You are on page 1of 28

water

Article
Dynamics of Sediment Transport and
Erosion-Deposition Patterns in the Locality of a
Detached Low-Crested Breakwater on a
Cohesive Coast
Arniza Fitri 1, *, Roslan Hashim 2, *, Soroush Abolfathi 3 and Khairul Nizam Abdul Maulud 4,5
1 School of Civil and Architectural Engineering, Nanchang Institute of Technology, Nanchang 330099, China
2 Department of Civil Engineering, Faculty of Engineering, Malaya University, Kuala Lumpur 50603, Malaysia
3 School of Engineering, University of Warwick, Coventry CV4 7AL, UK
4 Earth Observation Centre, Institute for Climate Change, Universiti Kebangsaan Malaysia,
Bangi 43600, Selangor, Malaysia
5 Smart and Sustainable Township Research Centre, Faculty of Engineering and Built Environment,
Universiti Kebangsaan Malaysia, Bangi 43600, Selangor, Malaysia
* Correspondence: arnizafitri@gmail.com (A.F.); roslan@um.edu.my (R.H.)

Received: 8 June 2019; Accepted: 13 August 2019; Published: 19 August 2019 

Abstract: Understanding the dynamics of sediment transport and erosion-deposition patterns in the
locality of a coastal structure is vital to evaluating the performance of coastal structures and predicting
the changes in coastal dynamics caused by a specific structure. The nearshore hydro-morphodynamic
responses to coastal structures vary widely, as these responses are complex functions with numerous
parameters, including structural design, sediment and wave dynamics, angle of approach, slope of
the coast and the materials making up the beach and structures. This study investigated the sediment
transport and erosion-deposition patterns in the locality of a detached low-crested breakwater
protecting the cohesive shore of Carey Island, Malaysia. The data used for this study were collected
from field measurements and secondary sources from 2014 to 2015. Sea-bed elevations were monitored
every two months starting from December 2014 to October 2015, in order to quantify the sea-bed
changes and investigate the erosion-deposition patterns of the cohesive sediment due to the existence
of the breakwater. In addition, numerical modelling was also performed to understand the impacts
of the breakwater on the nearshore hydrodynamics and investigate the dynamics of fine sediment
transport around the breakwater structure. A coupled two-dimensional hydrodynamics-sediment
transport model based on Reynolds averaged Navier-Stokes (RANS) equations and cell-centered
finite volume method with flexible meshing approach was adopted for this study. Analysis of the
results showed that the detached breakwater reduced both current speed and wave height behind the
structure by an average of 0.12 m/s and 0.1 m, respectively. Also, the breakwater made it possible for
trapped suspended sediment to settle in a sheltered area by approximately 8 cm in height near to
the first main segment of the breakwater, from 1 year after its construction. The numerical results
were in line with the field measurements, where sediment accumulations were concentrated in the
landward area behind the breakwater. In particular, sediment accumulations were concentrated along
the main segments of the breakwater structure during the Northeast (NE) season, while concentration
near the first main segment of the breakwater were recorded during the Southwest (SW) season.
The assessment illustrated that the depositional patterns were influenced strongly by the variations
in seasonal hydrodynamic conditions, sediment type, sediment supply and the structural design.
Detached breakwaters are rarely considered for cohesive shores; hence, this study provides new,
significant benefits for engineers, scientists and coastal management authorities with regard to
seasonal dynamic changes affected by a detached breakwater and its performance on a cohesive coast.

Water 2019, 11, 1721; doi:10.3390/w11081721 www.mdpi.com/journal/water


Water 2019, 11, 1721 2 of 28

Keywords: coastal dynamics; coastal resilience; cohesive sediment; detached low-crested breakwater;
erosion-deposition pattern; artifitial reef

1. Introduction
Detached breakwaters are parallel barrier structures placed in the shallow nearshore water column,
to protect any landform area behind them from the direct impacts of wave attack, currents, tides and
storms surges [1,2]. Detached shore-parallel breakwaters are also known as artificial reefs. Low-crest
breakwaters have small relative freeboard and are commonly built of quarry materials of homogeneous
size [3]. Detached breakwaters are being used increasingly worldwide in the intertidal zones of
non-cohesive (sand) coasts, to provide protection measures and mitigate erosion problems [4–7].
However, such coastal structures cause local changes to nearshore flow hydrodynamics and sediment
dynamics in the coastal zone [7–9]; these are site-specific and affected by parameters such as sediment
characteristics, climate conditions, structural configurations (design, shape, dimension, and material)
and nearshore hydro-morphodynamics [2,5,10]. The complexity of morphodynamic changes due to
the presence of coastal structures is higher when coastal structures are located on intertidal areas of
cohesive shores, where significant temporal and spatial variation of the water depths exist [11].
Sediment transport in coastal areas is mainly influenced by dynamic nearshore processes and
site-specific environmental conditions, including sediment characteristics, wind, currents, waves,
tides and the exchange processes between estuaries and nearshore regions [12–16]. Coastal protection
structures, such as a breakwater, can change flow patterns, hydrodynamics and sediment transport,
which could then impact erosion-deposition patterns in the coastal zone [17–19]. Breakwaters can
reduce the incident wave energy and impact on the sediment transport capacity, allowing sediment
deposition on the shoreward sides of the structure [20,21]. An accurate description of the sediment
transport around a protection structure is vital for predicting the effects of coastal structures on
morphological changes in coastal zone [22].
Previous studies performed on shoreline changes due to the presence of detached breakwaters have
mainly focused on non-cohesive (sand) coasts [23–28], while detached breakwaters are infrequently
used at cohesive coasts due to the lack of understanding of complicated cohesive sediment dynamics
impacts on nearshore morphodynamics [21,27–29].
The sloping bottom of the nearshore acts like a natural defence and protects the shore against
waves, currents and storms [30,31] by dissipating the wave energy through wave breaking phenomena
and depending on bottom friction coefficient [32]. The breaking wave may re-form and break again
several times before finally rushing up the foreshore areas in the swash zone. However, in the presence
of a breakwater in the nearshore area, the structure will interact with incident waves and dissipate some
of the wave energy while the remaining turbulent enegy will be reflected and transmitted through the
breakwater [33,34]. For submerged breakwater, the waves may simply transmit over the breakwater’s
structure. However, when the crest of the breakwater is above the mean water level, waves generate
a secondary flow at the toe of the structure or transmit through breakwater when the structure is
permeable [35].
Previous research has reported that cohesive sediments have low settling velocities and therefore
could be transported over a long distance before settling through water column [36–38]. However,
the cohesive behaviour of fine sediments allow them to agglomerate and form bigger- sized aggregates
called flocks, with higher settling velocities compared to single-particle fine sediments. Thus, flocks can
settle in areas where fine sediments can never be deposited [39]. Generally, the formation of flocks can
occur when suspended sediments in the water column exceeds 0.01 kg/m3 [39].
On a non-cohesive shore, presence of breakwaters lead to sand deposition on the lee side of the
structure and formation of a sandbar (tombolo) [35,40]. The tombolo grows from the shore towards
the structure. Prior to the sand deposition reaching the structure, the sandform is referred to as
Water 2019, 11, 1721 3 of 28

salient [35,40]. If a breakwater is built perpendicular to the shoreline, the longshore sediment transport
is cut at the breakwater; thus, significant morphological changes occur near the breakwater [22].
For a cohesive shore, the hydro-morphodynamic behaviour of sediments in presence of a detached
breakwater is not yet fully understood and more research is needed. This paper provides valuable new
information on the dynamic of cohesive sediment in presence of a detached low-crested breakwater,
and improves our understanding of the complex hydro-morphodynamics of cohesive coasts which
enables coastal planners, managers and engineers, to conduct accurate environmental assessment and
design suitable coastal defence structures for cohesive shores.
In the last few decades, the cohesive shore of Carey Island, Malaysia has been experiencing
mangrove degradation and erosion problems due to human interventions in the coastal zone.
To reduce erosion problems, a rehabilitation project has been carried out by constructing a detached
low-crest breakwater near the degraded mangrove area. The impact of the detached breakwater on
sediment transport dynamics and erosion-deposition patterns on the cohesive shore had not been
fully investigated to this date. Hence, this study aims to understand the sediment transport and
erosion-deposition patterns in the locality of a detached low-crested breakwater on the cohesive shore
of Carey Island, Malaysia under seasonal variation of hydrodynamic conditions. This study specifically
investigates: (i) the characteristics of, and changes in, the hydrodynamics of the cohesive shore of
Carey Island due to the existence of the detached breakwater during the NE and SW seasons, (ii) the
pattern of suspended sediment concentration (SSC) around the breakwater structure during the NE
and SW seasons, and (iii) the variation of sea-bed elevation and the erosion-deposition pattern in the
vicinity of the breakwater structure on the cohesive shore of Carey Island.

2. Study Area
The study site is located in degraded environment in an intertidal area of Carey Island on the
west coast of peninsular Malaysia, within the Strait of Malacca (Figure 1). The Carey Island coast
is a mangrove forest reserve, with a semi-diurnal tidal system that receives daily tidal inundations
of maximum 2.96 m MSL of neap rise and 4.33 m MSL of spring rise. For the most part, the Carey
Island coast is covered by cohesive sediment. Protection of the landward against tidal inundation and
prevention of salt water intrusion during high tides are accomplished by Earth dyke constructed by
the Department of Irrigation and Drainage (DID) of Malaysia, along the coastline of the island.
According to the Malaysian Department of Meteorology data, the temperatures at the Carey
Island have fluctuated between Tmax = 35 ◦ C and Tmin = 26 ◦ C, with an average humidity of 92%,
since the year 2000 [40]. Based on the rainfall data collected by the Sime Darby Plantation Berhad in
2015, the total annual rainfall at the study site was 1718 mm and the minimum and maximum monthly
rainfalls were 63 mm and 281 mm, respectively (2015). The Langat River is situated near the study site
(degraded mangrove area), and the estuary of the river is located approximately 8 km from the study
site (Figure 1a). The suspended sediments are carried by Langat River to the Malacca Strait as well as
the area of the study site.
Water 2019, 11, 1721 4 of 28
Water 2019, 11, x FOR PEER REVIEW 4 of 28

(a)

Earth dyke

5m Earth dyke

CG1
S1
Earth dyke
Circulation Gap CG2 Main body
S2

CG3 S3

(without L-block)

(b)

Figure 1. View of the study area: (a) location of Carey Island, Malaysia; (b) study site.
Figure 1. View of the study area: (a) location of Carey Island, Malaysia; (b) study site.
The problems of erosion and mangrove degradation have plagued the intertidal area of Carey
The problems of erosion and mangrove degradation have plagued the intertidal area of Carey
Island since 1995. To tackle this problem, a mangrove rehabilitation project was carried in the
Island since 1995. To tackle this problem, a mangrove rehabilitation project was carried in the
intertidal area in early 2009 which included constructing an 85 m-long stretch of low-crested, detached
intertidal area in early 2009 which included constructing an 85 m-long stretch of low-crested,
breakwaters consisting of three main segments (main body; S1, S2, S3) and three circulation gaps
detached breakwaters consisting of three main segments (main body; S1, S2, S3) and three circulation
gaps (CG1, CG2, CG3) (Figure 1b) [22]. The distance between the dyke structure and the northern
Water 2019, 11, 1721 5 of 28

