Next Article in Journal
Sustainable Healthcare Education as a Practice of Governmentality?
Next Article in Special Issue
Mathematical Modelling of Biogas Production in a Controlled Landfill: Characterization, Valorization Study and Energy Potential
Previous Article in Journal
The Impact of Surface Water Seepage on Seismicity and Rockbursting in Mines
Previous Article in Special Issue
Characterization of Hydrologic Sand and Dust Storm Sources in the Middle East
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Monitoring Coastal Changes and Assessing Protection Structures at the Damietta Promontory, Nile Delta, Egypt, to Secure Sustainability in the Context of Climate Changes

by
Hesham M. El-Asmar
1,* and
Maysa M. N. Taha
2
1
Department of Geology, Damietta University, Damietta 34517, Egypt
2
Department of Geology, Helwan University, Helwan 11795, Egypt
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(22), 15415; https://doi.org/10.3390/su142215415
Submission received: 22 August 2022 / Revised: 7 November 2022 / Accepted: 8 November 2022 / Published: 20 November 2022

Abstract

:
The Damietta Promontory is a distinct coastal region in the Nile Delta Egypt, which comprises several communities with strategic economic projects. The promontory has experienced numerous inundation crises due to anthropogenic intervention and/or sea level rise (SLR). The recorded rate of erosion detected is from −18 to −53 m/yr., and −28 to −210 m/yr. along the promontory’s western and eastern coasts, respectively, with a total loss of about 3 km during the past century. It is critical to ensure sustainability of this coastal region in case of future climate changes and expected SLR; accordingly, the state has implemented a long-term plan of coastal protection. The current study updates the coastal changes and assesses the efficiency of the protection structures. For such study, Ikonos satellite images of 1 m high-resolution were acquired on 30 July 2014 and 10 August 2022, respectively. These were compared to multitemporal Landsat images dated 30 June 2015, 29 September 1987, 15 October 1984, and the Landsat 4 MSS images dated 20 October 1972. The results confirm the presence of accretion along the western jetty of the Damietta Harbor with an average of +10.91 m/yr., while erosion of −4.7 m/yr. was detected at the east of the eastern harbor jetty. At the detached breakwaters along Ras El-Bar, an accretion of +4 m/yr. was detected, and then erosion was measured westward to the tip of the detached breakwaters with an average of −1.77 m/yr. At the eastern coast of the promontory, eastward erosion was recorded with rates of −44.16, −34.33, and −20.33 m/yr., respectively, then the erosion stopped after the construction of the seawall. The current study confirms the efficiency of the detached breakwaters and seawalls as coastal protection structures. However, the seawalls lack swimming-friendly long, wide beaches like those found on the detached breakwaters. The groins seem ineffective with rips and reversed currents like those at Ras El -Bar. To develop a fishing community at the Manzala triangle similar in nature to Venice, it is recommended to extend the seawall to 12 km and then construct detached breakwaters eastward to the El-Diba inlet. To secure sustainability of the coast, a continuous maintenance of the protection structures to keep their elevations between 4–6 m above sea level (a.s.l.) is a critical task, in order to reduce the potential risks that could arise from a tsunami, with sand nourishment as a preferred strategy.

1. Introduction

The Nile Delta is a classical globally studied river mouth, formed from the interplay of the River Nile and the Mediterranean Sea across the geologic time scale. The Nile Delta has seen significant development in recent years as part of the national strategy for sustainable development 2030 [1]. The Nile Delta coast is a subaerial part of the fluvio-marine Bilqas Formation of the Holocene age [2]. The Holocene is composed of fluvial sediments of prodelta clays and delta-front sands and silts, with marine and subaerial coastal sediments such as beach ridges and coastal sand dunes [2]. The structural sequence shows a continuous prograding delta with a coarsening upward sequence [2,3]. Study of the coastal geomorphology reflects that the delta beaches are described as having fine to extremely fine sands and being fully dissipative, divergent, smooth, wide, gently sloped beach faces [4]. Due to SLR and anthropogenic intervention, the Nile Delta coast is currently neither dissipative nor divergent [3,4], and new wave and current patterns have resulted in new coastal segmentations related to the direction of the shoreline, the relationship to induced waves and currents, and the implemented protection structures. The studies of LC/LU of the coastal Nile Delta reveals the presence of geomorphic units including the beach ridges, coastal dunes, and interdune depressions with sabkhas [5,6]. Most of these features were obliterated through the coastal development [5,6]. The beaches show cuspate features, steep berm gradients, and shell distribution [4], and these features, supported with the results of satellite images, help in reliable interpretation of coastal dynamics and understanding the role of the interplay of different factors along the coast [4].
The Damietta Promontory (31°27′55.15′′ N and 31°2′13.51′′ E to 31°22′22.8′′ N and 32°2′32.66′′ E) was selected for the current study a number of reasons, including the fact that it supports the economy through industries such as tourism and fishing and is expected to attract additional investment. This area is prone to seasonal inundation, which may threaten ongoing development. The Damietta branch of the Nile is like an artery that runs through the body of the Damietta Promontory, and meets the Mediterranean Sea at the Damietta mouth (Figure 1a), dividing the promontory into two coastal segments (Figure 1). The western one is oriented NE–SW, comprising Ras El-Bar Resort and Damietta Harbor. The eastern one is oriented NW–SE, and comprises Ezbet El-Borg City and the Manzala Lagoon (Figure 1a,c), [7]. The Damietta branch, along with the historical branches, discharges more than 150 million m3 of sediment into the Mediterranean annually [3]. Such huge discharge has led to accretion, with the coast protruding into the sea for 2–3 km [4,5]. Of these sediments, 0.6 to 1.8 million m3/yr. moves along the coast, forming the promontory [5]. The extension of the promontory into the sea reflects the quantity of discharge reaching the coast from the Damietta branch relative to the drift by longshore transport [5].
The wave propagations at Ras El-Bar come principally from the NW (78%), NE (21%), and SW (1%) directions [8]. The resultant longshore currents have SW and NE directions, and are produced from N, NNW, and NW (50°) and WWN (40°) waves, respectively. On the other hand, wave propagations at Ezbet El-Borg come from the N, NNW, and NE (90°), resulting in a unidirectional NE longshore current without reversals (Figure 1a) [7,8]. Additionally, ground subsidence of up to 5 mm/yr. [9,10] and SLR of 5–6.9 mm/yr [11] both have contributed to coastline retreat [12,13] and a generally termed relative sea level rise [14].
Ras El-Bar and Ezbet El-Borg cities (Figure 1a,b) comprise 10% of the Damietta province’s total population of about 1,600,000, of whom 25,000 live at Ras El Bar, and make their livelihood by serving from 250,000 to one million tourists and visitors a year at resort facilities and by providing summer entertainment. On the other hand, Ezbet El-Burg City is inhabited by 100,000 people, of whom 10,000 are anglers (1% of Egypt’s total). It is one of the world’s most famous cities in the fields of fishing and shipping [15]. It is the base of Egypt’s largest fishing boat fleet, including boats of the traditional felucca type. The city itself hosts 65% of the Egyptian naval fleet, approximately 926 fishing boats, and receives a number of other ships, bringing the number of boats operating along the Damietta coast to 1350 boats. The city is also home to a sardine-canning factory operated by the Edfina Company [15].
The fishing sector provides the main source of income for the locals. Many of the fishing boats venture far into the Eastern Mediterranean and the Red Sea [15]. In addition, the city has interesting historical values with several archaeology sites and possible underwater antiquities. The annual inundation of these cities by winter storms represents a critical crisis for these benefited people and influences their way of life. A possible SLR through climate changes may represent a catastrophic episode for these vulnerable areas and residences [12,13,16].
In 1981, construction of the Damietta Harbor (Figure 1c) began in order to meet the needs of increasing international trade in the Eastern Mediterranean [17]. It was also planned to build the harbor inland to minimize the effects of waves and currents and to avoid storm conditions [18]. The location of the harbor was selected along an embayment with minimum wave energy and shoreline changes [19,20]. The navigation channel of the harbor was protected by construction of two jetties, the western (WJ) and the eastern (EJ), later extended to 1300 m and 600 m, respectively (Figure 1c) [21]. The annual net rate of littoral drift on the western side of the harbor is about 1.43 × 105 m3 (accretion with an average 2.13 m/yr.) while the annual net rate of littoral drift on the eastern side is about 2.54 × 105 m3 (−92 m erosion on average over the past 45 years) (Figure 1c) [20,22]. Therefore, the harbor is threatened by problems of erosion along its eastern jetty and sedimentation along its navigation channel. The orientation of the navigation channel and its depth represent a sinking area receiving more west-to-east drifted sediments, reducing the channel depth form 14 m to10 m, which impedes the navigation of ships along the channel [23]. It is intended to construct a new section of harbor to increase its capacity to accommodate the greater cargo trade and to ensure a depth of 18 m in the navigation channel to allow passage of giant ships, required for transferring the products of the liquid natural gas (LNG) plant, the methane and formaldehyde industry, and for a free industrial zone [24].
In order to assess the coastal dynamic and shoreline changes, the current study’s main novelty is the use of high-resolution Geo-Eye (Ikonos) satellite images that compare to other Landsat images, and a real field investigation including identification of beach and nearshore geo-features expressing the coastal dynamic and shoreline changes. Such high-resolution results will offer ± 1 m accuracy of shoreline measurements, and allow an excellent opportunity to assess the current situation and the efficiency of the protection works. It will be a prototype for sustainable development of relevant fragile coasts suffering erosion problems. Among the current study’s objectives are to: (1) update rates of shoreline changes; (2) assess the implemented protection structures based on a comparison of shoreline changes; (3) demonstrate their efficiency to protect the coast from inundation; (4) determine the role of SLR, land subsidence, and the anthropogenic intervention in creating the problem of coastal inundation; (5) propose a role model of investment of the islands of the Manzala triangle; (6) confirm the role of Integrated Coastal Zone Management (ICZM) as a sustainable approach in dealing with coastal areas.

