Open access
Technical Papers
Apr 5, 2024

Relating Geotechnical Sediment Properties and Erodibility at a Sandy Beach

Publication: Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 150, Issue 4

Abstract

Geotechnical sediment properties, morphological change, and hydrodynamics were measured as part of the During Nearshore Event Experiment (DUNEX) in October 2021 at the sandy Atlantic side beach in Duck, North Carolina. In this study, direct relationships were explored between in situ soil properties and direct erodibility measurements in the context of morphological change. Moisture content, grain size, total density, relative density, void ratio, and sediment strength were compared to bed-level change using ground-based LiDAR and erodibility parameters from laboratory jet erosion tests (JETs) conducted along a cross-shore transect stretching from the dune toe to the lower intertidal zone. Directly relating changes in sediment properties to changes in morphology from LiDAR proved inconclusive due to the complex interactions between sediments, hydrodynamics, and morphology, even on a local scale, but initial observations and possibly impacting factors were discussed. Void ratio and total unit weight correlated well with the detachment rate coefficient from JETs, with denser sediments testing as less erodible. In situ sediment strength measurements related—as expected—to total unit weight, void ratio, and water content, with increases in firmness factor associated with increases in total unit weight, void ratio, and water content. These strength measurements were also found to have a correlation with the detachment coefficient from the JET, with stronger/firmer sediments being less erodible than weaker ones.

Introduction

Coastal areas are subject to increasingly difficult challenges due to climate change. Sea level rise and increases in the frequency and severity of coastal storms are expected to lead to increasing beach erosion, shoreline retreat, and flood events (Sallenger 2000; Zhang et al. 2004; Jones et al. 2009; Bender et al. 2010; Knutson et al. 2010). Central to improving the prediction and mitigation capabilities of these impacts is enhancing the understanding of sediment dynamics in coastal areas. Specifically, the role of sediment properties beyond grain size, for example, geotechnical sediment properties and their relationship to driving mechanisms of erosion, is still not well understood. Such relationships have the potential to improve the prediction of sediment dynamics and to help with mitigating climate-related coastal change and impacts on coastal communities.
Erosion is typically modeled as a fluid–particle interaction, and thus, grain size and particle density predominantly govern the prediction of erodibility (Shields 1936; Briaud et al. 2001). For fine-grained sediments, there have been numerous attempts to correlate a wide variety of sediment properties with erodibility (NASEM 2019). Properties such as water content, unit weight, plasticity index, undrained shear strength, void ratio, swell potential, and fines fraction, among others have been tested with various degrees of success (Cao et al. 2002; NASEM 2019). Developing empirical relationships between critical shear stress and the aforementioned soil properties from remolded laboratory samples represents the state of the art for determining the erosion behavior of fine-grained sediments, and these types of relationships are often site-specific, owing to the complex nature of the erosion problem and the influence of a large number of sediment-specific properties on erodibility (Briaud 2013). For coarse-grained sediments, sands, and gravels, grain size is accepted as the controlling parameter for erodibility relationships (Briaud 2013). For many beach environments, sand is the predominant sediment type, and thus, sediment dynamics are predicted mostly based on grain size, if sediment properties are considered at all. For example, Cheng et al. (2023) examined the relationship between beach morphology, sediment size, and sediment shape and found that sediment size and shape had an impact on sediment transport regimes and could indicate different types of beach morphology. However, geotechnical sediment properties such as moisture content, density, and shear strength have been shown to be important for a variety of related coastal processes such as dune retreat, tidal flat evolution, and Aeolian sediment transport (Davidson-Arnott et al. 2005; Erikson et al. 2007; Sassa and Watabe 2007; Zambrano-Cruzatty et al. 2019) and may be particularly important in the intertidal zone that is subject to saturation and drainage and variations in forcing conditions on different time scales (waves, tides, inundation events). While geomorphodynamics of coastal areas are driven by local hydrodynamics (Dalrymple 1992; Masselink and Short 1993; Dean and Dalrymple 2004; Masselink et al. 2006; van der Lugt et al. 2019; Hsu et al. 2023), it is hypothesized that including geotechnical sediment properties in erodibility relationships could further improve the prediction of coastal change by understanding sediment stabilizing forces better.
Challenges of conducting geotechnical measurements in energetic coastal environments have been approached in recent years. Tools such as portable free-fall penetrometers (PFFPs) have enabled conducting geotechnical surveys of energetic coastal areas (Stark and Kopf 2011; Stark et al. 2014; Albatal and Stark 2017; Albatal et al. 2019). When deployed in submerged coastal areas, PFFPs have been used for sediment characterization and for estimating the bearing strength, relative density of sands, undrained shear strength, and soil behavior under dynamic loading (Akal and Stoll 1995; Chow et al. 2018; Lucking et al. 2017; Albatal et al. 2020; Kiptoo et al. 2020). More recently, PFFPs have been tested for geotechnical surveying in emerged intertidal environments (Reeve et al. 2018; Brilli and Stark 2020). The ability of a single tool to collect geotechnical measurements seamlessly from the emerged to submerged coastal zones is essential to integrate geotechnical sediment properties in sediment transport prediction and simulation at beaches. The goal of this study was to relate geotechnical sediment properties, erodibility parameters, morphodynamics, and hydrodynamic regime in a sandy beach environment. To this end, a field survey was conducted at the United States Army Corp of Engineers Field Research Facility (USACE FRF) in October 2021 in the framework of the During Nearshore Event Experiment (DUNEX). The field survey included in situ testing and the collection of sediment samples for subsequent laboratory erodibility testing. Based on the collected field data and physical sample analysis, sediment properties and mobility were assessed and related to bed-elevation change.