Water 2019, 11, x FOR PEER REVIEW 5 of 28


(CG1, CG2, CG3) (Figure 1b) [22]. The distance between the dyke structure and the northern section
of the breakwater is approximiately 5 m. The lengths of the three main segments of breakwater
section of the breakwater is approximiately 5 m. The lengths of the three main segments of
structure (S1, S2, S3) are 20, 30 and 20 m, respectively, with a height of 1.4 m and a width of 2.5 m.
breakwater structure (S1, S2, S3) are 20, 30 and 20 m, respectively, with a height of 1.4 m and a width
The circulation gaps (CG1, CG2, CG3) are each 5 m long, with a height of 0.8 m, and a width of
of 2.5 m. The circulation gaps (CG1, CG2, CG3) are each 5 m long, with a height of 0.8 m, and a width
2.0 m (Figure 2) [22]. The breakwater was constructed at longitude 101◦ 200 23.5” to 101◦ 200 24.5” E,
of 2.0 m (Figure◦ 2)0[22]. The ◦breakwater was constructed at longitude 101°20′23.5” to 101°20′24.5” E,
and latitude 02 49 27” to 02 490 28.5” N, with the aim reducing current speeds and wave actions and
and latitude 02°49′27” to 02°49′28.5” N, with the aim reducing current speeds and wave actions and
increasing sediment accumulation and sea-bed elevation in the protected mangrove degradation areas
increasing sediment accumulation and sea-bed elevation in the protected mangrove degradation
behind the breakwater in order to create a suitable tidal regime for restoring the lost mangroves.
areas behind the breakwater in order to create a suitable tidal regime for restoring the lost mangroves.
The detached breakwater is a rubble mound armoured by L-blocks (Figure 2). In addition, there are
The detached breakwater is a rubble mound armoured by L-blocks (Figure 2). In addition, there are
stones (without L-blocks) with an average diameter of 15 to 20 cm fitted neatly to the right side of the
stones (without L-blocks) with an average diameter of 15 to 20 cm fitted neatly to the right side of the
breakwater in mounds approximately 0.8 m high, 2.5 m wide and 20 m long. The stones are arranged
breakwater in mounds approximately 0.8 m high, 2.5 m wide and 20 m long. The stones are arranged
to trap more sediment within the study site [22]. The breakwater is submerged during the spring tide
to trap more sediment within the study site [22]. The breakwater is submerged during the spring tide
and emerges during the neap tide.
and emerges during the neap tide.

Figure 2. Schematic of detached breakwater structure.


Figure 2. Schematic of detached breakwater structure.
3. Methods
3. Methods
3.1. Data Collection
3.1. Data Collection
In this study, field-measured and secondary data were collected and analysed to understand the
hydrodynamic and sediment transport patterns. Field-measured data include sediment samplings
In this study, field-measured and secondary data were collected and analysed to understand the
along the coastline of the Carey Island; fine resolution bathymetry data survey along Langat River and
hydrodynamic and sediment transport patterns. Field-measured data include sediment samplings
Carey Island coastline; measurement of currents, waves, water level, suspended sediment concentration
along the coastline of the Carey Island; fine resolution bathymetry data survey along Langat River
(SSC) characteristics near the study site area; water samplings at the mouth of Langat River and
and Carey Island coastline; measurement of currents, waves, water level, suspended sediment
concentration (SSC) characteristics near the study site area; water samplings at the mouth of Langat
River and near the study site; and monitoring of coastal sea-bed profiles in the vicinity of the
breakwaters. The secondary data included the weather data consist of wind and wave data (2014–
Water 2019, 11, 1721 6 of 28
Water 2019, 11, x FOR PEER REVIEW 6 of 28

near therainfall
2015); study site;
dataand monitoring
(2014–2015); tideofatcoastal
Lumutsea-bed
station, profiles
Belawaninstation,
the vicinity
Tanjung of Keling
the breakwaters.
station and
The
Dumai station (2014–2015); and bathymetry data for offshore zones (approximately 45 × 160rainfall
secondary data included the weather data consist of wind and wave data (2014–2015); km).
data (2014–2015); tide at Lumut station, Belawan station, Tanjung Keling station
The typical weather on Carey Island, Malaysia is primarily influenced by the Northeast season and Dumai station
(2014–2015);
(NE) from and bathymetry
November data for
to March andoffshore zones (approximately
the Southwest season (SW) from 45 ×May160 km).
to September. April and
The typical weather on Carey Island, Malaysia is primarily
October represent transition periods between the two seasons. The weather influenced by the Northeast season (NE)
data, consisting of the
from November to March and the Southwest season (SW) from May to September.
daily wind and wave characteristics were collected at lat. 2°–3° N, long. 101°–102° E by the Malaysian April and October
represent
Departmenttransition periods between
of Meteorology. The the two seasons.
3-hour wave climateThe weather
conditionsdata,were
consisting
also of the daily
obtained wind
from the
and wave characteristics were collected at lat. 2◦ –3 ◦ N, long. 101 ◦ –102 ◦ E by the Malaysian Department
European Centre for Weather Forecasts of Medium Range (ECMWF), in grid format for longitude
of99.75°
Meteorology.
E–101.45°TheE 3-hour wave 2°
and latitude climate
N–3.25°conditions
N. Hourly were also obtained
rainfall data for from the European
the Klang region wasCentre for
obtained
Weather Forecasts of Medium Range (ECMWF), in grid format for longitude 99.75 ◦ E–101.45◦ E and
from the West Estate Office, Sime Darby Plantation Berhad, Carey Island, and hourly tidal data was
latitude
obtained2◦ N–3.25 ◦
from the N. Hourly rainfall
Malaysian dataMapping
Survey and for the Klang region was
Department obtained
(JUPEM). from
Figure the Westthe
3 presents Estate
wind
Office, Sime Darby Plantation Berhad, Carey Island, and hourly tidal
roses during the NE and SW seasons based on the analysis of weather data in 2015. The figuredata was obtained from the
Malaysian
illustrates the dominant wind speed and direction during both NE and SW seasons. The weatherNE
Survey and Mapping Department (JUPEM). Figure 3 presents the wind roses during the data
and SW seasons based on the analysis of weather data in 2015. The figure
were used for setting up the hydrodynamic conditions in the numerical modelling study (Section illustrates the dominant
wind
3.3).speed and direction during both NE and SW seasons. The weather data were used for setting up
the hydrodynamic conditions in the numerical modelling study (Section 3.3).

(b)

(b)
(a)

Figure 3. Wind roses in 2015: (a) NE season; (b) SW season (sources of data: Malaysian Department
Figure 3. Wind roses in 2015: (a) NE season; (b) SW season (sources of data: Malaysian Department
of Meteorology).
of Meteorology).
During the field measurement campaign, twenty sediment samples were collected along the Carey
During the field measurement campaign, twenty sediment samples were collected along the
Island coastline and the area surrounding the breakwater, at depths of 0–100 cm to determine the size
Carey Island coastline and the area surrounding the breakwater, at depths of 0–100 cm to determine
distribution of the sediment particles.
the size distribution of the sediment particles.
Water samples and flow velocity were collected from the mouth of the Langat River (during the
Water samples and flow velocity were collected from the mouth of the Langat River (during the
period of 23 December 2014 to 7 January 2015) to determine the total suspended solids (TSS1) values
period of 23 December 2014 to 7 January 2015) to determine the total suspended solids (TSS1) values
and the discharge values of the river. For each TSS test, 100 mL water of each sample was filtered
and the discharge values of the river. For each TSS test, 100 mL water of each sample was filtered
through a pre-weighed glass fibre filter. The filters were then dried at 105 ◦ C in an oven and weighed.
through a pre-weighed glass fibre filter. The filters were then dried at 105 °C in an oven and weighed.
The mass differences between the pre and post-filtration of the filter were measured to determine the
The mass differences between the pre and post-filtration of the filter were measured to determine the
total suspended sediment in 100 mL of water. The TSS analyses were performed in accordance with the
total suspended sediment in 100 mL of water. The TSS analyses were performed in accordance with
APHA 2540D standard method. Furthermore, in order to obtain information about the average amount
the APHA 2540D standard method. Furthermore, in order to obtain information about the average
of TSS in the Strait of Malacca (TSS 2), which can affect the sediment transport around the study site,
amount of TSS in the Strait of Malacca (TSS 2), which can affect the sediment transport around the
ten water samples were collected monthly from a site located at latitude 2◦ 49’38.38” N and longitude
study site, ten water samples were collected monthly from a site located at latitude 2°49’38.38” N and
Water 2019, 11, 1721 7 of 28

101◦ 19’55.39” E, which is approximately 500 m seaward of the breakwater. Table 1 summarises the
monthly climate conditions at the study site in 2015.

Table 1. Monthly climate conditions in 2015.

Month
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
No. of rainfall
4 5 4 5 9 4 4 12 7 9 11 14
events
Rainfall
63 64 118 60 221 128 65 281 254 139 192 133
(mm/month)
Significant wave
1.2 0.75 1 1 0.75 0.75 0.5 0.75 1 1.5 1.2 1
height (m)
Mean wave
5 4 5 4 3 3 3 4 5 5 5 4
period (s)
Dominant wind
9.1 7.5 8.5 7.3 6.2 5.3 5.5 6.5 5.0 6.5 7.5 8.5
speed (m/s)
Dominant wind
300 330 320 270 120 130 140 130 120 290 310 300
direction (degree)
Sources of data: Rainfall data were obtained from Sime Darby Plantation Berhad, Carey Island; Wind and wave
data were obtained from Malaysian Meteorology Department and ECMWF.

Based on weather data obtained from Malaysian Meteorology Department (Figure 3), the wind
during the NE season predominantly blew from the Northwest (approximately 300◦ –330◦ ) at a
dominant speed range of 7.5–10 m/s, while the wind during the SW season predominantly blew from a
Southeastern direction (approximately 120◦ –150◦ ) at a dominant speed range of 5–7.5 m/s. Monthly
climate monitoring (Table 1), indicates the significant wave heights and intensity of wind speed were
higher during the Northeast season compared to the Southwest season.
In addition to the climatic data mentioned above, two Acoustic Wave and Current Profiler (AWAC)
units and four optical backscatter sensors (OBS-3A) were installed at the study site (see Figure 4) to
measure the suspended sediment concentration (SSC), water level fluctuations, currents and waves
characteristics. Field measurements using AWAC and OBS-3A were carried out from 23 December
2014 to 7 January 2015, covering both the neap and spring tides. The data collected during the field
study campaign, were used for the model setup, calibration and validation of the numerical model
(Section 3.3). Table 2 shows the coordinates and water depth for the AWACs, OBS-3As, water samples
and water discharge measurements.

Table 2. Coordinates of field measurements.

