6 GSD was obtained, except for the RTK-CORS method which presented an RMSEZ close to 9 GSD. According to these results, the altimetric accuracy met the horizontal accuracy geometry at a scale level of 1:5,000 class and can be considered as high-precision geospatial data [23].

The maximum elevation difference of the photogrammetric products observed was 0.377 m when using the RTK-CORS method. This error can be considered as low vertical accuracy and could be related to some disadvantages or limitations observed with this technology. Since CORS transmits its real-time correction data via a cellular network, weak signal coverage could have compromised the availability and reliability of real-time corrections due to latency or delay in receiving the correction positioning data, reducing the accuracy of direct georeferencing. Despite this elevation difference, Elkhrachy [36] reported accuracies greater than 1 m for photogrammetric products that are not specific for topography or do not use GCP for ortho-rectification.

In this study, the influence of surface morphology on the planimetric and altimetric accuracy of RTK methods was evidenced. Greater elevation differences in altimetry were identified at CPs where slopes, ridges, valleys, bumps, and irregularities were present. Based on the results obtained in this study, surface irregularities introduced a change in photo scale and perspective distortions in the captured images. Therefore, the photo scale varied across the image in surfaces that showed significant relief or irregularities due to the changing distance between the camera and different parts of the surface. This variation in the photo scale affected the accurate measurement of distances and heights, leading to errors and inaccuracies in the resulting photogrammetric products [28,62].

Finally, several studies in the literature have proven that the use of UAVs with a large number of GCPs produces photogrammetric products with high accuracies, where RMSE X,Y,Z of 0.06 m can be achieved [36]. According to Agüera-Vega et al. [61], when more GCPs are available, more triangulation can be performed between GCPs and image pixels. The direct georeferencing methods used in our study did not use GCPs for the development of the photogrammetric products, except for the RTK-DRTK2 + GCP method which used six GCPs. Despite this situation, a high planimetric and altimetric accuracy was obtained, similar to the accuracy of the photogrammetric products obtained using UAVs with a large number of GCPs or the one obtained in topographic surveys. Therefore, direct georeferencing methods demonstrated their feasibility of use in digital aerial photogrammetry with the advantages of low cost, time conservation, and minimum fieldwork [63,64]. Thereby, the RTK-CORS + PPK method can be used with full security, replacing the traditional methodologies, as long as there is a CORS antenna correction and internet signal. However, in remote locations where there is no internet, the RTK-DRTK2 and RTK-DRTK2 + GCP methods are suggested.

5 Conclusion

This study assesses the effect of traditional and current measurement technologies and methods, such as RTK, PPK, CORS, and RTK + GCP on the altimetric and planimetric accuracy of orthophotos and DEMs. According to the results obtained, the four methods showed acceptable accuracies in the planimetry, with a maximum value of RMSE X,Y of 0.0529 m equivalent to almost 2.5 GSD. The RTK-CORS + PPK method was the most accurate in planimetry with an RMSE X,Y of 0.0164 m (equal to 0.74 GSD).

In altimetry, three real-time positioning methods (RTK-CORS + PPK, RTK-DRTK2, RTK-DRTK2 + GCP) showed highly accurate photogrammetric products, with RMSEZ values ranging from 0.0201 to 0.0334 m. Likewise, the results showed that the most accurate method was the RTK-CORS + PPK, which presented an RMSE X,Y,Z of 0.0372 m for the three coordinate axes, while the RTK-CORS method was the least accurate.

This study found that three RTK methods (CORS-RTK + PPK, RTK-D 2RTK, and RTK-D 2RTK + GCP) effectively acquired topographic data. The planimetric accuracy of these methods achieved less than one GSD, while the altimetric accuracy was better than 1.5 times the value of the GSD. The accuracies obtained with these methods are comparable with topographic surveys that are measured in a traditional way using total stations or GNSS antennas. The photogrammetric products presented a high accuracy in DEM and orthophotography, with the advantage of a higher level of automation of these processes.

The present study demonstrated that various photogrammetric works can be performed with the RTK-CORS + PPK method without using GCP since this method produced acceptable results with RMSE of approximately 1–2 GSD in planimetry and 1.5–2.5 GSD in altimetry. This statement does not imply the elimination of CPs from photogrammetric flights, as these points are considered essential to validate the results obtained.

Acknowledgements

The authors would like to thank the Universidad Autónoma de Sinaloa, specially to the Facultad de Ciencias de la Tierra y el Espacio (México) for providing Hardware and Software Technologies. This work was also carried out in collaboration with the Laboratorio Nacional CONAHCYT de Tecnologías de la Información Geoespacial para los Sistemas Socioecológicos Resilientes (LaNCTIGeSSR). The authors also acknowledge the Sistema Nacional de Investigadoras e Investigadores (SNII), México City, México.

  1. Funding information: This research has not received any external funding.

  2. Author contributions: Conceptualization: S.A.M.A. and J.G.R.P.; methodology: S.A.M.A., J.G.R.P., and W.P.R; validation: S.A.M.A., A.J.S.G, and Z.D.M.F.; formal analysis: S.A.M.A., J.G.R.P., and S.A.R.G.; investigation: S.A.M.A., J.G.R.P., and Z.D.M.F.; resources: S.A.M.A. and W.P.R.; data processing: S.A.M.A., J.G.R.P., Z.D.M.F., and A.J.S.G; writing: S.A.M.A., J.G.R.P., and W.P.R.; review and editing of the draft: S.A.R.G. and W.P.R.; visualization: Z.D.M.F. and A.J.S.G.; and supervision: S.A.M.A. and W.P.R. All authors have read and accepted the published version of this manuscript. The authors applied the SDC approach for the sequence of authors.