2. Data and Methods

The resolution of satellite images is increasing: observations are becoming more frequent, and this trend is expected to continue in the near future [25]. However, images always contain some uncertainty, which is one of the most common challenges [25]. Thus, quantitative uncertainty information associated with the data is required. It is also a critical scientific effort for both data producers and end users, as the process will reveal error characteristics that will guide further improvements in data production and rational data use [26]. To avoid uncertainty of remote sensing results, the selected images should not include more than 10% cloud. Geometric correction is used to correct the variation between the actual location coordinates and the raw image data on the ground or target image [27]. Geometric correction was performed through image-to-image geo-referencing in the Universal Transverse Mercator Projection (UTM/zone 36 WGS 84). In the current study, at least 10 prominent well-distributed ground control points (GCP) were selected in the master images, located in the other images, and then a nearest-neighbor resampling method was applied in addition to ground verification of several well-known points that were easily identified and observed in the field. In the current study, three points were detected and verified; one at the eastern jetty of the Damietta Harbor, the second at the tip of the eighth breakwaters, and the third at the lighthouse of the western jetty of the Damietta branch of the Nile. It was considered for remote sensing applications to use images taken at various dates prior to and after the construction of the protection structures to demonstrate the efficiency of the structures in protecting the coast concerned. We needed data from at least two dates to track changes in coastline locations. The impact of shifting shorelines is not as obvious, and a large set of data is required to track changes in coastline locations [27,28]. Atmospheric correction was applied using Fast Line-of-sight Atmospheric Analysis of Spectral Hypercubes (FLAASH) as a method in ENVI for retrieving spectral reflectance from hyperspectral radiance images. FLAASH incorporates the MODTRAN radiation transfer model to compensate for atmospheric effects [27,29].
Multitemporal Landsat images were used in the current study (Table 1) to identify environmental changes: (1) the Multispectral Landsat 8-OLI (dated 30 June 2015, path 176, row 38), (2) the Landsat 5 TM (dated 29 September 1987 and 15 October 1984, path 176, row 38), and (3) the Landsat 4 MSS (dated 20 October 1972, path 189, row 38). The data were obtained from the USGS’s GLOVIS and EARTH EXPLORER websites [30]. All image scenes were processed using ENVI 5.1 and arc GIS 10.4 software. These were compared with the results deduced from the high-resolution 1 m ground IKONOS satellite images acquired on 30 July 2014 and 10 August 2022, respectively. The image data were acquired from the World Imagery Wayback—ArcGIS Living Atlas sites [31]. Images of the two shorelines at different dates were extracted using photo-interpretation techniques covering 8 years (2014 and 2022, respectively).
The DSAS tool was used to determine the rate of shoreline displacement (Figure 2). DSAS is an add-on for Esri ArcGIS that enables users to compute shoreline change rate data from numerous coastlines with various dates [32,33]. This feature enables the creation of transects along the shore that are positioned perpendicular to the baseline at the desired user spacing [34]. A new transect layer was made using the Digital Shoreline Analysis System (DSAS) tool, with a transect spacing of 100 m. DSAS extracted the required data from the input baseline, shoreline, and transect files covering a 5 km long section of shoreline.
The methodology for calculating rates was based on observing how shorelines changed over time. A quick and effective way for estimating the rate of shoreline changes is the end point rate (EPR) (Figure 2). The mean annual rate of shoreline change (meters per year) for the oldest and most recent coastlines is calculated by dividing the distance of shoreline movement by the amount of elapsed time [35]. EPR is one of six methods used by DSAS to calculate rate-of-change statistics for a time series of shoreline vector data [36]. EPR is used to assess the annual rate of shoreline alterations for periods between 2014 and 2022. Similar assessments and predictions of shoreline change using EPR and linear LRR tools in DSAS have been utilized on the Andhra Pradesh coastline in India [37].

3. Results and Discussion

The global SLR in most of the world’s oceans and seas, including the Mediterranean, may have serious consequences on the low-lying land of fragile coasts, including the Nile Delta. The studies of coastal changes along these vulnerable coasts are usually of global interest. Such studies update our understanding of SLR in the context of climate change, land subsidence, and other anthropogenic interventions. This section is dedicated to disclosing the current high-resolution Geo-Eye (Ikonos) satellite images compared with the historical results in order to update the records, determine the changes, understand the situations that led to the implementation of several protection structures, and then to recommend the preferred model for coastal stability. The recorded results are listed in Figure 3 and compared with the current study’s results.