Station Longitude (x) Latitude (y) Water Depth (m)


Station 1 (AWAC 1 and OBS 1) 101◦ 200 11.18” E 02◦ 480 40.02” N 10.324
Station 2 (AWAC 2 and OBS 2) 101◦ 180 58.14” E 02◦ 490 26” N 12.557
Station 3 (OBS 3) 101◦ 260 10.06” E 02◦ 400 7.910 N 15.221
Station 4 (OBS 4) 101◦ 060 44.34” E 03◦ 80 36.81” N 10.483
Water samples (TSS 1) and water
101◦ 240 6.24” E 2◦ 480 2.72” N 6.242
discharges from LR
Water 2019,
Water 11, 1721
2019, 11, x FOR PEER REVIEW 8 of
8 28
of 28

3.25°

3.00°

(a)
2.50°

2.00°

100.00° E 101° E 102° E

W1
(b)

Figure 4. (a) Locations of AWAC and OBS-3A units within the study site, (b) location of water and bed
samples
Figurein4.the
(a)study site (Sof=AWAC
Locations soil sample, W = water
and OBS-3A sample
units and
within thedischarge).
study site, (b) location of water and
bed samples in the study site (S = soil sample, W = water sample and discharge).
A fine resolution bathymetry survey was conducted around the coastline of Carey Island and the
Langat River,
A finecovering
resolution area of 17.5 ×
anbathymetry 5 km with
survey lines spacedaround
was conducted at 20–500
themcoastline
intervals.
of The survey
Carey was
Island and
carried out using a boat equipped with an echo-sounder and a DGPS during the spring tide on 8
the Langat River, covering an area of 17.5 × 5 km with lines spaced at 20–500 m intervals. The survey to 12
December 2014.
was carried Bathymetry
out data
using a boat were processed
equipped and prepared and
with an echo-sounder for setting
a DGPSup the numerical
during the springmodel
tide on
using flexible meshing technique.
8 to 12 December 2014. Bathymetry data were processed and prepared for setting up the numerical
model using flexible meshing technique.

3.2. Monitoring of Sea-Bed Elevation


Water 2019, 11, 1721 9 of 28

Water 2019, 11, x FOR PEER REVIEW 9 of 28


3.2. Monitoring of Sea-Bed Elevation
The sea-bed elevations in the mangrove degraded area of the Carey Island intertidal region were
The sea-bed
monitored using elevations in the(theodolite)
a Total Station mangrove degraded area of the
and bed-profiler forCarey Island intertidal
a two-month period, region
duringwerethe
monitored using a Total Station (theodolite) and bed-profiler for a
period from December 2014 to October 2015. The sea-bed elevation monitoring was two-month period, during themainly
period
from December
concentrated 2014
at the to Octoberregion,
breakwater 2015. especially
The sea-bed theelevation
mangrove monitoring
degradedwas areamainly concentrated
at the landward of theat
the breakwater region, especially the mangrove degraded area at the landward
breakwater, as more significant sea-bed level changes was expected in that area. A total of twenty- of the breakwater, as
moreprofile
four significant sea-bed
lines (CS1 level changes
to CS24), was expected
perpendicular in that area.
to the shoreline, wereA surveyed
total of twenty-four
and monitoredprofile lines
during
(CS1
the to CS24),
study period perpendicular to the shoreline,
(Figure 5). Monitoring sea-bedwere surveyed
levels and monitored
on a cohesive shore is aduring thetask
difficult study period
since any
(Figure 5). Monitoring sea-bed levels on a cohesive shore is a difficult task since
movement on the cohesive sea-bed could disturb the sediment surfaces significantly. In this study, any movement on the
cohesivepoles
wooden sea-bedwerecould disturb
pushed intothe
thesediment surfaces significantly.
cohesive sediments around the In this study,area
monitoring wooden
at 5 mpoles were
distances
pushed
along intoprofile
each the cohesive
line tosediments around the
obtain accurate bedmonitoring areaThe
surface data. at 5 mbed distances
profile along each profile
measurements line
were
to obtain accurate bed surface data. The bed profile measurements were conducted
conducted during low tide exposure, according to the datum provided by JUPEM near the study site. during low tide
exposure,
The according
profiling outcomes to the
weredatum
used provided
to analysebythe
JUPEM near the studypatterns
erosion-deposition site. The
inprofiling outcomes
the locality of the
were used to analyse the erosion-deposition patterns in the locality of
breakwater’s structure by determining the elevation differences between the bed profiling the breakwater’s structure by
determining the elevation differences between the bed profiling
measurements obtained and the initial sea-bed elevations in December 2014. measurements obtained and the initial
sea-bed elevations in December 2014.

Figure 5. Bed profiling method in the study site (24 profiles).


Figure 5. Bed profiling method in the study site (24 profiles).
3.3. Numerical Modelling
3.3. Numerical
Accurate Modelling
prediction of hydrodynamic and sediment transport characteristics under the influence
of wave-current interactions
Accurate prediction of in coastal regionsand
hydrodynamic is a sediment
challenging task duecharacteristics
transport to multifacetedunder
nearshore
the
dynamic processes which are varying on both temporal and spatial scales [19,41,42].
influence of wave-current interactions in coastal regions is a challenging task due to multifaceted Previous studies
have shown
nearshore that modelling
dynamic processes nearshore problems
which are varying based on the
on both shallow and
temporal waterspatial
equation provides
scales high
[19,41,42].
accuracy and efficiency [11,43,44]. In this study, the MIKE 21 Flow Model FM
Previous studies have shown that modelling nearshore problems based on the shallow water including Hydrodynamic,
Spectra Wave
equation and high
provides Mudaccuracy
Transportand modules was[11,43,44].
efficiency adopted for simulating
In this theMIKE
study, the flow hydrodynamics
21 Flow Model FM and
cohesive sediment transport for the case study described in Section 2. MIKE
including Hydrodynamic, Spectra Wave and Mud Transport modules was adopted for simulating 21 Flow model FM is a
complete coastal modelling suite, which is capable of designing the data assessment
the flow hydrodynamics and cohesive sediment transport for the case study described in Section 2. for coastal and
offshore
MIKE 21 structures;
Flow model andFMenvironmental
is a completeimpact
coastalassessment
modellingofsuite,
marine infrastructures
which is capable ofbased on flexible
designing the
mesh approach. This model was established by the Danish Hydraulic Institute
data assessment for coastal and offshore structures; and environmental impact assessment of marine (DHI), Denmark.
MIKE 21 Hydrodynamic
infrastructures based on isflexible
the basicmesh
module of the MIKE
approach. This 21 Flow Model
model FM systemby
was established for the
free Danish
surface
flows which
Hydraulic simulates
Institute water
(DHI), level fluctuations
Denmark. MIKE 21 and flows in response
Hydrodynamic is thetobasic
a variety of forcing
module functions
of the MIKE 21
Flow Model FM system for free surface flows which simulates water level fluctuations and flows in
Water 2019, 11, 1721 10 of 28

in lakes, estuaries, bays and coastal areas. The details of the numerical model used for this study are
as follows:
The modelling approach adopted for this study is based on the numerical solution of the two
dimensional incompressible Reynolds averaged Navier-Stokes equations with the assumption of
Boussinesq and hydrostatic pressure.
The governing equations used for the two-dimesional numerical model in Cartesian coordinates
are continuity (Equation (1)) and momentum (Equations (2) and (3)):

∂u ∂v ∂w
+ + =S (1)
∂x ∂y ∂z

∂ρ η
∂u ∂u2 ∂vu ∂wu ∂η 1 ∂pa ∂ ∂u
Z !
g
+ + + = fv − g − − dz + Fu + νt + us S (2)
∂t ∂x ∂y ∂z ∂x ρ0 ∂x ρ0 z ∂x ∂z ∂z
Z η
∂v ∂v2 ∂vu ∂wu ∂η 1 ∂pa ∂ρ ∂ ∂v
!
g
+ + + = −fu − g − − dz + Fv + νt + us S (3)
∂t ∂y ∂x ∂z ∂y ρ0 ∂y ρ0 z ∂y ∂z ∂z
where t denotes time and u, v, w are velocity components in x, y, z Cartesian coordinates, respectively.
d denotes the still water depth and h is the total water depth (= η + d), g is the gravitational acceleration,
pa is atmospheric pressure, ρ is density, ρ0 is reference density of water, νt is the vertical diffusivity, f is
Coriolis parameter and S is point-source discharge magnitude. Fu and Fv are gradient-stress relations
described is Equations (4) and (5), respectively.

∂ ∂u ∂ ∂u ∂v
!!
Fu = (2A ) + A + (4)
∂x ∂x ∂y ∂y ∂x

∂ ∂u ∂v ∂ ∂v
!! !
Fv = A + + 2A (5)
∂x ∂y ∂x ∂y ∂y
where A is the horizontal eddy viscosity.
The total water depth, h is determined based on the kinematic boundary condition at the surface
by use of robust vertical integration of local continuity equation (Equation (6)).

∂h ∂hu ∂hv
+ + = hS + P̂ − Ê (6)
∂t ∂x ∂y

where P̂ and Ê are the precipitation and evaporation rates, respectively and u and v are the
depth-averaged velocities in the two-dimensional domain. The model does not consider compressibility
of water. A standard k − ε model is used for turbulence modelling and the vertical eddy diffusivity is
derived based on Equation (7).
k2
νt = cµ (7)
ε
where k is the turbulent kinetic energy per unit mass, cµ is an empirical constant and ε is the dissipation
rate of turbulent kinetic energy.
The discretisation of the numerical domain is performed with use of unstructured grid approach
to achieve an optimal degree of flexibility in the representation of geometry of the detached breakwater,
enabling a smooth representation of offshore and onshore boundaries. The flexible meshing approach
used for this study enabled depth-adaptive and boundary-fitted mesh and provided adequate resolution
of the bathymetry of study site, as well as high accuracy wave and current generation. MIKE 21 Spectral
Wave (SW), a spectral wind-wave model, is used to simulate the growth, decay and transformation of
wind-generated waves and swells in the numerical domain for the described case-study. The model
solves the spectral wave action balance equation formulated by Komen et al. [45] and Young [46].
The hydrodynamic module was coupled with a sediment transport module capable of simulating
fine sediments, to investigate the impact of detached breakwater on the coastal waters of Carey Island
Water 2019, 11, 1721 11 of 28

and the sediment deposition-erosion patterns landwards of the breakwater. MIKE 21 Mud Transport
module solves an advection-dispersion equation, based on Mehta [47], and the impact of waves
and currents are introduced with bed shear stress. The cohesive sediment transport is described by
Equation (8):
 
∂  νTy ∂ci 
   
∂ci ∂uci ∂vci ∂wci ∂ws ci ∂  νTx ∂ci  ∂  νTz ∂ci 
+ + + − = + +   + Si (8)
∂t ∂x ∂y ∂z ∂z ∂x  σi ∂x  ∂y  σi ∂y  ∂z  σi ∂z 
 
Tx Ty Tz

where ci is the ith component of the mass concentration, νTx is the anistropic eddy viscosity, Si is the
source term, σiTx is turbulent Schmitt number, and ws is the settling velocity. The model is capable of
considering flocculation as a function of suspended sediment concentration (Equation (9)).
c
ws = k × ( )γ (9)
ρsediment

where k is a constant, ρsediment is sediment density, and γ is the settling index coefficient. The deposition
is described in the model by Equation (10):

SD = ws cb pD (10)

where ws denotes settling velocity of suspended sediment, pD describes the probability of deposition
(= 1 − τb /τcd ) and cb is suspended sediment concentration near the bed. The sediment transport model
is soved by spatial discretisation of the primitive equations with use of cell-centred finite volume
method. An unstructured gird approach is used for the horizontal plane, while in the vertical domain
structured mesh is used.
The numerical model described above has been adopted successfully in several studies. Jose and
Stone [48] and Jose et al. [49] investigated the characteristics of wave transformation with MIKE 21
Spectra Wave model for south-central Louisiana, USA [49]. Patra et al. [44] investigated and validated
the characteristics of offshore waves in the Bay of Bengal, India using MIKE 21 Spectra Wave FM.
In addition, a study of suspended sediment transport was carried out sucessfully by Sravanthi et al. [10]
along the cohesive shore of Central Kerala, India using MIKE 21 Mud Transport FM. Previous studies
have shown that MIKE 21 is capable of simulating hydrodynamic processes with over 85% accuracy
when compared to field data [11,44,48–50].