  3. Conflict of interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this article.

  4. Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

[1] La Salandra M, Miniello G, Nicotri S, Italiano A, Donvito G, Maggi G, et al. Generating UAV high-resolution topographic data within a FOSS photogrammetric workflow using high-performance computing clusters. Int J Appl Earth Observ Geoinf. 2021;105:102600. 10.1016/J.JAG.2021.102600.Search in Google Scholar

[2] Mancini F, Dubbini M, Gattelli M, Stecchi F, Fabbri S, Gabbianelli G. Using unmanned aerial vehicles (UAV) for high-resolution reconstruction of topography: the structure from motion approach on coastal environments. Remote Sens. 2013;5:6880–98. 10.3390/RS5126880.Search in Google Scholar

[3] Medvedev AA, Telnova NO, Kudikov AV, Alekseenko NA. Use of photogrammetric point clouds for the analysis and mapping of structural variables in sparse northern boreal forests. Sovremennye problemy distantsionnogo zondirovaniya Zemli iz kosmosa. 2020;17:150–63. 10.21046/2070-7401-2020-17-1-150-163.Search in Google Scholar

[4] Nuijten RJG, Coops NC, Watson C, Theberge D. Monitoring the structure of regenerating vegetation using drone-based digital aerial photogrammetry. Remote Sens. 2021;13:1942. 10.3390/RS13101942.Search in Google Scholar

[5] Hristov G, Kinaneva D, Georgiev G, Zahariev P, Kyuchukov P. An overview of the use of unmanned aerial vehicles for precision agriculture. Proceedings of the International Conference on Biomedical Innovations and Applications. Vol. 2020, BIA; 2020. p. 137–40. 10.1109/BIA50171.2020.9244519.Search in Google Scholar

[6] Montes de Oca A, Flores G. The AgriQ: A low-cost unmanned aerial system for precision agriculture. Expert Syst Appl. 2021;182:115163. 10.1016/J.ESWA.2021.115163.Search in Google Scholar

[7] Pecci A. Digital survey from drone in archaeology: potentiality, limits, territorial archaeological context and variables. IOP Conf Series: Mater Sci Eng. 2020;949:012075. 10.1088/1757-899X/949/1/012075.Search in Google Scholar

[8] Wright AE, Conlin DL, Shope SM. Assessing the accuracy of underwater photogrammetry for archaeology: a comparison of structure from motion photogrammetry and real time kinematic survey at the east key construction wreck. J Mar Sci Eng. 2020;8:849. 10.3390/JMSE8110849.Search in Google Scholar

[9] Jalinoos F, Amjadian M, Agrawal AK, Brooks C, Banach D. Experimental evaluation of unmanned aerial system for measuring bridge movement. J Bridge Eng. 2019;25:04019132. 10.1061/(ASCE)BE.1943-5592.0001508.Search in Google Scholar

[10] Khuc T, Nguyen TA, Dao H, Catbas FN. Swaying displacement measurement for structural monitoring using computer vision and an unmanned aerial vehicle. Measurement. 2020;159:107769. 10.1016/J.MEASUREMENT.2020.107769.Search in Google Scholar

[11] Zhao S, Kang F, Li J, Ma C. Structural health monitoring and inspection of dams based on UAV photogrammetry with image 3D reconstruction. Autom Constr. 2021;130:103832. 10.1016/J.AUTCON.2021.103832.Search in Google Scholar

[12] Oh YK. A study on the survey methodology in riverbed private use using integration drone photogrammetry and cadastral information. J Cadastre Land Informatix. 2017;47:135–44. 10.22640/LXSIRI.2017.47.2.135.Search in Google Scholar

[13] Laporte-Fauret Q, Marieu V, Castelle B, Michalet R, Bujan S, Rosebery D. Low-cost UAV for high-resolution and large-scale coastal dune change monitoring using photogrammetry. J Mar Sci Eng. 2019;7:63. 10.3390/JMSE7030063.Search in Google Scholar

[14] Zanutta A, Lambertini A, Vittuari L. UAV photogrammetry and ground surveys as a mapping tool for quickly monitoring shoreline and beach changes. J Mar Sci Eng. 2020;8:52. 10.3390/JMSE8010052.Search in Google Scholar

[15] Carrivick JL, Smith MW. Fluvial and aquatic applications of structure from motion photogrammetry and unmanned aerial vehicle/drone technology. WIREs Water. 2019;6:1–17. 10.1002/WAT2.1328.Search in Google Scholar

[16] Li T, Zhang B, Xiao W, Cheng X, Li Z, Zhao J. UAV-Based photogrammetry and LiDAR for the characterization of ice morphology evolution. IEEE J Sel Top Appl Earth Obs Remote Sens. 2020;13:4188–99. 10.1109/JSTARS.2020.3010069.Search in Google Scholar