3.1. The Coastal Dynamics and Protection Structures

To measure the coastal dynamics and shoreline changes we must take into consideration some aspects of interest, among which are the physical conditions and the nature of coastal materials and whether these are rocks or soft sediments, and the orientation of the coastline in relation to waves and currents [4]. While the western side of the promontory is oriented NE–SW, and most waves come from NW propagation, these have resulted in two opposing SW and NE longshore currents. The eastern one is oriented NW–SE and the waves oriented N, NNW, and NE have resulted in a unidirectional NE longshore current (Figure 1a) [7,8]. The presence of bidirectional currents may little impede the drifting of sediment, which does not apply in the unidirectional current [4]. This may explain why erosion along both sides of the promontory appears different: while the recorded rate of erosion on the promontory’s western coast was between −18 and −53 m/yr., at the same time, rates of −28 to −210 m/yr. were detected on the promontory’s eastern coast [5,7,8].
Since 1902, coastal changes have been attributed to the development of numerous barrages and dams to regulate water flow, which altered the river system. The promontory then underwent erosion and shoreline retreat (Figure 1a, Figure 3 and Figure 4a), with rates of −53 m/yr. and −210 m/yr., on the promontory’s western and eastern coasts, respectively [38,39].
From 1909 to 1980, these rates changed to −18 m/yr. and −28 m/yr., respectively [5,7]. Beginning in 1972, the promontory’s erosion rates increased to −23 m/yr. and −52 m/yr. on the west and east promontory coasts, respectively [7]. This was strongly linked to the entrapment of sediment discharges after the construction of the Aswan High Dam [40,46,49,50]. This confirms that the problem of erosion along the Nile Delta coast is old and started as early as the start of the 20th century, which enforced the state to react to the problem and build several protection structures. Some of these structures were damaging to the coast due to several physical and geological conditions.
The history of protection structures began around 1929, when the Damietta estuary was constantly subjected to sedimentation and shifting of the river mouth because of longshore drift, causing shoaling in the estuary and preventing fishing boats from entering across the river mouth to the sea and vice versa [46,51]. This prompted the construction of two jetties of 240 m and 290 m in length (Figure 4d), to keep the river mouth open [7,8,51], which were later renewed in 2007. Between 1984 and 1991, the promontory retreated by an average of −196 m (e.g., −28 m/yr.), and total erosion of the promontory from 1984 to 2015 was calculated as −1631 m (Figure 4a), with an average erosion rate of −20.61 m/yr. [7,8]. Other results detected between 1971 and 1990 were that the shoreline retreated −24 m/yr. to the west, and −41 m/yr. to the east of the promontory [5,43], and the lengths decreased by −1.9 km to the west, and −3.3 km to the east [5,43]. Rates of erosion were recorded at −15 m/yr., −39 to −11 m/yr. [16,51], and at −47.99 m/yr., −43 m/yr. and −53.30 m/yr. near the promontory’s tip [41,42,51] (Figure 3), which led to the building of a seawall at the eastern promontory.
A 6 km long and 4 m a.s.l. seawall was constructed along the eastern side of the Damietta Promontory in 2000 [42,43,52] (Figure 4a,b). It is made up of prefabricated 4 to 7 tons dolos alternated with carbonate and basaltic rocks to protect the promontory from erosion (Figure 4c) [47,49]. However, severe erosion was detected at the down-drift of the seawall [53] (Figures 4a and 7c for comparison), which led the state to add 6 km in length to the wall to cover the erosional coast at the tip of the Manzala lagoon (the Manzala triangle, Figure 1b). The erosion then transferred eastward, eroding a considerable part of the Damietta Spit (Figure 1b and Figure 2a). Rates of erosion between −24 m/yr. to −65 m/yr. were detected at the sand spit from 1972 to 2018 (Figure 3), [44,45,48].
In 1963, a concrete seawall was constructed (Figure 5a) along the western side of the promontory at the Ras El-Bar resort to reduce erosion along the southern edge of the western jetty of the Damietta Promontory. This seawall was subjected to sediment undermining, which caused a partial collapse (Figure 5a). The seawall was later rebuilt as a rip-rap structure that stands around 3.5 m a.s.l. (Figure 5b) [48,53]. To the west of this seawall, three 120 m long concrete groins were built in 1971. These were constructed as prefabricated blocks that were protected by concrete dolos (Figure 5c), but when erosion persisted, the groin’s blocks broke and fell. Two of these groins were replaced in 2010, and they were extended to form a small bay with a tiny exit (Figure 5d) [4,49]. Some of the drifting sands ended up in this bay, forming a new sandy beach with a highly safe swimming location.
Until 1990 Ras El-Bar coast was still eroded, with rates up to −83.20 m/yr. recorded prior to the construction of the detached breakwaters [4,49] (Figure 3). Four detached breakwaters were built between 1991 and 1994. By the year 2000, four more breakwaters were added, bringing the total number to eight (about 400 m from the shoreline and a depth of 4 m b.s.l.). Each breakwater is 200 m long, with a 200 m space, and 3 m elevation a.s.l. (Figure 5e). Beyond these breakwaters, a shadow area was created with reduced (only 25%) wave energy (Figure 5e), ensuring coastal accretion of +4 m/yr. (Figure 6 and Figure 7b), and a total area of 0.46 km 2 (Figure 1d).
The beach in front of the summer residences inundated during 1986 was restored as a wide flat beach during 2000 (Figure 6a–d). The detached breakwater structures showed rates of accretion of +12–14 m/yr. [16,53], and up to 21 m/yr. [49] (Figure 3). According to El-Asmar [49] a new reversing eddy current formed westward of the detached breakwaters resulted in erosion on the coast with an annual rate of −4.7 m/yr., and a total erosional area of 0.57 km2 (Figure 1d). The development of a cuspate beach with shell accumulations and rip currents (Figure 5f,g) confirm the evidence of such erosion [4].
Inspection of this area of the coast shows evidences of sediments undermining sands, resulting in local subsidence (Figure 6e) of the detached breakwater with considerable sedimentation around the breakers’ body (Figure 6f). At the present time, the elevation of these breakers reached in some places less than 1 m (Figure 6e–f). Annual sand nourishment, as well as maintenance and support of the breakwaters with replenishment materials to secure their elevation at not less than 6 m, are critical tasks.
To protect the Damietta Harbor from physical conditions (waves, currents, and littoral drift), two jetties were built in 1981; these jetties are the western (WJ) and the eastern jetties (EJ). From 1984 to 1987, the shoreline west of the western jetty accreted at an average rate of +13.00 m/yr. (Figure 3). From 1987 to 2015, it displayed continuous accretion of +12.4 m/yr. (Figure 3). Later, the accretion continued at a rate of +15 m/yr. at the west of the harbor’s western jetty [16,53]. Other rates were reported as +9–16 m/yr. and an area of +0.244 km2 up to 2018 [8,16,22,44,47,51]. During the years 1984 to 1987, the coastline at the eastern jetty increased by an average of +10.13 m/yr. (Figure 1c EJ; Figure 7a Area A). A few years later, the erosion along this coast reached an average of −4.7 m/yr. (Figure 3), which threat the inland body of the eastern harbor’s jetty. This prompted the 2014 announcement of construction of a new project by Egyptian General Authority of Coastal Protection.