3.3.1. Model Setup


The numerical model described in Section 3.3 is used to investigate the impact of breakwater on
the hydrodynamics and sediment transport patterns in the locality of the breakwater. In this study,
the hydrodynamic characteristics, including currents and waves, were simulated in presence of the
breakwater and without the breakwater for both seasons (northeast season and southwest season) to
quantify the impact of the breakwater on the nearshore processes. Separate models were developed
to simulate the conditions for the neap and spring tides. The suspended sediment concentrations
(SSCs) were also simulated in the vicinity of the breakwater for both seasons and tidal conditions.
The numerical results were compared to the data obtained from field measurements, for calibration
and validation purposes.

3.3.2. Model Input


Model input data consisted of bathymetry data; climate data, including wind characteristics,
water level and wave characteristics; sediment data, including sediment characteristics, SSC and TSS;
and water discharge from the Langat River. The bathymetry data for the Strait of Malacca at coarse
resolution were obtained from the Malaysian National Hydrography Centre, based on navigation
survey from 2012. These data were further integrated with bathymetry data provided by DHI in
Water 2019, 11, 1721 12 of 28

Water 2019, 11, x FOR PEER REVIEW 12 of 28


C-MAP (2014). Bathymetry data for the Carey Island coast and Langat River areas were measured at
fine
fineresolution
resolutionduring
duringthethefieldwork
fieldwork(8(8toto1212December
December2014).
2014).Following
Followingsuccessful
successfulmodel
modelcalibration
calibration
and
andvalidation,
validation,thetheNE
NEseason
seasonsimulations
simulationswere
wereset
setup
upby
byusing
usingthe
thedominant
dominantwind
wind(8.5
(8.5m/s
m/sofofspeed
speed
and 310◦ofofdirection)
and310° direction)and
andwave
wavecharacteristics
characteristics(1(1mmheight
heightand
and5 5s speriod)
period)during
duringthetheNE
NEseason
seasoninin
2015
2015and
and the average
averageSSC.
SSC.ForForthethe
SWSW season
season simulations,
simulations, the model
the model was
was set up set up bytheusing
by using the
dominant
dominant
wind (6.5wind (6.5
m/s of m/s and
speed of speed ◦
140 of and 140° of direction)
direction) and waveand wave characteristics
characteristics (0.75 and
(0.75 m height m height and
3 s period)
3 during
s period)SWduring
seasonSW season
in 2015 andinthe
2015 and the
average average
SSC. SSC. Both
Both models usedmodels used
SSC data SSC data
collected collected
from from
23 December
232014
December 2014 to
to 7 January 7 January 2015.
2015.

3.3.3.Computational
3.3.3. ComputationalDomain
Domain
Thecomputational
The computationaldomain
domainwas wasdeveloped
developedwith withthe
theuse
useofofbathymetry
bathymetrydatadataandandby byadopting
adoptinga a
flexiblemesh
flexible meshtechnique.
technique.Figure
Figure6 6depicts
depictsthe
thecomputational
computationaldomain domaindeveloped
developedfor forhydrodynamic
hydrodynamic
andsediment
and sedimenttransport
transportmodelling.
modelling.This Thisstudy
studysetsetthe
the computationaldomains
computational domainsbeyond
beyondthe the study
study area,
area,
according to the guideline
according to the guideline for coastal hydraulic studies published by the Department
coastal hydraulic studies published by the Department of Irrigation of Irrigation and
Drainage
and Drainage(DID), Malaysia
(DID), [51]. Therefore,
Malaysia the computational
[51]. Therefore, domain domain
the computational developed for the simulation
developed for the
(MIKE 21 Hydrodynamic and MIKE 21 Spectra Wave) was expanded to
simulation (MIKE 21 Hydrodynamic and MIKE 21 Spectra Wave) was expanded to include the include the Tanjung Keling and
Lumut areas
Tanjung Keling (Figure 6a). Toareas
and Lumut improve the cost
(Figure 6a). efficiency
To improve of the modelling, smaller
cost efficiency computational
of the domain
modelling, smaller
were generated
computational for the
domain sediment
were transport
generated for themodel
sediment (Figure 6b). model
transport Figure (Figure
6 depicts6b).the computational
Figure 6 depicts
domain
the developed
computational for hydrodynamic,
domain developed forwave and sediment
hydrodynamic, transport
wave modelling.
and sediment transport modelling.

(a)

Figure 6. Cont.
Water
Water 11,11,
2019,
2019, 1721 PEER REVIEW
x FOR 13 13
of of
2828

Study Area

(b)

Figure 6. Computational domain: (a) hydrodynamic and wave models; (b) mud transport model.
Figure 6. Computational domain: (a) hydrodynamic and wave models; (b) mud transport model.
3.3.4. Model Calibration and Validation
3.3.4. Model Calibration and Validation
To check the accuracy of the simulation results from the models and to provide confidence in the
To checkofthe
simulations theaccuracy of the and
current, waves simulation
patternsresults from the
of suspended modelsconcentration
sediment and to provide confidence
in the locality ofinthe
the simulations of the current, waves and patterns of suspended sediment concentration
existing detached breakwater, the simulation results obtained from the models were calibrated initially in the
locality
againstofmeasured
the existing detachedonbreakwater,
conditions 23 December the2014
simulation
to 7 Januaryresults
2015obtained from
at station the 02
1 (lat: models wereN,
◦ 480 40.02”
calibrated initially against measured conditions on 23 December 2014 to 7 January
long: 101◦ 200 11.18” E). The models were further validated against field measurements taken between 2015 at station 1
(lat: 02°48′40.02”
23 December 2014N,and long: 101°20′11.18”
7 January E). The
2015 at station models
2 (lat: 02◦ 49were
0 26” N,further
long: 101 validated
◦ 180 58.14” against
E. field
measurements
The calibration of the hydrodynamic model was carried out by adjusting the values of the N,
taken between 23 December 2014 and 7 January 2015 at station 2 (lat: 02°49′26” bed
long: 101°18′58.14” E.
roughness/Manning number over the whole computational domain [52]. The calibration of the spectra
wave Themodel
calibration of theout
was carried hydrodynamic
by adjustingmodel was of
the values carried
the wave out breaking
by adjusting the values
parameters, of thefriction
bottom bed
roughness/Manning number over the whole computational domain
parameters and white-capping (deep water wave breaking) parameters [53]. The calibration of [52]. The calibration of the
the
spectra wave model was carried out by adjusting the values of the wave breaking
mud transport model was carried out by adjusting the values of the erosion coefficient, power of parameters, bottom
friction
erosion,parameters and white-capping
settling velocity (deep
and critical shear water
stress forwave breaking)
deposition parameters
and erosion [39].[53]. The calibration
To check the accuracy
ofof
the mud transport model was carried out by adjusting the values of the erosion
2
the simulation results, the Theil’s inequality coefficients, R Squared (R ), and Root Mean Squared coefficient, power
ofError
erosion, settling
(RMSE) werevelocity and[44,48,54].
calculated critical shear stress for deposition and erosion [39]. To check the
accuracy of the simulation results,
Figure 7 compares the simulated the Theil’s
andinequality
measuredcoefficients,
water level, R Squared
current (R
2), and Root Mean
characteristics, wave
Squared Error (RMSE) were calculated [44,48,54].
characteristics and SSC at station 1. Figure 8 depicts the comparison between the simulated and
Figure water
measured 7 compares the simulated
level, current and measured
characteristics, water level,
wave characteristics current
and SSC characteristics,
at station wave
2. Table 3 presents
characteristics and SSC at station 1. Figure 8 depicts the comparison between
the values of Thiel’s coefficient, R2 and RMSE obtained during the model calibration and validation. the simulated and
measured
Based onwater level, current
the standard characteristics,
error allowed by the DIDwave characteristics
(2013) and SSCand
for hydrodynamic at station
sediment 2. transport
Table 3
presents the values of Thiel’s coefficient, R 2 and RMSE obtained during the model calibration and
modelling, the values proved that the models were calibrated and validated well.
validation. Based on the standard error allowed by the DID (2013) for hydrodynamic and sediment
transport modelling, the values proved that the models were calibrated and validated well.
Water 2019, 11, 1721 14 of 28
Water 2019, 11, x FOR PEER REVIEW 14 of 28

0.8 Measured

Current Speed
0.6 Predicted
0.4
(m/s) 0.2
0.0
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(a)

360 Measured
Current Direction

280 Predicted
200
(degree)

120
40
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(b)

3.0 Measured
1.0 Predicted
Water Level

-1.0
(m)

-3.0
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(c)

Measured
0.8
Significant Wave

Predicted
0.6
Height (m)

0.4
0.2
0.0
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Data

(d)

Measured
360
Direction (degree)

Predicted
280
Mean Wave

200
120
40
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(e)

Figure 7. Cont.
Water 2019, 11, x FOR PEER REVIEW 15 of 28

Water
Water 2019,
2019, 11,11, x FOR PEER REVIEW
1721 15ofof2828
15

0.18 Measured
0.15 Predicted
0.12 Measured
3)
0.18
SSC (kg/m 0.09
0.15 Predicted
0.06
0.12
SSC (kg/m 3)

0.03
0.09
0.00
0.06
0.03
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
0.00 Date
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date
(f)

Figure 7. Comparison between simulation results and (f) field measurements on 23 December 2014 to 7
January
Figure 2015 at station
7. Comparison 1 (long:
between 101°20′11.18”
simulation E,and
results lat: field
02°48′40.02” N), (a) on
measurements current speed; (b)
23 December current
2014 to
Figure 7. Comparison between simulation results and field measurements on 23 December 2014 to 7
direction; (c) water level; (d) significant
◦ 0 wave height; (e)
◦ mean
0 wave direction; (f) SSC.
7 January 2015 at station 1 (long: 101 20 11.18” E, lat: 02 48 40.02” N), (a) current speed; (b) current
January 2015 at station 1 (long: 101°20′11.18” E, lat: 02°48′40.02” N), (a) current speed; (b) current
direction; (c) water level; (d) significant wave height; (e) mean wave direction; (f) SSC.
direction; (c) water level; (d) significant wave height; (e) mean wave direction; (f) SSC.
0.6 Measured
SpeedSpeed

0.4 Predicted
0.6 Measured
(m/s)(m/s)

0.2 Predicted
Current

0.4
0.0
0.2
Current

21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15


0.0 Date
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date
(a)

(a)
360
Direction

Measured
280
Predicted
(degree)