[17] Śledź S, Ewertowski MW, Piekarczyk J. Applications of unmanned aerial vehicle (UAV) surveys and structure from motion photogrammetry in glacial and periglacial geomorphology. Geomorphology. 2021;378:107620. 10.1016/J.GEOMORPH.2021.107620.Search in Google Scholar

[18] Yao H, Qin R, Chen X. Unmanned aerial vehicle for remote sensing applications – a review. Remote Sens. 2019;11:1443. 10.3390/RS11121443.Search in Google Scholar

[19] Štroner M, Urban R, Reindl T, Seidl J, Brouček J. Evaluation of the georeferencing accuracy of a photogrammetric model using a quadrocopter with onboard GNSS RTK. Sensors. 2020;20:2318. 10.3390/S20082318.Search in Google Scholar PubMed PubMed Central

[20] Mora-Felix ZD, Sanhouse-Garcia AJ, Bustos-Terrones YA, Loaiza JG, Monjardin-Armenta SA, Rangel-Peraza JG. Effect of photogrammetric RPAS flight parameters on plani-altimetric accuracy of DTM. Open Geosci. 2020;12:1017–35. 10.1515/GEO-2020-0189/MACHINEREADABLECITATION/RIS.Search in Google Scholar

[21] Taddia Y, Stecchi F, Pellegrinelli A. Coastal mapping using DJI phantom 4 RTK in post-processing kinematic mode. Drones. 2020;4:9. 10.3390/DRONES4020009.Search in Google Scholar

[22] Pirti A. Evaluating the accuracy of post-processed kinematic (PPK) positioning technique. Geodesy Cartogr. 2021;47:66–70. 10.3846/GAC.2021.12269.Search in Google Scholar

[23] Syetiawan A, Gularso H, Kusnadi GI, Pramudita GN. Precise topographic mapping using direct georeferencing in UAV. IOP Conf Series: Earth Environ Sci. 2020;500:012029. 10.1088/1755-1315/500/1/012029.Search in Google Scholar

[24] Costa SMA, de Almeida L, de Moura NJ, Abreu MA, da Silva AL, Fortes LPS, et al. RBMC in real time via NTRIP and its benefits in RTK and DGPS surveys. Int Assoc Geodesy Symp. 2012;136:917–22. 10.1007/978-3-642-20338-1_115/COVER.Search in Google Scholar

[25] Zeybek M. Accuracy assessment of direct georeferencing UAV images with onboard global navigation satellite system and comparison of CORS/RTK surveying methods. Meas Sci Technol. 2021;32:065402. 10.1088/1361-6501/ABF25D.Search in Google Scholar

[26] Teppati Losè L, Chiabrando F, Giulio Tonolo F. Are measured ground control points still required in UAV based large scale mapping? assessing the positional accuracy of an RTK multi-rotor platform. International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences - ISPRS Archives. Int Soc Photogramm Remote Sens. 2020;43:507–14. 10.5194/ISPRS-ARCHIVES-XLIII-B1-2020-507-2020.Search in Google Scholar

[27] Türk T, Tunalioglu N, Erdogan B, Ocalan T, Gurturk M. Accuracy assessment of UAV-post-processing kinematic (PPK) and UAV-traditional (with ground control points) georeferencing methods. Environ Monit Assess. 2022;194:1–12. 10.1007/S10661-022-10170-0/TABLES/4.Search in Google Scholar

[28] Štroner M, Urban R, Seidl J, Reindl T, Brouček J. Photogrammetry using UAV-mounted GNSS RTK: georeferencing strategies without GCPs. Remote Sens. 2021;13:1336. 10.3390/RS13071336.Search in Google Scholar

[29] Eker R, Alkan E, Aydin AA. Comparative analysis of UAV-RTK and UAV-PPK methods in mapping different surface types. Eur J For Eng. 2021;7:12–25. 10.33904/EJFE.938067.Search in Google Scholar

[30] Cho JM, Lee BK. GCP and PPK utilization plan to deal with RTK signal interruption in RTK-UAV photogrammetry. Drones. 2023;7:265. 10.3390/DRONES7040265.Search in Google Scholar

[31] Tomaštík J, Mokroš M, Surový P, Grznárová A, Merganič J. UAV RTK/PPK method – an optimal solution for mapping inaccessible forested areas. Remote Sens. 2019;11:721. 10.3390/RS11060721.Search in Google Scholar

[32] Kefauver C, Zeybek M, Ta¸skayata¸skaya S, Elkhrachy I, Tarolli P. Improving the spatial accuracy of UAV platforms using direct georeferencing methods: an application for steep slopes. Remote Sens. 2023;15:2700. 10.3390/RS15102700.Search in Google Scholar

[33] Martínez-Carricondo P, Agüera-Vega F, Carvajal-Ramírez F, Mart Inez-Carricondo P, AgUera-Vega F, Carvajal-Ram Irez F. Accuracy assessment of RTK/PPK UAV-photogrammetry projects using differential corrections from multiple GNSS fixed base stations. 2023;38:1–16. 10.1080/10106049.2023.2197507.Search in Google Scholar

[34] Pfeifer N, Glira P, Briese C. Direct georeferencing with on board navigation components of light weight UAV platforms. Int Arch Photogramm, Remote Sens Spat Inf Sci. 2012;XXXIX-B7:487–92. 10.5194/ISPRSARCHIVES-XXXIX-B7-487-2012.Search in Google Scholar