3.2. High Resolution Geo-Eye (Ikonos) Satellite Images

More than 20 years of monitoring this coastal area has been supported by research, ICZM, and coastal geomorphology, and supplemented with data from Geo-Eye (Ikonos) satellite images taken on 30 July 2014, and 2 August 2022, with a ground resolution of 1 m. In addition, a massive database was compiled through long-term research along the Nile Delta’s coast. These enabled a more accurate assessment of the current coastal situation at the Damietta Promontory, as well as an evaluation of the protection structures put in place to ensure sustainability in the face of climate change and sea-level rise. The data collected from the Geo-Eye (Ikonos) satellite images of the western and eastern coasts of the Damietta Promontory from 2014 to 2022 are listed in Figure 3, and satellite images are illustrated in Figure 7. It is worth mentioning that no results have been detected for the area west of the harbor’s western jetty (Figure 1c WJ), due to the ongoing construction of the harbor extension.
During 2014, four Y groins were built, three of the groins designed to be 170 m in length and the fourth 120 m, with 400 m space in between (area A, Figure 7a). The satellite images ensure the actual lengths at 188, 177, 149, and 125 m from east to west, respectively (Figure 5h and Figure 7a), whereas the gaps set at 430 m. Erosion areas between groins east of the eastern jetty are −14,434m2, −15,762m2, and −18,456m2 with an average of −2405, −2626, −3076 m2/yr. (Figure 3 and Figure 7a). While areas of accretion were observed at the groin bodies (Figure 7a) with areas of +2345 m2, 2590 m2, 7517 m2 and 10,587 m2 (Figure 7a, Area A) and an average of +391, +432, +1253, +1765 m2/yr. (Figure 3). These are about to create tombolos. It was the first time the Y groins have been used as a protection structure along this coastal region. It is too early to evaluate their performance, with new figures on erosion/accretion. However, sand nourishment for the groin’s gaps is important [54,55]. The favored solutions include ideas that combines both hard and soft structures, such as detachable breakwaters with sand nourishment to secure beaches safe for swimming [54,55,56].
The current study has proven that the groins cannot protect the beach when rip currents are present (Figure 5g) or from two opposing currents. Sand undermining has caused fracture and collapse (Figure 5c) and groins have failed to maintain the shoreline against erosion. A similar conclusion emerged when 15 groins were constructed along the Rosetta Promontory, which negatively affected the coastal morphology of the Rosetta Promontory [57,58], with an average of −30.8 m/yr. lost [50]. There are 47 traditional groins covering a total length of 3652 m and another 31 groins covering a coastal length of 4421 m at R. Morto Nuovo, Italy. Most of these groins are either completely or partially submerged, and the coast lost resort and tourism facilities [59].
The detached breakwaters seem to be relatively effective in protecting the coast [54,56,60] and freeing up more beaches for swimming. Both the submerged and emerging detached breakwaters are successful protective measures that keep beaches safe, which is a requirement of the tourism industry and recreational beaches [56,60]. The constraint of the detached breakwater is the adjustment of the shadow area beyond the breakers’ body to prevent tombolo formation, as seen in Tunisia [59] or at Baltim and Port Said (Figure 8a,b). If they comply with environmental regulations and address drowning issues, detached breakwaters meet communities’ preferences for long-term coastal protection [61] while also providing recreational facilities needed by the tourism industry.
While sand nourishment has been unsuccessful along the Port Said coast (Figure 8c), the current study recommends the use of sand nourishment, which is the least expensive method of coastal protection [45,54]. The success of sand nourishment depends on the coastal nature (coastal extension and beach gradient), the presence of plenty of sand (unit sand volume m3/m, per meter of the beach length) and a critical point is the size of the sand used compared with the native sediments. Unsuccessful sand nourishment cases were mentioned in Israel [62], and explained as a result of the presence of wide coasts with strong waves that dredge most sands, and the use of small-grain sand—the size should be 1.5–2 times that of the native beach sands [62]—insufficient sand sources due to environmental restrictions of quarrying, and the lack of maintenance and replenishment [62]. On the Dutch coast, sand nourishment is considered a strategic approach and a state trend for coastal protection. By 2000, the area of sand was increased from 6.4 to 12 million m3/yr., adding to the coastline 432 km, which enabled the coastal zone to stay in equilibrium with sea level rise [63].
It is interesting to shed some light on the problem of coastal sand dune quarrying. Natural sand dunes and backshores along the coast of the Mediterranean region act as natural defense measures, reducing erosion and shoreline retreat [64]. Removal of these dunes is a risky course [28]. Most of the dunes along the Nile Delta, between Baltim and Damietta, have been removed for urbanization [65], including the construction of recreational resorts and new communities [66,67]. Such anthropogenic intervention, while claiming to be sustainable development and investment in the tourism industry, is actually a major threat to the environment and the stability of the coast. Due to the lack of awareness of ICZM, several strategic projects were built in inconvenient coastal areas. Focusing on human demands for entertainment and investments such as tourism-accommodating beaches and recreational facilities in face of strategic industries imply risks and pollution. Such imbalance causes deterioration of highly vulnerable systems, compromising both ecosystem integrity and health, as well as tourism itself [68].
Finally, a discussion about the possible inundation scenarios of the northern Nile Delta coast seems very interesting. Recent research has postulated two scenarios of 0.5 m and 1 m SLR of future inundated land areas and the affected populations. The extreme scenario of SLR to 1 m will affect about 3900 km2 of cropland, 1280 km2 of vegetation, 205 km2 of wetland, 146 km2 of urban areas at the northern borders, and cause more than 6 million people to lose their houses [69]. However, a more optimistic scenario is based on the measurements of a SLR gauge and referred to a value ranges between 6.7–6.9 mm/yr. [70]. An interpolation process has been performed for the SLR rates up to year 2050 (a time period of 35 yeas) has been calculated at 0.24 m. In addition to the average subsidence rate (calculated at Damietta Promontory to be 2.6 mm/yr., which appears consistent with the global positioning system rate of 3.5 mm/yrm. Accordingly, the areas vulnerable to inundation by year 2050 will be 38.40 km2, 3.80 km2, 5.20 km2, and 2.60 km2 for the urban, agricultural lands, fishing farms, and bare areas, respectively [70].
Tenths of papers have been devoted to the effect of SLR and land subsidence since 1988, when Stanley and his colleagues [9,10,11] published their pioneering studies on the Nile Delta subsidence scenarios, with suggested inundated areas. Since then several scenarios have been published, estimating future inundation of hundreds of kilometres and the migration of millions of people from their first homes to other areas to the south [11,12,13]. A 35 years period has passed (1988 to 2022) during which a SLR equal to 0.33 m. is supposed to inundate 1/3 of the suggested areas, a case that has not yet occurred. The present situation undoubtfully forced us to conclude that such a gradual increase in SLR is a secondary player and the predicted 2050 [70] scenario will be similar to that of Stanley [9,10,11] unless a new strong and rapid player, such as Tsunami [71], enters the game. Therefore, the anthropogenic factors are strongly involved. Such as compaction of the Holocene sediments (8000 to 2500) [13], and urban-induced loading at major cities (12 to 20 mm/yr.). The major cities such as Cairo, Tanta, Mahala, Mansoura, Damietta, and Port Said, show estimated subsidence are around 6.4 ± 0.4 mm/year, 4.0 ± 0.6 mm/year, 4.8 ± 1.0 mm/year, 10.0 ± 1.2 mm/yr., 10.3 ± 1.6 mm/yr., and 4.9 ± 1.6 mm/yr. respectively [72,73]. In addition, the role of groundwater overexploitation (16 to 20 mm/yr.) that recorded at newly reclaimed lands and most agricultural lands (like Menoufia governorate) [74], and high subsidence rates (up to 0.7 mm/year) over onshore gas fields, particularly the Abu Madi [74]. The localization of subsidence signals on InSAR and GPS imited at the big cities [75], and about 15.56% of the coastal cities are affected by subsidence [76], the subsidences related to anthropogenic activities seem to be more effective rather than SLR [76].

4. Conclusions

Owing to its historical values, economic resources of oil and gas and dense population, the Nile Delta is a traditional global model for delta research. Several scenarios of inundation due to SLR are emerging, however; inspection of several areas of the coast confirms the role of anthropogenic interventions in coastal erosion problems rather than SLR and land subsidence. In unawareness with the ICZM several economic projects, resorts, and profession communities were constructed in inconvenient coastal areas, which were later subjected to erosion. These led the state to implement a plan for coastal protection. The current study used high-resolution satellite images to detect the shoreline changes at the Damietta Promontory and to assess the efficiency of the implemented protection structures in securing sustainability of the coast against SLR and climate changes.
  • The current study confirms the efficiency of detached breakwaters in protecting the coast, with annual sand nourishment preferred in order to secure wide and gently sloped beaches convenient for swimming. However, the dimensions of the breaker bodies, gaps, and shadow areas must be taken into considerations.
  • Seawalls, on the other hand, are effective in protecting the coast-facing non-available beaches. The case presented was at the promontory facing the Manzala triangle, where the seawall may offer stability of the coast against most storms, which allows better investments for the islands along the Manzala triangle in a model similar in nature to that of Venice, Italy.
  • The seawall at the Damietta Promontory should be extended to the El-Diba Inlet, the ancient mouth of the Nile’s Mendisian branch. Additionally, we encourage the presence of detached breakwaters to offer water circulation within the Manzala triangle and restore beaches for the new resort communities.
  • Sand nourishment is the preferred way for protection; this depends on the presence of a suitable source of sand of the right grain size.
  • Due to the absence of continuous maintenance, several areas of collapses are observed and parts of the breaker body reached an elevation of less than 1 m a.s.l. Such a situation represents a great risk and continuous maintenance is a critical task to keep the elevation of the breakers not less than 6 m a.s.l. in order to impede tsunamic waves.
  • The groins failed in protecting the coast in cases where there are multidirectional currents or rip currents perpendicular to the coast due to the resultant sand undermining, resulting in collapse as seen at Ras El-Bar.
  • It is still too early for the Y groins to be evaluated; however, these create a new pattern of erosion between groins and accretion along the groins’ bodies that will transform the groins to tombolos.
  • The protection structures must be built to cover an entire cell or sector, not just a small segment. A segment of coast between two bodies of water or outlying headlands is referred to as a segment in this definition.
  • Building the protection structures in an opposite direction to currents is not recommended. Such a situation creates unusual waves and currents that may threaten the works.
  • The current study confirms the role of anthropogenic intervention rather than SLR by comparing the 35 year time period from 1988 postulated by Stanley and another 35 year period of a new scenario for SLR through to 2050. In fact, nothing has changed along the coast since 1988, except the construction of protecting structures. The historical changes are linked to the entrapment of the discharge rather than action of SLR in addition to some anthropogenic activities along the coast near big cities. However, a sudden tsunami is not ruled out.