360
Direction

200 Measured
280
120 Predicted
(degree)
Current

200
40
120
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Current

40 Date
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date
(b)

(b)
3.0 Measured
LevelLevel

1.0 Predicted
3.0 Measured
(m) (m)

-1.0
1.0 Predicted
Water

-3.0
-1.0
Water

21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15


-3.0 Date
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date
(c)
8. Cont.
Figure(c)
Water
Water2019, 11,11,
2019, 1721
x FOR PEER REVIEW 1616ofof
28 28

0.8 Measured
Significant Wave
0.6 Predicted
Height (m)
0.4
0.2
0.0
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(d)

360 Measured
Direction (degree)

280 Predicted
Mean Wave

200
120
40
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(e)

0.18 Measured
0.15 Predicted
SSC (kg/m3)

0.12
0.09
0.06
0.03
0.00
21-Dec-14 25-Dec-14 29-Dec-14 2-Jan-15 6-Jan-15 10-Jan-15
Date

(f)
Figure 8. Comparison between simulation results and field measurements on 23 December 2014 to
Figure 8. Comparison between simulation results and field measurements on 23 December 2014 to 7
7 January 2015 at station 2 (long: 101◦ 180 58.14” E, lat: 02◦ 490 26” N), (a) current speed; (b) current
January 2015 at station 2 (long: 101°18′58.14” E, lat: 02°49′26” N), (a) current speed; (b) current
direction; (c) water level; (d) significant wave height; (e) mean wave direction; (f) SSC.
direction; (c) water level; (d) significant wave height; (e) mean wave direction; (f) SSC.
Table 3. Statistical results for the performance of the hydrodynamic, wave and mud transport models.
Table 3. Statistical results for the performance of the hydrodynamic, wave and mud transport models.
Calibration Validation
Parameter
R Squared
Calibration
Thiel’s Inequality
Validation
R Squared Thiel’s Inequality
RMSE 2 Thiel’s RMSE 2 Thiel’s
(R ) Coefficient (R ) Coefficient
Parameter R Squared R Squared
Current RMSE inequality RMSE inequality
0.07 m/s 0.92 (R2) 0.08 0.08 m/s 0.91 (R2) 0.08
Speeds coefficient coefficient
Current
Current 15◦ 0.94 0.05 17◦ 0.93 0.06
Directions 0.07 m/s 0.92 0.08 0.08 m/s 0.91 0.08
Speeds
Water Levels 0.05 m 0.95 0.04 0.06 m 0.94 0.05
Current
Significant
0.04 m15° 0.85 0.94 0.14 0.05 0.05 m 17° 0.83 0.93 0.16 0.06
wave heights
Directions
Mean Wave
Water Levels 18◦0.05 m
Directions
0.81 0.95 0.18 0.04 19◦ 0.06 m 0.80 0.94 0.19 0.05
Significant 0.004 0.005
SSC 0.04 m 0.83 0.85 0.16 0.14 0.05 m 0.82 0.83 0.17 0.16
wave heights kg/m3 kg/m3
Mean Wave
18° 0.81 0.18 19° 0.80 0.19
Directions
Water 2019, 11, x FOR PEER REVIEW 17 of 28

Water 2019, 0.004 0.005


SSC 11, 1721 0.83 0.16 0.82 0.1717 of 28
kg/m3 kg/m3

4. Results and Discussion


4. Results and Discussion
4.1. Sediment and Water Samples Analyses
4.1. Sediment andto
According Water SamplesofAnalyses
the results sediment particle analyses, the median grain diameter (D50 ) of the
cohesive sediments
According to theinresults
the intertidal area of
of sediment Careyanalyses,
particle Island, atthedepth of 0–40
median cm,
grain was determined
diameter to
(D50) of the
be 0.015–0.022 mm, which was consisted of 10% clay, 71% silt and 19% fine sand.
cohesive sediments in the intertidal area of Carey Island, at depth of 0–40 cm, was determined to be Furthermore,
the sediment
0.015–0.022 fraction
mm, whichatwasdepths of 40–100
consisted cm consisted
of 10% clay, 71% ofsiltstiff
andclay.
19%Based on water
fine sand. samples and
Furthermore, the
sediment fraction at depths of 40–100 cm consisted of stiff clay. Based on water samples River
velocity measurements at the mouth of the Langat River, TSS 1 showed that the Langat carries
and velocity
180–261 mg/L of
measurements at suspended
the mouth of sediments
the Langat to the Malacca
River, TSS 1 Strait
showed with 698–1130
that m3 /s
the Langat of water
River discharges
carries 180–261
during ebb tide and 121–479 m 3 /s during spring tide. Table 4 present TSS
mg/L of suspended sediments to the Malacca Strait with 698–1130 m /s of water discharges
3 2 analyses from theduring
water
samples collected at 500 m seaward of the breakwater.
ebb tide and 121–479 m3/s during spring tide. Table 4 present TSS 2 analyses from the water samples
collected at 500 m seaward of the breakwater.
Table 4. Monthly amount of TSS 2 in 2015.
Table 4. Monthly amount of TSS 2 in 2015.
Month
Jan Feb Mar Apr May JunMonth
Jul Aug Sep Oct Nov Dec
TSS 2 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
3) 0.047 0.043 0.058 0.049 0.079 0.071 0.06 0.081 0.076 0.058 0.061 0.056
(kg/m
TSS 2
0.047 0.043 0.058 0.049 0.079 0.071 0.06 0.081 0.076 0.058 0.061 0.056
(kg/m3)
Table 4 shows that the amounts of TSS2 were higher during the Southwest season, when the
Table 4 shows
monthly rainfall that thewere
intensities amounts of TSS2
also higher were
(see higher
Table during
1). High the Southwest
concentration season, when
of suspended the
sediment
monthly
from the rainfall
Langat intensities
River into were also higher
the Strait (see Table
of Malacca 1). High
is expected concentration
during of suspended
the Southwest sediment
season, due to the
from
higherthe Langatintensities.
rainfall River intoThe
the wind
Strait from
of Malacca is expected
the Southeast during
direction the Southwest
transports season, due
the suspended to the
sediment
higher
from therainfall
mouth intensities. The River
of the Langat wind from
to thethe Southeast
study site. direction transports the suspended sediment
from the mouth of the Langat River to the study site.
4.2. Hydrodynamic Changes in the Locality of the Breakwater Structure
4.2. Hydrodynamic Changes in the Locality of the Breakwater Structure
Figure 9 illustrates the simulation results for the currents and waves characteristics with and
without the 9breakwater
Figure illustratesduring the NE season,
the simulation resultsforforthe maximum
the currents tidal level ofcharacteristics
and waves the neap tide and spring
with and
tide conditions.
without Figure during
the breakwater 10 presents
the NEtheseason,
simulation
for theresults
maximumfor the currents
tidal and
level of thewaves characteristics
neap tide and spring
withconditions.
tide and without the breakwater
Figure 10 presents during SW seasonresults
the simulation for thefor
maximum tidaland
the currents levelwaves
of thecharacteristics
neap tide and
spring tide conditions.

(a) (b)

Figure 9. Cont.
Water 2019, 11, 1721 18 of 28
Water 2019, 11, x FOR PEER REVIEW 18 of 28

(c) (d)

(e) (f)

(g) (h)
Figure
Figure 9. 9.Simulated
Simulatedcurrent
current and
and wave
wavecharacteristics during
characteristics thethe
during Northeast season:
Northeast (a) CS(a)
season: during NT
CS during
(WT); (b) CS during NT (W); (c) CS during ST (WT); (d) CS during ST (W); (e) SWH during
NT (WT); (b) CS during NT (W); (c) CS during ST (WT); (d) CS during ST (W); (e) SWH during NT NT (WT);
(f) SWH
(WT); (f) SWH during NT NT
during (W);(W);
(g) (g)
SWH during
SWH ST ST
during (WT);
(WT);(h)(h)
SWHSWH during
duringSTST(W). Notes:
(W). Notes: WW ==with
with
breakwater, WT = without breakwater, CS = current speed, SWH = significant wave height, NT = neap
breakwater, WT = without breakwater, CS = current speed, SWH = significant wave height, NT = neap
tide, ST = spring tide.
tide, ST = spring tide.
Water 2019, 11, 1721 19 of 28

(a) Changes in Current and Wave Characteristics during the Northeast Season (NE)
The simulation results shown in Figure 9 demonstrate that during the NE season and without the
existence of the breakwater structure, the currents flow directly to the study site (degraded mangrove
area) from a Northwest direction at approximately 300◦ to 330◦ (magnetic) towards the Southeast
direction with velocity between 0.12 and 0.22 m/s. At the same time, incident waves impose force on
the mangrove degradation areas from a Southwest direction at approximately 230◦ to 250◦ (magnetic)
to the Northeast direction, with the significant wave heights between 0.10 and 0.21 m. In addition, the
simulation results also show that the current velocity and wave heights were reduced in the nearshore
area, before reaching the study site, due to the resistance forces of the cohesive sediment. Previous
studies have reported that cohesive bed slopes have higher friction coefficients compared to sandy
beach slopes [30,31]. Therefore, current and wave turbulent energies are reduced on cohesive coastal
bed slopes.
Due to the presence of the breakwater at the study site, the current and wave characteristics
changed dynamically around the degraded mangrove area. For the maximum tidal level during neap
tide conditions, the water levels were lower than the main segment’s (S1, S2, S3) crest levels and higher
than circulation gaps’ (CG1, CG2, CG3) crest levels (the breakwater was not submerged). Currents and
waves entered the mangrove degradation area via overtopping the circulation gap’s (CG1, CG2, CG3)
of the breakwater and through the available gaps between the breakwater and earth dyke or between
the breakwater and the stone (without L-blocks). At this moment, the current speeds on the landward
side of the breakwater decreased from 0.18 to 0.10 m/s (approximately 20 m from the breakwater
structure to the landward side area) and from 0.18 to 0.04 m/s (closest to the breakwater structure).
The wave heights also decreased on the landward side of the breakwater from approximately 0.165 to
0.06 m. In addition, the current speeds slightly increased around the circulation gaps in the breakwater
(CG1, CG2, CG3), above the stone (without L-blocks) and available gaps between the breakwater and
earth dyke from 0.20 to 0.24 m/s due to the turbulent flows in these areas as the result of return flow
occurrences, while the wave heights also increased slightly at the toe of the breakwater’s main segment
(S1, S2, S3) from 0.15 m to 0.18 m due to reflection incidences [41].
At the maximum tidal level during the spring tide, the water levels were higher than the
breakwater’s crest (the breakwater was submerged); consequently, the current and wave energies
could pass through the study site via overtopping the structure. The overall current speed and wave
heights on the landward side of the breakwater (mangrove degradation area) were reduced slightly
from approximately 0.18 to 0.16 m/s and from 0.18 to 0.15 m, respectively. However, the current speeds
on the landward side of the first main segment (S1) of the breakwater were found to be reduced by a
greater amount, i.e., from 0.18 to 0.06 m/s. Additionally, current speeds and wave heights at the toe of
the main body (direction the currents/waves come from) increased from 0.20 to 0.28 m/s and from 0.18
to 0.24 m, respectively. These increases might be due to refraction occurrences (due to the differences
in water depths) combining with reflection incidences when the current and waves pass through the
breakwater at an angle to underwater contours and through shallower depths [41].
Water 2019, 11, 1721 20 of 28
Water 2019, 11, x FOR PEER REVIEW 20 of 28

(a) (b)

(c) (d)

(e) (f)

Figure 10. Cont.