[35] Barba S, Barbarella M, Di Benedetto A, Fiani M, Gujski L, Limongiello M. Accuracy assessment of 3D photogrammetric models from an unmanned aerial vehicle. Drones. 2019;3:79. 10.3390/DRONES3040079.Search in Google Scholar

[36] Elkhrachy I. Accuracy assessment of low-cost unmanned aerial vehicle (UAV) photogrammetry. Alex Eng J. 2021;60:5579–90. 10.1016/J.AEJ.2021.04.011.Search in Google Scholar

[37] Salas López R, Terrones Murga RE, Silva-López JO, Rojas-Briceño NB, Gómez Fernández D, Oliva-Cruz M, et al. Accuracy assessment of direct georeferencing for photogrammetric applications based on UAS-GNSS for high andean urban environments. Drones. 2022;6:388. 10.3390/DRONES6120388.Search in Google Scholar

[38] Nesbit PR, Hubbard SM, Hugenholtz CH. Direct georeferencing UAV-SfM in high-relief topography: accuracy assessment and alternative ground control strategies along steep inaccessible rock slopes. Remote Sens. 2022;14:490. 10.3390/RS14030490.Search in Google Scholar

[39] Sanz-Ablanedo E, Chandler JH, Rodríguez-Pérez JR, Ordóñez C. Accuracy of unmanned aerial vehicle (UAV) and SfM photogrammetry survey as a function of the number and location of ground control points used. Remote Sens. 2018;10:1606. 10.3390/RS10101606.Search in Google Scholar

[40] Yu JJ, Kim DW, Lee EJ, Son SW. Determining the optimal number of ground control points for varying study sites through accuracy evaluation of unmanned aerial system-based 3D point clouds and digital surface models. Drones. 2020;4:49. 10.3390/DRONES4030049.Search in Google Scholar

[41] Famiglietti NA, Cecere G, Grasso C, Memmolo A, Vicari A. A test on the potential of a low cost unmanned aerial vehicle RTK/PPK solution for precision positioning. Sensors. 2021;21:3882. 10.3390/S21113882.Search in Google Scholar

[42] Zhang H, Aldana-Jague E, Clapuyt F, Wilken F, Vanacker V, Van Oost K. Evaluating the potential of post-processing kinematic (PPK) georeferencing for UAV-based structure-from-motion (SfM) photogrammetry and surface change detection. Earth Surf Dyn. 2019;7:807–27. 10.5194/ESURF-7-807-2019.Search in Google Scholar

[43] Taddia Y, González-García L, Zambello E, Pellegrinelli A. Quality assessment of photogrammetric models for façade and building reconstruction using DJI phantom 4 RTK. Remote Sens. 2020;12:3144. 10.3390/RS12193144.Search in Google Scholar

[44] Zumberge JF, Heflin MB, Jefferson DC, Watkins MM, Webb FH. Precise point positioning for the efficient and robust analysis of GPS data from large networks. J Geophys Res: Solid Earth. 1997;102:5005–17. 10.1029/96JB03860.Search in Google Scholar

[45] Erkoç MH, Doğan U. Datum definition for geodetic vertical velocity field derived from GNSS observations: a case study in western and southern Turkey. Bull Geophys Oceanogr. 2023;1:135–48. 10.4430/BGO00412.Search in Google Scholar

[46] Cavalheri EP, Dos Santos MC. Improved kinematic precise point positioning performance with the use of map constraints. J Appl Geodesy. 2020;14:191–204. 10.1515/JAG-2019-0034/MACHINEREADABLECITATION/RIS.Search in Google Scholar

[47] Zhang K, Okazawa H, Hayashi K, Hayashi T, Fiwa L, Maskey S. Optimization of ground control point distribution for unmanned aerial vehicle photogrammetry for inaccessible fields. Sustainability. 2022;14:9505. 10.3390/SU14159505.Search in Google Scholar

[48] Liu Y, Han K, Rasdorf W. Evaluation of UAS flight configuration factors and their impacts on photogrammetric survey accuracy. Construction Research Congress 2022: Computer Applications, Automation, and Data Analytics - Selected Papers from Construction Research Congress. 2-B, 2022. p. 225–34. 10.1061/9780784483961.024.Search in Google Scholar

[49] Liu Y, Han K, Rasdorf W. Assessment and prediction of impact of flight configuration factors on UAS-based photogrammetric survey accuracy. Remote Sens. 2022;14:4119. 10.3390/RS14164119.Search in Google Scholar

[50] McMahon C, Mora OE, Starek MJ. Evaluating the performance of sUAS photogrammetry with PPK positioning for infrastructure mapping. Drones. 2021;5:50. 10.3390/DRONES5020050.Search in Google Scholar

[51] Enterprise D DJI Phantom 4 RTK - Plataforma empresarial DJI en MEA n.d. https://thedronecentre.ae/dji-p4rtk/(, accessed March 2 2024).Search in Google Scholar

[52] Hu Y, Zhang N, Wang Q. Three-dimensional reconstruction and geo-positioning accuracy analysis based on UAV remote sensing images. Proceedings – 2022 2nd Asia-Pacific Conference on Communications Technology and Computer Science. 2022, ACCTCS; 2022. p. 127–31. 10.1109/ACCTCS53867.2022.00033.Search in Google Scholar