Author Contributions

Conceptualization, Investigation, Writing original draft, H.M.E.-A.; Remote Sensing Investigation and Software Processing M.M.N.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available on request from the corresponding author.

Acknowledgments

I would like to thank the unidentified reviewers who provided assistance throughout the review process. Many thanks to the guest editors who worked tirelessly to get this issue published. I would like to thank Mansoura University, Egypt for scientific support through funding the publishing fees.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moawad, M.B. Recent landscape changes at Damietta Promontory, Nile Delta-Egypt. Egypt. J. Remote Sens. Space Sci. 2021, 24, 525–535. [Google Scholar] [CrossRef]
  2. Coutellier, V.; Stanley, D.J. Late Quaternary Stratigraphy and Paleogeography of the Eastern Nile Delta Shelf. Mar. Geol. 1987, 77, 257–275. [Google Scholar] [CrossRef]
  3. Sestini, G. Nile Delta: A Review of Depositional Environments and Geological History. Geol. Soc. 1989, 41, 99–127. [Google Scholar] [CrossRef]
  4. El-Asmar, H.M.; Taha, M.M.N.; El-Sorogy, A.S. Morphodynamic Changes as an Impact of Human Intervention at the Ras El-Bar-Damietta Harbor Coast, NW Damietta Promontory, Nile Delta, Egypt. J. Afr. Earth Sci. 2016, 124, 323–339. [Google Scholar] [CrossRef]
  5. Frihy, O.E.; Khafagy, A.A. Climate and Human Induced Changes in Relation to Shoreline Migration Trends at the Nile Delta Promontories, Egypt. CATENA 1991, 18, 197–211. [Google Scholar] [CrossRef]
  6. Masria, A.; Nadaoka, K.; Negm, A.; Iskender, M. Detection of Shoreline and Land Cover Changes around Rosetta Promontory, Egypt, Based on Remote Sensing Analysis. Land 2015, 4, 216–230. [Google Scholar] [CrossRef] [Green Version]
  7. Frihy, O.E.; Debes, E.A.; El Sayed, W.R. Processes Reshaping the Nile delta Promontories of Egypt: Pre- and Post-Protection. Geomorphology 2003, 53, 263–279. [Google Scholar] [CrossRef]
  8. Frihy, O.E.; Lawrence, D. Evolution of the Modern Nile Delta Promontories: Development of Accretional Features during Shoreline Retreat. Environ. Geol. 2004, 46, 914–931. [Google Scholar] [CrossRef]
  9. Stanley, D.J. Recent Subsidence and Northeast Tilting of the Nile Delta, Egypt. Mar. Geol. 1990, 94, 147–154. [Google Scholar] [CrossRef]
  10. Stanley, D.J.; Warne, A.G. Nile Delta: Recent Geological Evolution and Human Impact. Science 1993, 260, 628–634. [Google Scholar] [CrossRef]
  11. Stanley, J.D.; Clemente, P.L. Increased land Subsidence and Sea-Level Rise are Submerging Egypt’s Nile Delta Coastal Margin. GSA Today 2016, 27, 4–11. [Google Scholar] [CrossRef] [Green Version]
  12. Marriner, N.; Flaux, C.; Morhangem, C.; Kaniewski, D. Nile Delta’s sinking past: Quantifiable Links with Holocene Compaction and Climate-Driven Changes in Sediment Supply. Geology 2012, 40, 1083–1086. [Google Scholar] [CrossRef]
  13. Becker, R.H.; Sultan, M. Land Subsidence in the Nile Delta: Inferences from Radar Interferometry. Holocene 2009, 19, 949–954. [Google Scholar] [CrossRef]
  14. El-Asmar, H.M. Short term coastal changes along Damietta-Port Said coast Northeast of the Nile Delta. Egypt. J. Coast. Res. 2002, 18, 433–441. [Google Scholar]
  15. Abaza, J.; Mostafa, M.; El-Fekki, A. The Forgotten Fishermen of Ezbet El-Borg. Daily News, 14 April 2015; 1/9–9/9. [Google Scholar]
  16. Aly, M.H.; Giardino, J.R.; Klein, A.G.; Zebker, H.A. InSAR study of shoreline change along the Damietta Promontory, Egypt. J. Coast. Res. 2012, 28, 1263–1269. [Google Scholar] [CrossRef]
  17. ASRT. Sedimentation in Damietta Harbor; Academy of Scientific Research and Technology Final Report; ASRT: Cairo, Egypt, 1988. [Google Scholar]
  18. Tetra Tech. Shoreline Master Plan for the Nile Delta Coast. Prog. Rep. 1984, 1, 143. [Google Scholar]
  19. Sogreah, M. Effects on the construction of the Port of Damietta on the evolution of the littoral drift. Consult. Rep. 1982, 9, 35/1202. [Google Scholar]
  20. AboZed, B.I. Effects of Waves and Currents on the Siltation Problem of Damietta harbor, Nile Delta coast, Egypt. Mediterr. Mar. Sci. 2007, 8, 33–47. [Google Scholar] [CrossRef] [Green Version]
  21. El-Asmar, H.M. Impact of protection structures on Physical and Sedimentary Parameters along the Damietta coastal area, Nile Delta, Egypt. J. Sedimentol. Egypt. 1995, 3, 111–124. [Google Scholar]
  22. Ezzeldin, M.M.; Rageh, O.S.; Saad, M.E. Assessment impact of the Damietta Harbor (Egypt) and its deep navigation channel on adjacent shorelines. J. Integr. Coast. Zone Manag. 2020, 20, 265–281. [Google Scholar] [CrossRef]
  23. El-Asmar, H.M.; White, K. Changes in coastal sediment transport processes due to construction of New Damietta Harbor, Nile Delta, Egypt. Coast. Eng. 2002, 46, 127–138. [Google Scholar] [CrossRef]
  24. Deabes, E.A. Sedimentation Processes at the Navigation Channel of the Liquefied Natural Gas (LNG) Port, Nile Delta, Egypt. Int. J. Geosci. 2010, 1, 14–20. [Google Scholar] [CrossRef] [Green Version]
  25. Stein, A.; Ge, Y.; Fabris-Rotelli, I. Introduction to the Special Issue “Uncertainty in Remote Sensing Image Analysis”. Remote Sens. 2018, 10, 1975. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, W.; Hall-Beyer, M.; Wu, C.; Fang, W.; Nsengiyumva, W. Uncertainty Problems in Image Change Detection. Sustainability 2020, 12, 274. [Google Scholar] [CrossRef]
  27. Wu, W.; Mhaimeed, A.S.; Al-Shafie, W.M.; Ziadat, F.; Dhehibi, B.; Nangia, V.; De Pauw, E. Mapping soil salinity changes using remote sensing in Central Iraq. Geoderma Reg. 2014, 2, 21–31. [Google Scholar] [CrossRef]
  28. Taha, M.M.N. Monitorig Changing Position of the Shoreline along Burullus–Baltim, Nile Delta, Egypt, and the Role of Coastal Sand dunes as a Natural Defense Against Erosion in a Geoheritage Site: Remote sensing application. Egypt. J. Geol. 2020, 64, 111–121. [Google Scholar]
  29. Do, A.T.; Vries, S.D.; Stive, M.J. The Estimation and Evaluation of Shoreline Locations, Shoreline-Change Rates, and Coastal Volume Changes Derived from Landsat Images. J. Coast. Res. 2019, 35, 56–71. [Google Scholar] [CrossRef] [Green Version]
  30. Available online: http://earthexplorer.usgs.gov/ (accessed on 21 August 2022).
  31. Available online: https://livingatlas.arcgis.com/wayback/#active=47471&ext=-115.34940,36.03763,-115.24760,36.09036 (accessed on 21 August 2022).
  32. Thieler, E.R.; Himmelstoss, E.A.; Zichichi, J.L.; Ergul, A. The Digital Shoreline Analysis System (DSAS) Version 4.0—An ArcGIS Extension for Calculating Shoreline Change; Open-File Report, 2008-1278; U.S. Geological Survey: Reston, VA, USA, 2009. [Google Scholar] [CrossRef]