Water 2019, 11, 1721 21 of 28
Water 2019, 11, x FOR PEER REVIEW 21 of 28

(g) (h)
Figure 10.
Figure 10. Simulated
Simulatedcurrent
currentand
andwave
wavecharacteristics during
characteristics during thethe
Southwest
Southwest season: (a) CS
season: (a) during NT
CS during
(WT);
NT (b) CS
(WT); (b)during NT (W);
CS during (c) CS
NT (W); (c)during ST (WT);
CS during (d) CS(d)
ST (WT); during ST (W);
CS during ST(e) SWH
(W); duringduring
(e) SWH NT (WT);
NT
(f) SWH
(WT); during
(f) SWH NT (W);
during (g) SWH
NT (W); during
(g) SWH ST ST
during (WT);
(WT);(h)(h)
SWHSWHduring
duringSTST(W).
(W). Notes:
Notes: W W== with
breakwater,WT
breakwater, WT== without
without breakwater,
breakwater,CS CS== current
current speed, SWH== significant
speed, SWH significant wave
wave height, NT == neap
height, NT
tide, ST =
= spring
spring tide.

(b) Changes
Changes in
in Current
Current and
and Wave
WaveCharacteristics
Characteristicsduring
duringthe
theSouthwest
SouthwestSeason
Season(SW)
(SW)
The simulation results shown in Figure 10 illustrate that without the coastal defence structure at
the site, the currents flow to the degraded mangrove area from a Southeast direction approximately
120 ◦ –150◦ (magnetic) towards the Northwest direction with speeds between 0.09 and 0.18 m/s,
120°–150° (magnetic) towards the Northwest direction with speeds between 0.09 and 0.18 m/s, while
while
waveswaves
force force the degraded
the degraded mangrove
mangrove areaareafrom fromthethe southbetween
south betweenapproximately
approximately170° 170◦ and 185°◦
and 185
(magnetic) withwith significant
significant heights
heights of of waves
waves between
between 0.1 0.1 and
and 0.225
0.225 m m during
duringthe theSouthwest
Southwestseason.season.
After
After construction
constructionofofthe thebreakwater
breakwater ininthetheintertidal area
intertidal of Carey
area of Carey Island, the currents
Island, the currents which flow
which
from Southeast
flow from directions
Southeast reachreach
directions the degraded
the degraded mangrove
mangrove areas first,first,
areas andand thenthen
theytheyflow seaward
flow seaward by
overtopping the circulation gaps’ of the breakwater (CG1, CG2, CG3)
by overtopping the circulation gaps’ of the breakwater (CG1, CG2, CG3) and through the available and through the available gaps
between the breakwater
gaps between the breakwaterand earth
anddyke
earthor between
dyke the breakwater
or between and stone
the breakwater and(without L-blocks)
stone (without during
L-blocks)
the neapthe
during tide. In this
neap case,
tide. currents
In this case,were reduced
currents were on reduced
the landward
on the side of the breakwater
landward side of the structure from
breakwater
approximately
structure from0.13 to 0.06 m/s, and
approximately 0.13they increased
to 0.06 m/s, and abovethey
the circulation
increased above gaps’ structures
the circulationand around
gaps’
the gaps between
structures the breakwater
and around and the stone
the gaps between (without L-blocks)/earth
the breakwater dyke from
and the stone (without approximatelydyke
L-blocks)/earth 0.13
to 0.18approximately
from m/s. During the spring
0.13 to 0.18tide,m/s.
the During
current speeds in the
the spring degraded
tide, the current mangrove
speeds areain were reduced
the degraded
from 0.17 toarea
mangrove 0.09 m/s,
werewhile the current
reduced from 0.17 speedsto slowed
0.09 m/s,quite significantly
while the current on the seaward
speeds side of
slowed the
quite
first main segment
significantly on the of seaward
the breakwater
side of(S1),
thei.e., bymain
first up tosegment
approximately
of the0.04 m/s. In addition,
breakwater (S1), i.e.,thebycurrent
up to
speeds increased0.04
approximately at the
m/s.toeInof addition,
the breakwater (direction
the current the current
speeds increased comes from)
at the toefrom
of theapproximately
breakwater
0.18 to 0.21the
(direction m/s.
current comes from) from approximately 0.18 to 0.21 m/s.
In terms of the the wave
wave characteristics,
characteristics, the wave heights in the degraded mangrove area were
reduced
reduced and andconcentrated
concentrated ononthethe
landward
landward side side
of theofbreakwater
the breakwaterstructure (from the
structure middle
(from thestructure
middle
to the earth
structure todyke).
the earthAlso,dyke).
wave Also,
heights increased
wave heights from 0.2 to 0.24
increased from m at0.2theto toe
0.24ofmtheatbreakwater
the toe of and the
stones (without
breakwater and L-blocks).
stones (without Such L-blocks).
an increase in wave
Such heightsinwas
an increase wave due to refraction
heights was dueand reflection
to refraction
incidences
and reflectionas indicated
incidences byasmechanics
indicated of bywave motionofinwave
mechanics shore protection
motion in shoremanual [41]. manual [41].
protection

4.3. Suspended Sediment


4.3. Suspended Sediment Transport
Transport in
in the
the Breakwater’s
Breakwater’s Surroundings
Surroundings
Figure
Figure 11
11 presents
presentsthethesimulations
simulationsofofsuspended
suspended sediment concentration
sediment during
concentration thethe
during NENE
season for
season
the
for maximum
the maximumtidal tidal
level of the of
level neap
thetide andtide
neap spring
andtide conditions
spring in the vicinity
tide conditions of the
in the breakwater.
vicinity of the
In addition, Figure 12 presents the simulations of suspended sediment concentration
breakwater. In addition, Figure 12 presents the simulations of suspended sediment concentration during the
during the SW season for the maximum tidal level of the neap tide and spring tide conditions around
the breakwater structure.
Water 2019, 11, 1721 22 of 28

SW season for the maximum tidal level of the neap tide and spring tide conditions around the
Water 2019, 11, x FOR PEER REVIEW 22 of 28
breakwater
Water 2019, 11, xstructure.
FOR PEER REVIEW 22 of 28

(a)
(a) (b)
(b)
Figure
Figure 11. Simulation of
of suspended sediment concentration during the Northeast season: (a)
(a) neap
Figure 11. Simulation of suspendedsediment
suspended sedimentconcentration
concentration during
during thethe Northeast
Northeast season:
season: (a) neap neap
tide;
tide;
tide;(b) spring
(b) springtide.
(b) spring tide. tide.

(a)
(a) (b)
(b)
Figure
Figure12.
Figure 12.Simulation
12. Simulationof
Simulation ofsuspended
of suspendedsediment
suspended sedimentconcentration
sediment concentrationduring
concentration duringthe
during theSouthwest
the Southwestseason:
Southwest season:(a)
season: (a)neap
(a) neap
neap
tide; (b)
tide; (b)
tide; spring
(b) springtide.
spring tide.
tide.

The averageconcentration
The concentration ofsuspended
suspended sedimentbrought brought andtransported
transported bycurrent
current andwave wave
Theaverage
average concentrationof of suspendedsediment
sediment broughtand and transportedby by currentand and wave
actions from
actions from the
theseaward sideside
of theofbreakwater to the mangrove degradationdegradation
area was approximately
actions from the seaward side of the breakwater to the 3mangrove degradation area was
seaward
3 , during the NE
the breakwater to the mangrove area was
0.048 kg/m
approximately 0.048 kg/m season
3, during the(Figure
NE 11) and
season 0.05211)
(Figure kg/m
and ,0.052
during the33, SW
kg/m season
during the (Figure
SW 12).
season
approximately 0.048 kg/m , during the NE season (Figure 11) and 0.052 kg/m , during the SW season
3
Hence, there
(Figure is a possibility that individual particles of the suspended sediments in waterincolumn
(Figure12).
12).Hence,
Hence,there
thereisisaapossibility
possibilitythatthatindividual
individualparticles
particlesof ofthe
thesuspended
suspendedsediments
sediments inwater water
becomebecome
column attachedattached
to each other and other
form flocks in theflocks
areas where the hydrodynamics conditions were
column become attached to each other and form flocks in the areas where the hydrodynamics
to each and form in the areas where the hydrodynamics
calm enough. Based
conditions on the simulation results, the concentrationsthe of suspended sediments brought
conditions were
were calmcalm enough.
enough. BasedBased on on the
the simulation
simulation results,
results, the concentrations
concentrations of of suspended
suspended
from the seaward are
sediments slightly reducing in line with the mean wave directionthe
towards the landward area.
sediments brought
brought fromfrom thethe seaward
seaward are are slightly
slightly reducing
reducing in in line
line with
with the meanmean wave
wave direction
direction
This result
towards may have been obtained because somebeensuspended sediment settles down on the sea-bed
towards the landward area. This result may have been obtained because some suspendedsediment
the landward area. This result may have obtained because some suspended sediment
settles
settlesdown
downon onthe
thesea-bed
sea-beddue duetotothe
thereduction
reductionof ofwave
waveandandcurrent
currentenergies
energiestriggered
triggeredby bythethehigh
high
bottom friction coefficients of the cohesive sediments
bottom friction coefficients of the cohesive sediments [31,55]. [31,55].
The
The simulation
simulation results
results of of mud
mud transport
transport module
module (Figures
(Figures 11 11 and
and 12)
12) show
show thatthat there
there are
are
reductions in the SSC of the water column at the study site after the construction
reductions in the SSC of the water column at the study site after the construction of the breakwater, of the breakwater,
Water 2019, 11, 1721 23 of 28

due to the reduction of wave and current energies triggered by the high bottom friction coefficients of
the cohesive sediments [30,31].
Water The
2019,simulation
11, x FOR PEER REVIEW
results 23 of 28
of mud transport module (Figures 11 and 12) show that there are reductions
in the SSC of the water column at the study site after the construction of the breakwater, especially in
especially
the landward in thearealandward area of the
of the breakwater breakwater
during during
the NE and SW the NE and
seasons. SW seasons.
In addition, In addition,
reductions in the
reductions in the SSC are higher around the first main segment of the breakwater (S1) (seaward
SSC are higher around the first main segment of the breakwater (S1) (seaward and landward areas) and
landward areas)
during the SW season. during the SW season.

4.4. Variation of Sea-Bed Levels


Levels and
and Erosion-Deposition
Erosion-DepositionPattern
Patternininthe
theLocality
Localityofofthe
theLow-Crested
Low-CrestedBreakwater
Breakwater
Figure 13 presents the variation of sea-bed levels and the pattern of erosion-deposition in the
Figure
locality of the13breakwater
presents in
thethe
variation
degradedofenvironment,
sea-bed levels
i.e., and the pattern
the cohesive of erosion-deposition
intertidal in
area of Carey Island,
the
measured every two months from December 2014 to October 2015.

(a) (b)

(c) (d)

Figure 13. Cont.