[53] Hossein J, Corcoran P. Accurate depth map estimation from small motions. Venecia, Italia: IEEE; 2017. p. 2453–61.10.1109/ICCVW.2017.289Search in Google Scholar

[54] Wang B, Shi W, Liu E. Robust methods for assessing the accuracy of linear interpolated DEM. Int J Appl Earth Observ Geoinf. 2015;34:198–206. 10.1016/J.JAG.2014.08.012.Search in Google Scholar

[55] Tien Bui D, Long NQ, Bui X-N, Nguyen V-N, Van Pham C, Van Le C, et al. Lightweight unmanned aerial vehicle and structure-from-motion photogrammetry for generating digital surface model for open-pit coal mine area and its accuracy assessment. In: Tien Bui D, Ngoc Do A, Bui HB, Hoang ND, editors. Advances and applications in geospatial technology and earth resources. Cham: Springer; 2018. p. 17–33. 10.1007/978-3-319-68240-2_2.Search in Google Scholar

[56] Uysal M, Toprak AS, Polat N. DEM generation with UAV photogrammetry and accuracy analysis in Sahitler hill. Measurement. 2015;73:539–43. 10.1016/J.MEASUREMENT.2015.06.010.Search in Google Scholar

[57] Elaksher A, Ali T, Alharthy A. A quantitative ASSESSMENT of LIDAR data accuracy. Remote Sens. 2023;15:442. 10.3390/RS15020442.Search in Google Scholar

[58] Ewertowski MW, Tomczyk AM, Evans DJA, Roberts DH, Ewertowski W. Operational framework for rapid, very-high resolution mapping of glacial geomorphology using low-cost unmanned aerial vehicles and structure-from-motion approach. Remote Sens. 2019;11:65. 10.3390/RS11010065.Search in Google Scholar

[59] Jiménez-Jiménez SI, Ojeda-Bustamante W, Marcial-Pablo MDJ, Enciso J. Digital terrain models generated with low-cost UAV photogrammetry: methodology and accuracy. ISPRS Int J Geo-Inform. 2021;10:285. 10.3390/IJGI10050285.Search in Google Scholar

[60] Santise M, Fornari M, Forlani G, Roncella R. Evaluation of DEM generation accuracy from UAS imagery. Int Arch Photogramm, Remote Sens Spat Inf Sci. 2014;XL–5:529–36. 10.5194/ISPRSARCHIVES-XL-5-529-2014.Search in Google Scholar

[61] Agüera-Vega F, Carvajal-Ramírez F, Martínez-Carricondo P, Sánchez-Hermosilla López J, Mesas-Carrascosa FJ, García-Ferrer A, et al. Reconstruction of extreme topography from UAV structure from motion photogrammetry. Measurement. 2018;121:127–38. 10.1016/J.measurement.2018.02.062.Search in Google Scholar

[62] Forlani G, Dall’Asta E, Diotri F, Cella di UM, Roncella R, Santise M. Quality assessment of DSMs produced from UAV flights georeferenced with on-board RTK positioning. Remote Sens. 2018;10:311. 10.3390/RS10020311.Search in Google Scholar

[63] Rabah M, Basiouny M, Ghanem E, Elhadary A. Using RTK and VRS in direct geo-referencing of the UAV imagery. NRIAG J Astron Geophys. 2019;7:220–6. 10.1016/J.NRJAG.2018.05.003.Search in Google Scholar

[64] Stott E, Williams RD, Hoey TB. Ground control point distribution for accurate kilometre-scale topographic mapping using an RTK-GNSS unmanned aerial vehicle and SfM photogrammetry. Drones. 2020;4:55. 10.3390/DRONES4030055.Search in Google Scholar

Received: 2023-11-30
Revised: 2024-04-21
Accepted: 2024-04-24
Published Online: 2024-09-23