  33. Dewidar, K.; Bayoumi, S. Forecasting Shoreline Changes along the Egyptian Nile Delta Coast using Landsat Image Series and Geographic Information System. Environ. Monit. Assess. 2021, 193, 429. [Google Scholar] [CrossRef]
  34. Fountoulis, I.; Vassilakis, E.; Mavroulis, S.; Alexopoulos, J.; Dilalos, S.; Erkeki, A. Synergy of Tectonic Geomorphology, Applied Geophysics and Remote Sensing Techniques Reveals New Data for Active Extensional Tectonism in NW Peloponnese (Greece). Geomorphology 2015, 237, 52–64. [Google Scholar] [CrossRef]
  35. Thieler, E.R.; Hammar-Klose, E.S. National Assessment of Coastal Vulnerability to Sea-Level Rise, U.S. Atlantic Coast; U.S. Geological Survey: Reston, VA, USA, 1999; Volume 1, pp. 99–593. [Google Scholar]
  36. Elkafrawy, S.B.; Basheer, M.A.; Mohamed, H.M.; Naguib, D.M. Applications of remote sensing and GIS techniques to evaluate the effectiveness of coastal structures along Burullus headland-Eastern Nile Delta, Egypt. Egypt. J. Remote Sens. Space Sci. 2021, 24, 247–254. [Google Scholar] [CrossRef]
  37. Basheer, A.K.K.; Chandra, A.P. Assessment and Prediction of Shoreline Change using Multi-temporal Satellite Data and Geostatistics: A Case Study on the Eastern Coast of India. J. Water Clim. Chang. 2022, 13, 147. [Google Scholar] [CrossRef]
  38. E1-Fishawi, N.M.; Badr, A.A. Volumetric changes of nearshore sediments between Rosettaand Damietta, Egypt. INQUA Comm. Quat. Shorel. Newsl. 1989, 11, 39–42. [Google Scholar]
  39. Fanos, A.M.; Khafagy, A.A.; El Kady, M.M. Variations of the Nile River discharge and sediment regime during the periodfrom 1800 to 2000 and its effects on the Nile Delta coast. In Proceedings of the 8th International Symposium on River Sedimentation, Cairo, Egypt, 3–5 November 2001. [Google Scholar]
  40. Frihy, O.E. Sea-level rise and shoreline retreat of the Nile Delta Promontories, Egypt. Nat. Hazards 1992, 5, 65–81. [Google Scholar] [CrossRef]
  41. El-Asmar, H.M.; El-Kafrawy, S.; Taha, M.M.N. Monitoring coastal changes along Damietta promontory and the barrier beach toward Port Said east of the Nile Delta. Egypt. J. Coast. Res. 2014, 30, 993–1005. [Google Scholar] [CrossRef]
  42. Ali, E.; El-Magd, I.A. Impact of Human Interventions and Coastal Processes along the Nile Delta Coast, Egypt during the Past Twenty-Five Years. Egypt. J. Aquat. Res. 2016, 42, 1–10. [Google Scholar] [CrossRef]
  43. Dewidar, K.; Frihy, O.E. Automated Techniques for Quantification of Beach Change Rates using Landsat Series along the North-Eastern Nile Delta, Egypt. J. Oceanog. Mar. Sci. 2010, 1, 28–39. [Google Scholar]
  44. Abou Samra, R.M.; Ali, R.R. Applying DSAS Tool to Detect Coastal Changes along the Nile Delta Egypt. Egypt. J. Remote Sesing Space Sci. 2021, 24, 463–470. [Google Scholar] [CrossRef]
  45. Frihy, O.E.; Komar, P.D. Long-term Shoreline Changes and the Concentration of Heavy Minerals in Beach Sands of the Nile Delta, Egypt. Mar. Geol. 1993, 115, 253–261. [Google Scholar] [CrossRef]
  46. El-Banna, M. Erosion and Accretion Rates and their Associated Sediment Characters along Ras El Bar Coast, Northeast Nile Delta, Egypt. Environ. Geol. 2007, 52, 41–49. [Google Scholar] [CrossRef]
  47. Fanos, A.M.; Khafagy. A.A.; Dean, R.G. Protective Works on the Nile Delta Coast. J. Coast. Res. 1995, 11, 516–528. [Google Scholar]
  48. Elnabwy, M.T.; Emad Elbeltagi, E.; El Banna, M.M.; Elshikh, M.M.Y.; Motawa, I.; Kaloop, M. An Approach Based on Landsat Images for Shoreline Monitoring to Support Integrated Coastal Management—A Case Study, Ezbet Elborg, Nile Delta, Egypt. ISPRS Int. J. Geo-Inf. 2020, 9, 199. [Google Scholar] [CrossRef] [Green Version]
  49. Ghoneim, E.; Mashaly, J.; Gamble, D.; Halls, J.; Abu Bakr, M. Nile Delta exhibited a spatial reversal in the rates of shoreline retreat on the Rosetta promontory comparing pre- and post-beach protection. Geomorphology 2015, 228, 1–14. [Google Scholar] [CrossRef]
  50. White, K.; El-Asmar, H.M. Monitoring Changing Position of Coastlines using Thematic Mapper Imagery, an example from the Nile Delta. Geomorphology 1999, 29, 93–105. [Google Scholar] [CrossRef]
  51. Masria, A.; Negm, A.; Iskender, M.; Savedra, O.C. Numircal Modeling of Rosetta River Mouth, Egypt. Int. Water Technol. J. IWTJ 2014, 4, 190. [Google Scholar]
  52. Deabes, E.A.M. Applying ArcGIS to Estimate the Rates of Shoreline and Back-Shore Area Changes along the Nile Delta Coast, Egypt. Int. J. Geosci. 2017, 8, 332–348. [Google Scholar] [CrossRef] [Green Version]
  53. El-Gamal, A.; Balbaa, S.H.; Rashed, M.A.; Mansour, A.S. Three Decades Monitoring of Shoreline Change Pattern of Damietta Promontory, Nile Delta, Egypt. Aquat. Sci. Technol. 2020, 8, 1–14. [Google Scholar] [CrossRef]
  54. Khalifa, A.M.; Soliman, M.R.; Yassin, A.A. Assessment of a combination between hard structures and sand nourishment eastern of Damietta harbor using numerical modeling. Alex. Eng. J. 2017, 56, 545–555. [Google Scholar] [CrossRef]
  55. Masria, A.; Nadaoka, K.; Kuriyama, Y.; Negm, A.; Iskender, M.; Savedra, O.C. Neashore and Beach nourishments Effects on the Stability of the Rosette Promontory, Egypt. In Proceedings of the Eighteenth International Water Technology Conference, IWTC18 2015, Sharm El-Sheikh, Egypt, 12–14 March 2015. [Google Scholar]
  56. Masria, A.; Iskender, M.; Negm, A.; Savedra, O.C. The Effect of potential Discharges on the Stability of the Rosetta Promontory, Egypt. J. Geogr. Tech. 2014, 2, 6976. [Google Scholar]
  57. Negm, A.; Masria, A.; Iskender, M.; Savedra, O.C. Towards a Sustainable Stability of Coastal Zone at Rosetta Promontory/Mouth, Egypt. Oceanography 2015, 3, 132. [Google Scholar] [CrossRef]
  58. Pranzini, E.; Anfuso, G.; Cinelli, I.; Marco Piccardi, M.; Vitale, G. Shore Protection Structures Increase and Evolution on the Northern Tuscany Coast (Italy): Influence of Tourism Industry. Water 2018, 10, 1647. [Google Scholar] [CrossRef] [Green Version]
  59. Saı¨di, H.; Souissi, R.; Zargouni, F. Environmental Impact of Detached Breakwaters on the Mediterranean Coastline of Soliman (NE of Tunisia) end. Fis. Acc. Lincei 2012, 23, 339–347. [Google Scholar] [CrossRef]
  60. Zidan, A.R.; Rageh, O.S.; Sarhan, T.E.; Esmail, M. Effect of Breakwaters on Wave Energy Dissipation: Case Study, Ras El-Bar beach, Egypt. Int. Water Technol. J. 2012, 2, 268–283. [Google Scholar]