Water 2019, 11, 1721 24 of 28
Water 2019, 11, x FOR PEER REVIEW

(e)
Figurelevels
Figure 13. Variation of the sea-bed 13. and
Variation of the sea-bed
erosion-deposition levels
pattern andlocality
in the erosion-deposition pattern in the locality of
of the breakwater
breakwater
structure over two-month periods fromstructure
Decemberover
2014two-month
to October periods from +
2015. Notes: December 2014
represents the to October 2015. Notes
represents
deposition in cm, (a,b) NE season, theSW
(c,d,e) deposition
season. in cm, (a, b) NE season, (c, d, e) SW season.

The erosion-deposition patterns for the cohesive patterns


The erosion-deposition sedimentfor in the cohesive
intertidalsediment
area of Careyin theIsland
intertidal area of Carey
caused by the construction of theby
caused 85-mthe long, low-crested
construction breakwater
of the 85-m long,are evident. Unlike
low-crested sand coasts,
breakwater are evident. Unlike sand
deposition patterns on cohesive
depositioncoasts are more
patterns dynamic.
on cohesive Depositions
coasts are found
are more dynamic. to be higher
Depositions are at
found to be higher a
lower sea-bed elevations in protected
sea-bed areas with
elevations calm hydrodynamic
in protected conditions.
areas with calm Based onconditions.
hydrodynamic Figure 13, Based on Figure
it is evident that there is evident
an increment of sea-bed
that there is an elevations
increment in of the degraded
sea-bed mangrove
elevations in thearea after
degraded mangrove ar
construction of the structure. During the NE season, the sediment accumulation rates were higher
construction of the structure. During the NE season, the sediment accumulation rates were hi
on the landward side of main segmentsside
the landward of theofbreakwater
main segments (S1, S2, S3), breakwater
of the while they were lower
(S1, S2, S3),on the they were lower
while
landward side of the circulation
landward gaps’ in of
side thethe
breakwater
circulation (CG1,
gaps’CG2,
in theCG3). During(CG1,
breakwater the SW season,
CG2, CG3).theDuring the SW sea
sediment accumulations are much higher
sediment surrounding
accumulations the first
are much main
higher segment of the
surrounding the first
breakwater (S1), of the breakwa
main segment
while lower amounts accumulate around
while lower the circulation
amounts accumulate gaps’ in thethe
around breakwater
circulation and thein
gaps’ gapthebetween
breakwater and the gap b
the breakwater and the stone (without L-blocks).
the breakwater and the stone (without L-blocks).
Overall, the sea-bed elevations in the
Overall, thedegraded
sea-bed environment
elevations ininthe thedegraded
intertidal environment
area of Carey Island
in the intertidal area o
increased by approximately 8 cm increased
Island near to theby first main segment8ofcm
approximately thenear
breakwater (S1),
to the first while
main other areas
segment of the breakwater (S1
on the landward side of the breakwater increased approximately 4 cm, on average, from December
other areas on the landward side of the breakwater increased approximately 4 cm, on averag
2014 to October 2015. December 2014 to October 2015.

4.5. Discussion 4.5. Discussion


The hydrodynamic characteristics, cohesive characteristics,
The hydrodynamic behaviour of the sediments,
cohesive concentration
behaviour of
of the sediments, concentra
suspended sediment and suspended
existence ofsediment
a low-crest detached breakwater at the study site have resulted
and existence of a low-crest detached breakwater at the study site have r
in specific patterns of sediment transport
in specific and
patterns oferosion-deposition
sediment transportinand
theerosion-deposition
cohesive intertidalinzone
the of
cohesive intertidal
Carey Island during both Carey
the SWIsland
and NE seasons.
during both the SW and NE seasons.

4.5.1. Sediment Transport and Erosion-Deposition Pattern during the Northeast Season
4.5.1. Sediment Transport and Erosion-Deposition Pattern during the Northeast Season
Water flows from the nearshore area bring suspended sediments to the study site.
Water flows from the nearshore area bring suspended sediments to the study site. Figur
Figures 9, 11 and 13 show the flow patterns, sediment transport and erosion-deposition in the locality
and 13 show the flow patterns, sediment transport and erosion-deposition in the locality
of the detached breakwater during NE season. During the neap tide, the detached breakwater obstructs
detached breakwater during NE season. During the neap tide, the detached breakwater obstr
the transportation of some suspended sediments from the seaward side to the mangrove degradation
transportation of some suspended sediments from the seaward side to the mangrove degr
area. It was found that the concentrations of suspended sediments are higher on the seaward side than
area. It was found that the concentrations of suspended sediments are higher on the seawa
the landward side of the breakwater. Some suspended sediments can enter the degraded mangrove
than the landward side of the breakwater. Some suspended sediments can enter the de
mangrove area through the gap between the breakwater and earth dyke or stone (without L
and also, by overtopping the circulation gaps’ structure. The concentrations of suspended sed
Water 2019, 11, 1721 25 of 28

area through the gap between the breakwater and earth dyke or stone (without L-blocks) and also,
by overtopping the circulation gaps’ structure. The concentrations of suspended sediments in the
water column were reduced further on the landward side of the breakwater because the sediment
settles in this area due to the calm hydrodynamic conditions created by the breakwater and sediments
are trapped [8,9].
During the spring tide, the suspended sediments from the seaward direction entered the degraded
mangrove area via overtopping the breakwater structure. Some suspended sediments in the water
column formed flocks and settle down in regions with lower hydrodynamic energies within the
degraded mangrove area, while others were transported back seawards, all depending on turbulent
properties resultant of the wave breaking and wave-current interactions. Suspended sediments
containing flocks have higher settling velocities compared to individual particles; hence, they can
settle and be deposited faster in regions with lower turbulent energies. The observed results were
in line with the findings of previous studies conducted by Pejrup and Zhu et al. [37,38]. As current
speeds and wave heights were reduced, the amount of sediment deposited increases in the landward
area behind the breakwater, thus, the sediment accumulations were also higher during the NE season.
The results were found to be in line with the field measurement (Figure 13a,b) which indicated that
sediment accumulations were concentrated in the landward area behind the breakwater.

4.5.2. Sediment Transport and Erosion-Deposition Pattern during the Southwest Season
During the neap tide and spring tide in the SW season, the detached breakwater trapped suspended
sediments before they flowed seaward from the degraded mangrove area. The suspended sediments
settled in areas with lower hydrodynamic conditions, especially in the vicinity of the first main segment
of the breakwater (S1). According to Figure 12, a lower amount of suspended sediment concentrations
appeared around the first main segment of the breakwater (S1) due to the calmer hydrodynamic
conditions (with current speeds of less than 0.07 m/s) and higher settling rate. The field measurement
confirmed that sediment accumulations were concentrated around the first main segment of the
breakwater structure and in the landward area behind the breakwater (Figure 13c–e).
Sediment accumulations were found to be slightly higher during the SW season compared to the
NE season. This higher amount could be attributed to the calmer hydrodynamic conditions during
the SW season. In addition, a higher amount of the suspended sediments were transported from the
seaward to the study site during SW season. Calm hydrodynamic conditions and high suspended
sediment concentrations could have accelerated the sediment deposition rate [4].
Finally, minor sea-bed elevation changes, with an average increment of 4 cm, were recorded
in the landward area behind the breakwater after a year, which could have been influenced by the
consolidation of the deposited cohesive sediment as indicated by several researchers as Kirby, Edmonds,
& Ranasinghe [27,36,55,56].

5. Conclusions
The numerical modelling study showed that without having the breakwater in the degraded
environment of the intertidal area, waves and curernts directly reach to the bare area at the study site
with higher turbulent energy which means that in absence of the breakwater more intense degredation
will occur in the coastal mangrove reserve area.
The numerical model found that the pattern of suspended sediment around the breakwater during
both seasons is strongly affected by the local hydrodynamic conditions. It was shown that less energetic
wave-current conditions effectively reduced the amount of SSC in the water column, as the suspended
sediments were deposited in the water column under calm hydrodynamic conditions. In addition,
the patterns of erosion-deposition in the locality of a low-crested breakwater on a cohesive coast were
more dynamically evolving compared to a sandy coasts. Also, it was found that the deposition rate
and sea-bed elevations increased more prominent under calm nearshore hydrodynamics.
Water 2019, 11, 1721 26 of 28

Both field measurements and numerical modelling indicated that sediment accumulation increased
in the degraded environment of the intertidal area of Carey Island with the presence of the detached
breakwater. Field study showed increase in the sea-bed elevations, particularly near the first segment
of breakwater structure. A smaller increased in bed elevations occured in the landward area behind
the breakwaters. The results showed that the presence of the low-crested detached breakwater lead
into sediment accumulations in the degraded environment in the intertidal area of Carey Island and
reduce the erosion problem at the site. Sediment deposition during the calm seasons could create
a suitable tidal regime for mangrove survival. The overall results indicated that conduction of the
breakwater structure is vital to reduce the erosion problems in cohesive coasts of Carey Island and
play a key role in success of mangrove rehabilitation projects in the region.

Author Contributions: A.F. and R.H. conceived and designed the study; A.F. performed data analyses; A.F. and
S.A. wrote the manuscript; all authors (A.F., S.A., K.M. and R.H.) revised the manuscript for publication.
Funding: This work was funded by the High Impact Research, University of Malaya [Grant Number
UM.C/HIR/MOHE/ENG/47] and the Ministry of Higher Education, Malaysia for TRGS/1/2015/UKM/02/5/1
& TRGS/1/2015/UKM/02/5/3.
Acknowledgments: The authors would like to thank the Department of Survey and Mapping Malaysia (JUPEM),
the Department of Meteorology, Malaysia and the National Hydrography Centre, Malaysia for sharing the
important data that enabled us to undertake this study.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Hashim, R.; Kamali, B.; Tamin, N.M.; Zakaria, R. An integrated approach to coastal rehabilitation: Mangrove
restoration in Sungai Haji Dorani, Malaysia. Estuar. Coast. Shelf Sci. 2010, 86, 118–124. [CrossRef]
2. Hashim, R.; Fitri, A.; Motamedi, S.; Hashim, A.M. Modeling of Coastal Hydrodynamic Associated with
Coastal Structures: A Review. Malays. J. Sci. 2013, 32, 149–154.
3. Scheffer, F. Rubble Mound Breakwaters for the New Port of Ennore (India)-Evaluation of Construction.
Master’s Thesis, Delft University of Technology, Delft, The Netherlands, 1999.
4. Saied, U.; Tsanis, I.K. A Coastal Area Morphodynamics Model. Environ. Model. Softw. 2008, 23, 35–49.
[CrossRef]
5. Fairley, I.; Davidson, M.; Kingston, K. The Morpho-Dynamics of a Beach Protected by Detached Breakwaters
in a High Energy Tidal Environment. J. Coast. Res. 2009, 56, 598–607.
6. Nam, P.T.; Larson, M.; Hanson, H. A Numerical Model of Beach Morphological Evolution Due to Waves and
Currents in the Vicinity of Coastal Structures. Coast. Eng. 2011, 58, 863–876. [CrossRef]
7. Nam, P.T.; Larson, M.; Hanson, H. Modeling Morphological Evolution in the Vicinity of Coastal Structures.
Coast. Eng. Proc. 2011, 1, 68. [CrossRef]
8. Barbaro, G.; Foti, G. Shoreline behind a Breakwater: Comparison between Theoretical Models and Field
Measurements for the Reggio Calabria Sea. J. Coast. Res. 2012, 29, 216–224. [CrossRef]
9. Abolfathi, S.; Yeganeh-Bakhtiary, A.; Hamze-Ziabari, S.M.; Borzooei, S. Wave Runup Prediction Using M5’
Model Tree Algorithm. Ocean Eng. 2016, 112, 76–81. [CrossRef]
10. Fan, D.; Guo, Y.; Wang, P.; Shi, J.Z. Cross-Shore Variations in Morphodynamic Processes of an Open-Coast
Mudflat in the Changjiang Delta, China: With an Emphasis on Storm Impacts. Cont. Shelf Res. 2006, 26,
517–538. [CrossRef]
11. Kamali, B. Design, Construction and Testing of L-Blok Detached Breakwater for Coastal Rehabilitation.
Ph.D. Thesis, University of Malaya, Kuala Lumpur, Malaysia, 2011.
12. Wang, H.; Wang, A.; Bi, N.; Zeng, X.; Xiao, H. Seasonal Distribution of Suspended Sediment in the Bohai Sea,
China. Cont. Shelf Res. 2014, 90, 17–32. [CrossRef]
13. Fitri, A.; Hashim, R.; Song, K.I.; Motamedi, S. Evaluation of Morphodynamic Changes in the Vicinity of
Low-Crested Breakwater on Cohesive Shore of Carey Island, Malaysia. Coast. Eng. J. 2015, 57, 1550023.
[CrossRef]
14. Pang, C.; Li, K.; Hu, D. Net Accumulation of Suspended Sediment and Its Seasonal Variability Dominated by
Shelf Circulation in the Yellow and East China Seas. Mar. Geol. 2016, 371, 33–43. [CrossRef]
Water 2019, 11, 1721 27 of 28