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Theoretical magnetotelluric response of stratiform earth consisting of alternative homogeneous and transitional layers
  3. The research of common drought indexes for the application to the drought monitoring in the region of Jin Sha river
  4. Evolutionary game analysis of government, businesses, and consumers in high-standard farmland low-carbon construction
  5. On the use of low-frequency passive seismic as a direct hydrocarbon indicator: A case study at Banyubang oil field, Indonesia
  6. Water transportation planning in connection with extreme weather conditions; case study – Port of Novi Sad, Serbia
  7. Zircon U–Pb ages of the Paleozoic volcaniclastic strata in the Junggar Basin, NW China
  8. Monitoring of mangrove forests vegetation based on optical versus microwave data: A case study western coast of Saudi Arabia
  9. Microfacies analysis of marine shale: A case study of the shales of the Wufeng–Longmaxi formation in the western Chongqing, Sichuan Basin, China
  10. Multisource remote sensing image fusion processing in plateau seismic region feature information extraction and application analysis – An example of the Menyuan Ms6.9 earthquake on January 8, 2022
  11. Identification of magnetic mineralogy and paleo-flow direction of the Miocene-quaternary volcanic products in the north of Lake Van, Eastern Turkey
  12. Impact of fully rotating steel casing bored pile on adjacent tunnels
  13. Adolescents’ consumption intentions toward leisure tourism in high-risk leisure environments in riverine areas
  14. Petrogenesis of Jurassic granitic rocks in South China Block: Implications for events related to subduction of Paleo-Pacific plate
  15. Differences in urban daytime and night block vitality based on mobile phone signaling data: A case study of Kunming’s urban district
  16. Random forest and artificial neural network-based tsunami forests classification using data fusion of Sentinel-2 and Airbus Vision-1 satellites: A case study of Garhi Chandan, Pakistan
  17. Integrated geophysical approach for detection and size-geometry characterization of a multiscale karst system in carbonate units, semiarid Brazil
  18. Spatial and temporal changes in ecosystem services value and analysis of driving factors in the Yangtze River Delta Region
  19. Deep fault sliding rates for Ka-Ping block of Xinjiang based on repeating earthquakes
  20. Improved deep learning segmentation of outdoor point clouds with different sampling strategies and using intensities
  21. Platform margin belt structure and sedimentation characteristics of Changxing Formation reefs on both sides of the Kaijiang-Liangping trough, eastern Sichuan Basin, China
  22. Enhancing attapulgite and cement-modified loess for effective landfill lining: A study on seepage prevention and Cu/Pb ion adsorption
  23. Flood risk assessment, a case study in an arid environment of Southeast Morocco
  24. Lower limits of physical properties and classification evaluation criteria of the tight reservoir in the Ahe Formation in the Dibei Area of the Kuqa depression
  25. Evaluation of Viaducts’ contribution to road network accessibility in the Yunnan–Guizhou area based on the node deletion method
  26. Permian tectonic switch of the southern Central Asian Orogenic Belt: Constraints from magmatism in the southern Alxa region, NW China
  27. Element geochemical differences in lower Cambrian black shales with hydrothermal sedimentation in the Yangtze block, South China
  28. Three-dimensional finite-memory quasi-Newton inversion of the magnetotelluric based on unstructured grids
  29. Obliquity-paced summer monsoon from the Shilou red clay section on the eastern Chinese Loess Plateau
  30. Classification and logging identification of reservoir space near the upper Ordovician pinch-out line in Tahe Oilfield
  31. Ultra-deep channel sand body target recognition method based on improved deep learning under UAV cluster
  32. New formula to determine flyrock distance on sedimentary rocks with low strength
  33. Assessing the ecological security of tourism in Northeast China
  34. Effective reservoir identification and sweet spot prediction in Chang 8 Member tight oil reservoirs in Huanjiang area, Ordos Basin
  35. Detecting heterogeneity of spatial accessibility to sports facilities for adolescents at fine scale: A case study in Changsha, China
  36. Effects of freeze–thaw cycles on soil nutrients by soft rock and sand remodeling
  37. Vibration prediction with a method based on the absorption property of blast-induced seismic waves: A case study
  38. A new look at the geodynamic development of the Ediacaran–early Cambrian forearc basalts of the Tannuola-Khamsara Island Arc (Central Asia, Russia): Conclusions from geological, geochemical, and Nd-isotope data
  39. Spatio-temporal analysis of the driving factors of urban land use expansion in China: A study of the Yangtze River Delta region
  40. Selection of Euler deconvolution solutions using the enhanced horizontal gradient and stable vertical differentiation
  41. Phase change of the Ordovician hydrocarbon in the Tarim Basin: A case study from the Halahatang–Shunbei area
  42. Using interpretative structure model and analytical network process for optimum site selection of airport locations in Delta Egypt
  43. Geochemistry of magnetite from Fe-skarn deposits along the central Loei Fold Belt, Thailand
  44. Functional typology of settlements in the Srem region, Serbia
  45. Hunger Games Search for the elucidation of gravity anomalies with application to geothermal energy investigations and volcanic activity studies
  46. Addressing incomplete tile phenomena in image tiling: Introducing the grid six-intersection model
  47. Evaluation and control model for resilience of water resource building system based on fuzzy comprehensive evaluation method and its application
  48. MIF and AHP methods for delineation of groundwater potential zones using remote sensing and GIS techniques in Tirunelveli, Tenkasi District, India
  49. New database for the estimation of dynamic coefficient of friction of snow
  50. Measuring urban growth dynamics: A study in Hue city, Vietnam
  51. Comparative models of support-vector machine, multilayer perceptron, and decision tree ‎predication approaches for landslide ‎susceptibility analysis
  52. Experimental study on the influence of clay content on the shear strength of silty soil and mechanism analysis