  61. Saengsupavanich, C. Detached Breakwaters: Communities’ Preferences for Sustainable Coastal Protection. J. Environ. Manag. 2013, 115, 106–113. [Google Scholar] [CrossRef] [PubMed]
  62. Bitan, M.; Zviely, D. Sand Beach nourishment: Experience from the Mediterranean Coast of Israel. J. Mar. Sci. Eng. 2020, 8, 273. [Google Scholar] [CrossRef] [Green Version]
  63. Brand, E.; Ramaeker, G.; Lodder, Q. Dutch Experience with Sand Nourishments for Dynamic Coastline Conservation- An Operational Overview. Ocean Coast. Manag. 2022, 217, 106008. [Google Scholar] [CrossRef]
  64. Enríquez, A.R.; Marcos, M.; Falqués, A.; Roelvink, D. Assessing Beach and Dune Erosion and Vulnerability under Sea Level Rise: A case Study in the Mediterranean Sea. Front. Mar. Sci. 2019, 6, 4. [Google Scholar] [CrossRef] [Green Version]
  65. Iskader, M. Stability of the Northern coast of Egypt under the effect of urbanization and climate change. Water Sci. 2021, 35, 1–10. [Google Scholar] [CrossRef]
  66. El-Banna, M. Nature and human impact on Nile Delta coastal sand dunes, Egypt. Environ. Geol. 2004, 45, 690–695. [Google Scholar] [CrossRef]
  67. El-Banna, M. Vulnerability and Fate of a Coastal Sand Dune Complex, Rosetta-Idku, northwestern Nile Delta, Egypt. Environ. Geol. 2008, 54, 1291–1299. [Google Scholar] [CrossRef]
  68. Della Bella, A.; Fantinato, E.; Scarton, F.; Buffa, G. Mediterranean Developed Coasts: What Future for the Foredune Restoration? J. Coast. Conserv. 2021, 25, 49. [Google Scholar] [CrossRef]
  69. Hasan, E.; Khan, S.I.; Yang, H. Investigation of Potential Sea Lvel Rise Impact on the Nile Delta, Egypt using Digital Elevation Models. Environ. Monit. Assess. 2015, 187, 649. [Google Scholar] [CrossRef] [PubMed]
  70. El-Quilish, M.; El-Ashquer, M.; Dawod, G.; El Fiky, G. Development of an Inundation Model for the Northern Coastal Zone of the Nile Delta Region, Egypt Using High-Resolution DEM. Arab. J. Sci. Eng. 2022. [Google Scholar] [CrossRef]
  71. Polonia, A.; Vaiani, S.C.; Lange, G.J. Did the A.D. 365 Crete Earthquake/Tsunami Trigger Synchronous Giant Turbidity Currents in the Mediterranean Sea? Geology 2016, 44, 191–194. [Google Scholar] [CrossRef]
  72. Rateb, A.; Abotalib, A.Z. Inferencing the Land Subsidence in the Nile Delta using Sentinel-1 Satellites and GPS between 2015 and 2019. Sci. Total Environ. 2020, 729, 138868. [Google Scholar] [CrossRef] [PubMed]
  73. El-Bastawesy, M.; Cherif, O.H.; Sultan, M. The Geomorphological Evidences of Subsidence in the Nile Delta: Analysis of High Resolution Topographic DEM and Multi-temporal Satellite Images. J. Afr. Earth Sci. 2017, 136, 252–261. [Google Scholar] [CrossRef]
  74. Gebremichael, E.; Sultan, M.; Becker, R.; El-Bastawesy, M.; Cherif, O.H.; Emil, M. Assessing Land Deformation and Sea Encroachment in the Nile Delta: A Radar Interferometric and Inundation Modeling Approach. J. Geophys. Res. Solid Earth 2018, 123, 3208–3224. [Google Scholar] [CrossRef]
  75. Saleh, M.; Becker, M. New estimation of Nile Delta Subsidence Rates from InSAR and GPS Analysis. Environ. Earth Sci. 2019, 78, 6. [Google Scholar] [CrossRef]
  76. Hassaan, M.A.; Abdrabo, M.A. Vulnerability of the Nile Delta Coastal Areas to Inundation by Sea Level Rise. Environ. Monit. Assess. 2013, 185, 6607–6616. [Google Scholar] [CrossRef]
Figure 1. Satellite image from 1972 of the Damietta Promontory with the western coast including Ras El-Bar resort and Damietta Harbor, with the eastern coast comprising the Ezbet El-Borg fishing city and the Manzala lagoon triangle (a,b). The black rectangle refers to the location of Figure 4c. The situation at 2015 and comparison with 1984 satellite image showing segments of accretion in red color and erosion in yellow (c) W.J. and E.J. are the western and eastern jetties of the Damietta Harbor. The yellow rectangle refers to the location of Figure 5f–h, while the red rectangle refers to the location of Figure 5a–e (c).
Figure 1. Satellite image from 1972 of the Damietta Promontory with the western coast including Ras El-Bar resort and Damietta Harbor, with the eastern coast comprising the Ezbet El-Borg fishing city and the Manzala lagoon triangle (a,b). The black rectangle refers to the location of Figure 4c. The situation at 2015 and comparison with 1984 satellite image showing segments of accretion in red color and erosion in yellow (c) W.J. and E.J. are the western and eastern jetties of the Damietta Harbor. The yellow rectangle refers to the location of Figure 5f–h, while the red rectangle refers to the location of Figure 5a–e (c).
Sustainability 14 15415 g001
Figure 2. Flow chart showing the procedures applied on the selected satellite images used in the current study.
Figure 2. Flow chart showing the procedures applied on the selected satellite images used in the current study.
Sustainability 14 15415 g002
Figure 3. Summary of the recorded data base of shoreline changes along the Damietta Promontory (references are listed in the last column) compared with the current results (first raw). The erosion is given in negative values while accretion rates expressed in positive values [1,4,5,8,16,22,38,39,40,41,42,43,44,45,46,47,48].
Figure 3. Summary of the recorded data base of shoreline changes along the Damietta Promontory (references are listed in the last column) compared with the current results (first raw). The erosion is given in negative values while accretion rates expressed in positive values [1,4,5,8,16,22,38,39,40,41,42,43,44,45,46,47,48].
Sustainability 14 15415 g003
Figure 4. Comparison of satellite images of 1984 and 2015 showing the values of eroded and accreted coast in meters. The yellow bar shows the location of the seawall at “b” (a). The seawall constructed in 2000; A–A´ is a location of cross section along the seawall (b) illustrated section at (c) [14], with basalt, dolomite and tetrapod dolos of 3 tons (c). The two jetties at the Damietta branch river mouth (d). The red circle and yellow rectangle at (b,d) refer to the eastern and western jetties at the Damietta branch, respectively.
Figure 4. Comparison of satellite images of 1984 and 2015 showing the values of eroded and accreted coast in meters. The yellow bar shows the location of the seawall at “b” (a). The seawall constructed in 2000; A–A´ is a location of cross section along the seawall (b) illustrated section at (c) [14], with basalt, dolomite and tetrapod dolos of 3 tons (c). The two jetties at the Damietta branch river mouth (d). The red circle and yellow rectangle at (b,d) refer to the eastern and western jetties at the Damietta branch, respectively.