15. Zhang, H. Transport of Microplastics in Coastal Seas. Estuar. Coast. Shelf Sci. 2017, 199, 74–86. [CrossRef]
16. Joshi, S.; Xu, Y.J. Bedload and Suspended Load Transport in the 140-km Reach Downstream of the Mississippi
River Avulsion to the Atchafalaya River. Water 2017, 9, 716. [CrossRef]
17. Fernández-Montblanc, T.; Izquierdo, A.; Quinn, R.; Bethencourt, M. Waves and Wrecks: A Computational
Fluid Dynamic Study in an Underwater Archaeological Site. Ocean Eng. 2018, 163, 232–250. [CrossRef]
18. Dong, S.; Salauddin, M.; Abolfathi, S.; Tan, Z.H.; Pearson, J.M. The Influence of Geometrical Shape Changes
on Wave Overtopping: A Laboratory and SPH Numerical Study. Coasts Mar. Struct. Break. 2018, 1217–1226.
[CrossRef]
19. Abolfathi, S.; Shudi, D.; Borzooei, S.; Yeganeh-Bakhtiari, A.; Pearson, J. Application of smoothed particle
hydrodynamics in evaluating the performance of coastal retrofit structures. Coast Eng. Proc. 2018. [CrossRef]
20. Dean, R.G.; Chen, R.J.; Browder, A.E. Full scale monitoring study of a submerged breakwater, Palm Beach,
Florida, USA. Coast. Eng. 1997, 29, 291–315. [CrossRef]
21. Van Rijn, L.C. Coastal erosion and control. Ocean Coast. Manag. 2011, 54, 867–887. [CrossRef]
22. Sravanthi, N.; Ramakrishnan, R.; Rajawat, A.S.; Narayana, A.C. Application of Numerical Model in
Suspended Sediment Transport Studies along the Central Kerala, West-Coast of India. Aquat. Procedia 2015,
4, 109–116. [CrossRef]
23. Archetti, R.; Zanuttigh, B. Integrated Monitoring of the Hydro-Morphodynamics of a Beach Protected by
Low Crested Detached Breakwaters. Coast. Eng. 2010, 57, 879–891. [CrossRef]
24. Lamberti, A.; Zanuttigh, B. An Integrated Approach to Beach Management in Lido Di Dante, Italy. Estuar. Coast.
Shelf Sci. 2005, 62, 441–451. [CrossRef]
25. Martinelli, L.; Zanuttigh, B.; Lamberti, A. Hydrodynamic and Morphodynamic Response of Isolated and
Multiple Low Crested Structures: Experiments and Simulations. Coast. Eng. 2006, 53, 363–379. [CrossRef]
26. Zanuttigh, B. Numerical Modelling of the Morphological Response Induced by Low-Crested Structures in
Lido Di Dante, Italy. Coast. Eng. 2007, 54, 31–47. [CrossRef]
27. Kirby, R. Chapter Four Distinguishing Accretion from Erosion-Dominated Muddy Coasts. In Muddy Coasts
of the World; Healy, T., Wang, Y., Healy, J.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2002; Volume 4,
pp. 61–81.
28. Baas, J.H.; Davies, A.G.; Malarkey, J. Bedform Development in Mixed Sand–mud: The Contrasting Role of
Cohesive Forces in Flow and Bed. Geomorphology 2013, 182, 19–32. [CrossRef]
29. Shi, Z.; Chen, J.Y. Morphodynamics and Sediment Dynamics on Intertidal Mudflats in China (1961–1994).
Cont. Shelf Res. 1996, 16, 1909–1926. [CrossRef]
30. Birben, A.R.; Ozolcer, I.H.; Karasu, S.; Komurcu, M.I. Investigation of the effects of offshore breakwater
parameters on sediment accumulation. Ocean Eng. 2007, 34, 284–302. [CrossRef]
31. United States Army, Corps of Engineers Coastal Engineering Research Center (U.S.). Shore Protection Manual:
Mechanics of Wave Motion: Volume 1, Chapter 2; Department of the Army, Waterways Experiment Station:
Washington, DC, USA, 1984.
32. Nikmanesh, M.R.; Talebbeydokhti, N. Numerical Simulation for Predicting Concentration Profiles of Cohesive
Sediments in Surf Zone. Sci. Iran. 2013, 20, 454–465. [CrossRef]
33. Chang, K.H.; Tsaur, D.H.; Huang, L.H. Accurate Solution to Diffraction around a Modified V-Shaped
Breakwater. Coast. Eng. 2012, 68, 56–66. [CrossRef]
34. De Jong, R.J. Wave Transmission at Low Crested Structures. Master’s Thesis, Delf University of Technology,
Delf, The Netherlands, 1996.
35. Young, D.M. A Laboratory Study on the Effects of Submerged Vertical and Semicircular Breakwaters on
Near-Field Hydrodynamics and Morphodynamics. Master’s Thesis, Clemson University, Clemson, SC,
USA, 2008.
36. Edmonds, D.A.; Slingerland, R.L. Significant effect of sediment cohesion on delta morphology. Nat. Geosci.
2010, 3, 105. [CrossRef]
37. Pejrup, M.; Mikkelsen, O.A. Factors controlling the field settling velocity of cohesive sediment in estuaries.
Estuar. Coast. Shelf Sci. 2010, 87, 177–185. [CrossRef]
38. Zhu, Z.; Yu, J.; Wang, H.; Dou, J.; Wang, C. Fractal dimension of cohesive sediment flocs at steady state under
seven shear flow conditions. Water 2015, 7, 4385–4408. [CrossRef]
39. DHI. MIKE 21 Mud Transport Scientific Documentation; DHI Water and Environment: Hesholm, Denmark, 2014.
Water 2019, 11, 1721 28 of 28

40. Jain, M. Wave Attenuation and Mud Entrainment in Shallow Waters. Ph.D. Thesis, University of Florida,
Gainesville, FL, USA, 2007.
41. Yeganeh-Bakhtiary, A.; Houshangi, H.; Hajivalie, F.; Abolfathi, S. A Numerical Study on Hydrodynamics
of Standing Waves in Front of Caisson Breakwaters with WCSPH Model. Coast. Eng. J. 2017, 59,
1750005-1–1750005-31. [CrossRef]
42. Abolfathi, S.; Pearson, J.M. Application of smoothed particle hydrodynamics (SPH) in nearshore mixing:
A comparison to laboratory data. Coast. Eng. Proc. 2017, 1, 16. [CrossRef]
43. Eissa, S.S.; Lebleb, A.A. Numerical Modeling of Nearshore Wave Conditions at Al Huwaisat Island, KSA.
Aquat. Procedia 2015, 4, 79–86. [CrossRef]
44. Patra, S.K.; Mohanty, P.K.; Mishra, P.; Pradhan, U.K. Estimation and Validation of Offshore Wave
Characteristics of Bay of Bengal Cyclones (2008–2009). Aquat. Procedia 2015, 4, 1522–1528. [CrossRef]
45. Komen, G.J.; Cavaleri, L.; Donelan, M.; Hasselmann, K.; Hasselmann, S.; Janssen, P. Dynamics and Modelling
of Ocean Waves; Cambridge University Press: Cambridge, UK, 1994.
46. Young, I. Wind Generated Ocean Waves; Elsevier: Amsterdam, The Netherlands, 1999; Volume 2.
47. Mehta, A.J.; Hayter, E.J.; Parker, W.R.; Krone, R.B.; Teeter, A.M. Cohesive sediment transport. I: Process
description. J. Hydraul. Eng. 1989, 115, 1076–1093. [CrossRef]
48. Jose, F.; Stone, G.W. Forecast of Nearshore Wave Parameters Using MIKE-21 Spectral Wave Model; Gulf Coast
Association of Geological Societies Transactions: Tulsa, OK, USA, 2006; Volume 56, pp. 323–327.
49. Jose, F.; Kobashi, D.; Stone, G.W. Spectral Wave Transformation over an Elongated Sand Shoal off South
Central Louisiana, USA. J. Coast. Res. 2007, 50, 757–761.
50. Sørensen, O.R.; Kofoed-Hansen, H.; Rugbjerg, M.; Sørensen, L.S. A third-generation spectral wave model
using an unstructured finite volume technique. Coast. Eng. 2004, 4, 894–906.
51. Department of Irrigation and Drainage. Guideline for Hydrodynamic Modelling; Department of Irrigation and
Drainage: Kuala Lumpur, Malayisa, 2013.
52. DHI. MIKE 21 Hydrodynamic Scientific Documentation; DHI Water and Environment: Hesholm, Denmark, 2014.
53. DHI. MIKE 21 Spectra Wave Scientific Documentation; DHI Water and Environment: Hesholm, Denmark, 2014.
54. Song, J.; Wei, L.; Ming, Y. A method for simulation model validation based on Theil’s inequality coefficient and
principal component analysis. In Proceedings of the Asian Simulation Conference, Singapore, 6–8 November
2013; Springer: Berlin/Heidelberg, Germany, 2013; pp. 126–135.
55. Ranasinghe, R.; Turner, I.L. Shoreline response to submerged structures: A review. Coast. Eng. 2006, 53,
65–79. [CrossRef]
56. Ranasingle, R.; Larson, M.; Savioli, J. Shoreline response to a single shore-parallel submerged breakwater.
Coast. Eng. 2010, 57, 1006–1017. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like