  53. Geosite assessment as a contribution to the sustainable development of Babušnica, Serbia
  54. Using fuzzy analytical hierarchy process for road transportation services management based on remote sensing and GIS technology
  55. Accumulation mechanism of multi-type unconventional oil and gas reservoirs in Northern China: Taking Hari Sag of the Yin’e Basin as an example
  56. TOC prediction of source rocks based on the convolutional neural network and logging curves – A case study of Pinghu Formation in Xihu Sag
  57. A method for fast detection of wind farms from remote sensing images using deep learning and geospatial analysis
  58. Spatial distribution and driving factors of karst rocky desertification in Southwest China based on GIS and geodetector
  59. Physicochemical and mineralogical composition studies of clays from Share and Tshonga areas, Northern Bida Basin, Nigeria: Implications for Geophagia
  60. Geochemical sedimentary records of eutrophication and environmental change in Chaohu Lake, East China
  61. Research progress of freeze–thaw rock using bibliometric analysis
  62. Mixed irrigation affects the composition and diversity of the soil bacterial community
  63. Examining the swelling potential of cohesive soils with high plasticity according to their index properties using GIS
  64. Geological genesis and identification of high-porosity and low-permeability sandstones in the Cretaceous Bashkirchik Formation, northern Tarim Basin
  65. Usability of PPGIS tools exemplified by geodiscussion – a tool for public participation in shaping public space
  66. Efficient development technology of Upper Paleozoic Lower Shihezi tight sandstone gas reservoir in northeastern Ordos Basin
  67. Assessment of soil resources of agricultural landscapes in Turkestan region of the Republic of Kazakhstan based on agrochemical indexes
  68. Evaluating the impact of DEM interpolation algorithms on relief index for soil resource management
  69. Petrogenetic relationship between plutonic and subvolcanic rocks in the Jurassic Shuikoushan complex, South China
  70. A novel workflow for shale lithology identification – A case study in the Gulong Depression, Songliao Basin, China
  71. Characteristics and main controlling factors of dolomite reservoirs in Fei-3 Member of Feixianguan Formation of Lower Triassic, Puguang area
  72. Impact of high-speed railway network on county-level accessibility and economic linkage in Jiangxi Province, China: A spatio-temporal data analysis
  73. Estimation model of wild fractional vegetation cover based on RGB vegetation index and its application
  74. Lithofacies, petrography, and geochemistry of the Lamphun oceanic plate stratigraphy: As a record of the subduction history of Paleo-Tethys in Chiang Mai-Chiang Rai Suture Zone of Thailand
  75. Structural features and tectonic activity of the Weihe Fault, central China
  76. Application of the wavelet transform and Hilbert–Huang transform in stratigraphic sequence division of Jurassic Shaximiao Formation in Southwest Sichuan Basin
  77. Structural detachment influences the shale gas preservation in the Wufeng-Longmaxi Formation, Northern Guizhou Province
  78. Distribution law of Chang 7 Member tight oil in the western Ordos Basin based on geological, logging and numerical simulation techniques
  79. Evaluation of alteration in the geothermal province west of Cappadocia, Türkiye: Mineralogical, petrographical, geochemical, and remote sensing data
  80. Numerical modeling of site response at large strains with simplified nonlinear models: Application to Lotung seismic array
  81. Quantitative characterization of granite failure intensity under dynamic disturbance from energy standpoint
  82. Characteristics of debris flow dynamics and prediction of the hazardous area in Bangou Village, Yanqing District, Beijing, China
  83. Rockfall mapping and susceptibility evaluation based on UAV high-resolution imagery and support vector machine method
  84. Statistical comparison analysis of different real-time kinematic methods for the development of photogrammetric products: CORS-RTK, CORS-RTK + PPK, RTK-DRTK2, and RTK + DRTK2 + GCP
  85. Hydrogeological mapping of fracture networks using earth observation data to improve rainfall–runoff modeling in arid mountains, Saudi Arabia
  86. Petrography and geochemistry of pegmatite and leucogranite of Ntega-Marangara area, Burundi, in relation to rare metal mineralisation
  87. Prediction of formation fracture pressure based on reinforcement learning and XGBoost
  88. Hazard zonation for potential earthquake-induced landslide in the eastern East Kunlun fault zone
  89. Monitoring water infiltration in multiple layers of sandstone coal mining model with cracks using ERT
  90. Study of the patterns of ice lake variation and the factors influencing these changes in the western Nyingchi area
  91. Productive conservation at the landslide prone area under the threat of rapid land cover changes
  92. Sedimentary processes and patterns in deposits corresponding to freshwater lake-facies of hyperpycnal flow – An experimental study based on flume depositional simulations
  93. Study on time-dependent injectability evaluation of mudstone considering the self-healing effect
  94. Detection of objects with diverse geometric shapes in GPR images using deep-learning methods
  95. Behavior of trace metals in sedimentary cores from marine and lacustrine environments in Algeria
  96. Spatiotemporal variation pattern and spatial coupling relationship between NDVI and LST in Mu Us Sandy Land
  97. Formation mechanism and oil-bearing properties of gravity flow sand body of Chang 63 sub-member of Yanchang Formation in Huaqing area, Ordos Basin
  98. Diagenesis of marine-continental transitional shale from the Upper Permian Longtan Formation in southern Sichuan Basin, China
  99. Vertical high-velocity structures and seismic activity in western Shandong Rise, China: Case study inspired by double-difference seismic tomography
  100. Spatial coupling relationship between metamorphic core complex and gold deposits: Constraints from geophysical electromagnetics
  101. Disparities in the geospatial allocation of public facilities from the perspective of living circles
  102. Research on spatial correlation structure of war heritage based on field theory. A case study of Jinzhai County, China
  103. Formation mechanisms of Qiaoba-Zhongdu Danxia landforms in southwestern Sichuan Province, China