Sustainability 14 15415 g004
Figure 5. The coast west of the promontory at Ras El-Bar showing the concrete seawall. (a) The yellow rectangle refers to the lighthouse; the modified 1200 m seawall extends westward of the Damietta Nile branch. (b) One of the three 120 m long concrete groins that were constructed to the west of the seawall at “b” and (c) later renewed. (d) The successful detached breakwaters that create shadow zones of 25% of the waves’ energy. (e) Eastward evidence of erosion including shell accumulation, cuspate beach “red arrows” (f) at the coast east of the harbor eastern jetty (see Figure 1c yellow area), with developed rip currents (yellow arrows) (g) and waves rolled mud balls along the eroded coast (pink arrows) (g), [4]. The new Y groins east of the harbor jetty (blue arrows) (h) with accretion along the groins and erosion along the spaces in between.
Figure 5. The coast west of the promontory at Ras El-Bar showing the concrete seawall. (a) The yellow rectangle refers to the lighthouse; the modified 1200 m seawall extends westward of the Damietta Nile branch. (b) One of the three 120 m long concrete groins that were constructed to the west of the seawall at “b” and (c) later renewed. (d) The successful detached breakwaters that create shadow zones of 25% of the waves’ energy. (e) Eastward evidence of erosion including shell accumulation, cuspate beach “red arrows” (f) at the coast east of the harbor eastern jetty (see Figure 1c yellow area), with developed rip currents (yellow arrows) (g) and waves rolled mud balls along the eroded coast (pink arrows) (g), [4]. The new Y groins east of the harbor jetty (blue arrows) (h) with accretion along the groins and erosion along the spaces in between.
Sustainability 14 15415 g005
Figure 6. Comparison of summer houses at Ras El-Bar during inundation in 1986 where water sea water attacks the houses and they lose the swimming beaches, and later in 2000 after the construction of the detached breakwaters the same house have regained a wide sandy beach (ad). Partial collapse “red arrows” and sedimentation “yellow arrow” at the protection structures, such collapse lowered the elevation of the breakwaters to less than 1 m a.s.l. (e,f).
Figure 6. Comparison of summer houses at Ras El-Bar during inundation in 1986 where water sea water attacks the houses and they lose the swimming beaches, and later in 2000 after the construction of the detached breakwaters the same house have regained a wide sandy beach (ad). Partial collapse “red arrows” and sedimentation “yellow arrow” at the protection structures, such collapse lowered the elevation of the breakwaters to less than 1 m a.s.l. (e,f).
Sustainability 14 15415 g006
Figure 7. GeoEye (Ikonos) satellite images captured 30 July 2014 and 10 August 2022 showing the changes in three coastal zones A, B, and C (a) the erosion and accretion around the Y-groins (a) and at the detached breakwaters Area B (b). The changes after the construction of the seawall “the yellow bar location” and erosion to the down drift “pink shadow” and at the sand spit (c).
Figure 7. GeoEye (Ikonos) satellite images captured 30 July 2014 and 10 August 2022 showing the changes in three coastal zones A, B, and C (a) the erosion and accretion around the Y-groins (a) and at the detached breakwaters Area B (b). The changes after the construction of the seawall “the yellow bar location” and erosion to the down drift “pink shadow” and at the sand spit (c).
Sustainability 14 15415 g007
Figure 8. Two photos showing the unsuccessful detached breakwaters at Pot Said (a), and at Baltim (b) due to short length of the shadow area. Field photos showing unsuccessful sand nourishment along El-Gamiel tourist village at Port Said the waves attacked the coast and obliterated all sands due to the non-efficient quantities of sands m3/m and the grain size (c). Finally, the coastal dunes natural defense measure subjected to erosion (d) at Baltim, threats the International coastal road and the Burullus coast.
Figure 8. Two photos showing the unsuccessful detached breakwaters at Pot Said (a), and at Baltim (b) due to short length of the shadow area. Field photos showing unsuccessful sand nourishment along El-Gamiel tourist village at Port Said the waves attacked the coast and obliterated all sands due to the non-efficient quantities of sands m3/m and the grain size (c). Finally, the coastal dunes natural defense measure subjected to erosion (d) at Baltim, threats the International coastal road and the Burullus coast.
Sustainability 14 15415 g008
Table 1. Satellite data acquired and utilized in the present study in a chronological order from 1972 to 2022.
Table 1. Satellite data acquired and utilized in the present study in a chronological order from 1972 to 2022.
SatelliteSensorsResolution (m)DateSpectral Bands
Landsat 4MSS6020 October 19720.5–0.6 μm visible green
0.6–0.7 μm visible red
0.7–0.8 μm near infrared
0.8–1.1 μm infrared
Landsat 5TM3015 October 1984
29 September 1987
0.45–0.52 μm visible blue
0.52–0.60 μm visible green
0.63–0.69 μm visible red
0.76–0.90 μm near infrared
1.55–1.75 μm infrared
10.4–12.5 μm thermal
Landsat 8OLI3030 June 20150.43–0.45 μm coastal
0.45–0.51 μm visible blue
0.53–0.59 μm visible green
0.63–0.67 μm visible red
0.85–0.87 μm near infrared
1.56–1.65 μm infrared
10.6–11.1 μm thermal
11.5–12.5 μm thermal
2.10–2.294 μm infrared
0.503–0.676 μm panchromatic
1.363–1.384μm cirrus
IKONOSIKONOS1–430 July 2014
10 August 2022
450–520 nm Blue
510–600 nm Green
630–700 nm Red
760–850 nm Near-IR
530–930 nm Panchromatic
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

El-Asmar, H.M.; Taha, M.M.N. Monitoring Coastal Changes and Assessing Protection Structures at the Damietta Promontory, Nile Delta, Egypt, to Secure Sustainability in the Context of Climate Changes. Sustainability 2022, 14, 15415. https://doi.org/10.3390/su142215415

AMA Style

El-Asmar HM, Taha MMN. Monitoring Coastal Changes and Assessing Protection Structures at the Damietta Promontory, Nile Delta, Egypt, to Secure Sustainability in the Context of Climate Changes. Sustainability. 2022; 14(22):15415. https://doi.org/10.3390/su142215415

Chicago/Turabian Style

El-Asmar, Hesham M., and Maysa M. N. Taha. 2022. "Monitoring Coastal Changes and Assessing Protection Structures at the Damietta Promontory, Nile Delta, Egypt, to Secure Sustainability in the Context of Climate Changes" Sustainability 14, no. 22: 15415. https://doi.org/10.3390/su142215415

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