  104. Magnetic data interpretation: Implication for structure and hydrocarbon potentiality at Delta Wadi Diit, Southeastern Egypt
  105. Deeply buried clastic rock diagenesis evolution mechanism of Dongdaohaizi sag in the center of Junggar fault basin, Northwest China
  106. Application of LS-RAPID to simulate the motion of two contrasting landslides triggered by earthquakes
  107. The new insight of tectonic setting in Sunda–Banda transition zone using tomography seismic. Case study: 7.1 M deep earthquake 29 August 2023
  108. The critical role of c and φ in ensuring stability: A study on rockfill dams
  109. Evidence of late quaternary activity of the Weining-Shuicheng Fault in Guizhou, China
  110. Extreme hydroclimatic events and response of vegetation in the eastern QTP since 10 ka
  111. Spatial–temporal effect of sea–land gradient on landscape pattern and ecological risk in the coastal zone: A case study of Dalian City
  112. Study on the influence mechanism of land use on carbon storage under multiple scenarios: A case study of Wenzhou
  113. A new method for identifying reservoir fluid properties based on well logging data: A case study from PL block of Bohai Bay Basin, North China
  114. Comparison between thermal models across the Middle Magdalena Valley, Eastern Cordillera, and Eastern Llanos basins in Colombia
  115. Mineralogical and elemental analysis of Kazakh coals from three mines: Preliminary insights from mode of occurrence to environmental impacts
  116. Chlorite-induced porosity evolution in multi-source tight sandstone reservoirs: A case study of the Shaximiao Formation in western Sichuan Basin
  117. Predicting stability factors for rotational failures in earth slopes and embankments using artificial intelligence techniques
  118. Origin of Late Cretaceous A-type granitoids in South China: Response to the rollback and retreat of the Paleo-Pacific plate
  119. Modification of dolomitization on reservoir spaces in reef–shoal complex: A case study of Permian Changxing Formation, Sichuan Basin, SW China
  120. Geological characteristics of the Daduhe gold belt, western Sichuan, China: Implications for exploration
  121. Rock physics model for deep coal-bed methane reservoir based on equivalent medium theory: A case study of Carboniferous-Permian in Eastern Ordos Basin
  122. Enhancing the total-field magnetic anomaly using the normalized source strength
  123. Shear wave velocity profiling of Riyadh City, Saudi Arabia, utilizing the multi-channel analysis of surface waves method
  124. Effect of coal facies on pore structure heterogeneity of coal measures: Quantitative characterization and comparative study
  125. Inversion method of organic matter content of different types of soils in black soil area based on hyperspectral indices
  126. Detection of seepage zones in artificial levees: A case study at the Körös River, Hungary
  127. Tight sandstone fluid detection technology based on multi-wave seismic data
  128. Characteristics and control techniques of soft rock tunnel lining cracks in high geo-stress environments: Case study of Wushaoling tunnel group
  129. Influence of pore structure characteristics on the Permian Shan-1 reservoir in Longdong, Southwest Ordos Basin, China
  130. Study on sedimentary model of Shanxi Formation – Lower Shihezi Formation in Da 17 well area of Daniudi gas field, Ordos Basin
  131. Multi-scenario territorial spatial simulation and dynamic changes: A case study of Jilin Province in China from 1985 to 2030
  132. Review Articles
  133. Major ascidian species with negative impacts on bivalve aquaculture: Current knowledge and future research aims
  134. Prediction and assessment of meteorological drought in southwest China using long short-term memory model
  135. Communication
  136. Essential questions in earth and geosciences according to large language models
  137. Erratum
  138. Erratum to “Random forest and artificial neural network-based tsunami forests classification using data fusion of Sentinel-2 and Airbus Vision-1 satellites: A case study of Garhi Chandan, Pakistan”
  139. Special Issue: Natural Resources and Environmental Risks: Towards a Sustainable Future - Part I
  140. Spatial-temporal and trend analysis of traffic accidents in AP Vojvodina (North Serbia)
  141. Exploring environmental awareness, knowledge, and safety: A comparative study among students in Montenegro and North Macedonia
  142. Determinants influencing tourists’ willingness to visit Türkiye – Impact of earthquake hazards on Serbian visitors’ preferences
  143. Application of remote sensing in monitoring land degradation: A case study of Stanari municipality (Bosnia and Herzegovina)
  144. Optimizing agricultural land use: A GIS-based assessment of suitability in the Sana River Basin, Bosnia and Herzegovina
  145. Assessing risk-prone areas in the Kratovska Reka catchment (North Macedonia) by integrating advanced geospatial analytics and flash flood potential index
  146. Analysis of the intensity of erosive processes and state of vegetation cover in the zone of influence of the Kolubara Mining Basin
  147. GIS-based spatial modeling of landslide susceptibility using BWM-LSI: A case study – city of Smederevo (Serbia)
  148. Geospatial modeling of wildfire susceptibility on a national scale in Montenegro: A comparative evaluation of F-AHP and FR methodologies
  149. Geosite assessment as the first step for the development of canyoning activities in North Montenegro
  150. Urban geoheritage and degradation risk assessment of the Sokograd fortress (Sokobanja, Eastern Serbia)
  151. Multi-hazard modeling of erosion and landslide susceptibility at the national scale in the example of North Macedonia
  152. Understanding seismic hazard resilience in Montenegro: A qualitative analysis of community preparedness and response capabilities
  153. Forest soil CO2 emission in Quercus robur level II monitoring site
  154. Characterization of glomalin proteins in soil: A potential indicator of erosion intensity
  155. Power of Terroir: Case study of Grašac at the Fruška Gora wine region (North Serbia)
  156. Special Issue: Geospatial and Environmental Dynamics - Part I
  157. Qualitative insights into cultural heritage protection in Serbia: Addressing legal and institutional gaps for disaster risk resilience
Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/geo-2022-0650/html
Scroll to top button