Methods

Field Measurements

From October 11 to 15, 2021, three transects (A–C) were surveyed daily on the Atlantic-side sand beach in Duck, North Carolina, located on the Outer Banks barrier islands (Fig. 1, top right corner). Each transect was surveyed at 5-m increments from the local dune toe to the waterline at low tide (Fig. 1). Measurements at each station involved moisture content measurements using a handheld moisture gauge calibrated for the effects of seawater, 10 cm push cores for density and grain size analysis, and estimates of sediment strength from a PFFP. Additionally, thirteen 25-cm push cores were collected randomly throughout the week at stations of interest, and these cores were saved to use in a jet erosion test (JET). The longer core length for JET samples was taken to ensure enough in situ sample was preserved for laboratory testing because the JET can accommodate samples longer than 10 cm. An additional 10-cm core was taken at the same location as the JET samples to provide sediment properties because the JET is a destructive test. Table 1 provides a detailed breakdown of the field campaign, highlighting the survey times and number of measurements performed along each transect. Over the 5-day survey period, 118 individual stations were surveyed, with 5 PFFP deployments, moisture content readings, and a push core taken per station. Measurements with multiple readings per station were averaged to produce a single value per station.
Fig. 1. Google Earth satellite image of the field site in Duck, North Carolina, highlighting the Measurement transects A, B, and C, locations of the LiDAR scanners (stars), and locations of ADVs (triangles) (lower right corner: N36°10′54.13″ W75°45′7.30″).
(Map data © 2017 Google, Terrametrics.)
Table 1. Summary of survey times and types/numbers of measurements performed
DateSurvey timeTransect ATransect BTransect C
StationsPFFP (5/stn)Moisture content (5/stn)Push coresStationsPFFP (5/stn)Moisture content (5/stn)Push coresStationsPFFP (5/stn)Moisture content (5/stn)Push cores
October 11PM low tide630306735357735357
October 12AM low tide630306735357735357
October 12PM high tide315153525255525255
October 13AM low tide630306735357735357
October 14AM low tide735357735357945456
October 15AM low tide630306735357945459
Moisture contents were measured using a Dynamax SM150T soil moisture sensor, which measures the dielectric permittivity of the substrate to estimate the in situ moisture content. The device was calibrated by preparing sand from the site to specific moisture contents and salinities, and developing a site-specific calibration curve that also accounts for the effects of seawater (Brilli et al. 2022). Local morphology was surveyed via terrestrial LiDAR measured throughout the week (stars, Fig. 1). LiDAR scans were collected from two sites: from the dune using a Riegl VZ-1000 LiDAR scanner (left star) and from the pier using a Riegl Z390i LiDAR scanner (right star). Three-dimensional spatial scans of the beach were performed hourly, with a peak laser pulse repetition rate of 70 kHz and point densities ranging from 101 to 104 points per square meter, depending on the distance from the scanner (O’Dea et al. 2019; US Army Corp of Engineers 2022).
Characterization of in situ sediment strength was performed using the PFFP blueDrop. The instrument is approximately streamlined in shape, with a length of 63.1 cm and a mass of 7.71 kg when equipped with a conical tip (Fig. 2). Five onboard vertical accelerometers ranging from 2 to 250g (where g is the acceleration due to gravity) measure continuously, along with two horizontal ±55g tilt accelerometers.
Fig. 2. PFFP blueDrop deployed in the swash at the USACE FRF.
The deceleration of the probe from impact with the sediment until stoppage (dec) is used with the mass of the probe (m) in Newton’s second law to obtain a sediment resistance force (Fsoil), and dividing by the area of the tip (A) gives an estimate of the dynamic bearing capacity of the soil, qdyn:
qdyn=FsoilA=m×decA
(1)
The deceleration profile (dec) can be single- and double-integrated to obtain the velocity and penetration depth, respectively. In emerged environments, the probe impacts at 2–5 m/s, which is more than 100 times faster than other standard testing methods such as the cone penetration test (CPT), which penetrates at 2 cm/s. These high strain rates can cause an increase in the measured bearing capacity compared to results from tests at lower strain rates (Dayal and Allen 1973). This effect can be accounted for by introducing a strain-rate correction (fsr) to normalize the results to a constant, slower strain rate. The following equation presents the logarithmic form of this equation, often used in submerged sandy environments (Stoll et al. 2007; Stark et al. 2009; Stephan et al. 2015; Albatal et al. 2020):
fsr=1+Klog10(vvref)
(2)
where K = empirical strain-rate correction factor; v = dynamic velocity; and vref = reference velocity typically taken as 2 cm/s to model a CPT. K is soil-dependent and is typically taken as 0–1.5 for submerged sandy environments (True 1976; Dayal et al. 1975; Stark et al. 2012; Stephan et al. 2015; Albatal et al. 2020). For emerged beach environments at the same site, good agreement with theoretically modeled bearing capacity was found using a velocity-dependent correction factor, Kmod=0.31×vi, where vi is the impact velocity in m/s (Brilli 2023). It should be noted that this relationship was found to break down at volumetric water contents >25% (Brilli 2023). This modification of K can be used in Eq. (2) to give a modified strain rate correction for emerged beach environments (fsr−mod). qdyn is divided by the strain rate correction to give an estimate of the equivalent static strength of the sediment, called the quasi-static bearing capacity (qsbc) (Stark et al. 2012), and is given by the following equation:
qsbc=qdynfsrmod
(3)
The modified strain rate correction is used in this study because it is presently the only strain-rate correction explored for PFFP’s in emerged, partially saturated environments. An additional metric used in this study for sediment characterization from PFFP measurements is the firmness factor (FF), developed by Mulukutla et al. (2011) for submerged environments:
FF=amaxvigtp
(4)
where amax = maximum measured deceleration during penetration; and tp = penetration time. This parameter is a proxy for soil stiffness and attempts to normalize the PFFP record for the effects of impact velocity because the measurement is sensitive to changes in vi. The firmness factor has been used in conjunction with qsbc to successfully characterize sediment types based on PFFP deployments in submerged environments (Albatal and Stark 2017).

Laboratory Testing

The 10-cm push cores were carefully collected at the beach and were processed in the lab to obtain grain size and density information about the samples. Samples were collected in tubes, sealed, and tested in their in situ condition. No samples were reconstituted for testing. Total density (ρtotal) and total unit weight (γtotal) were computed using the mass of sediment and water and the volume of the tubes (−350 cm3), given by the following equation:
γtotal=gρtotal=gMTVT
(5)
where g = acceleration due to gravity; MT = total mass of sediment and water; and VT = total volume of the tube. The sample was oven-dried and weighed to obtain dry density (ρd) and dry unit weight (γd), computed via Eq. (6) and converted to the void ratio (e) using Eq. (7):
γd=gρd=gMdVT
(6)
e=VvVs=Gsγwγd1
(7)
where Md = dry mass of sediment; Vv = volume of voids; Vs = volume of solids; Gs = specific gravity of the sediment taken as 2.65 for the quartz sand at this site; and γw = unit weight of salt water. Once density testing was completed, the remaining dry sample was used to determine grain size statistics, namely, the median grain size (d50), via sieve stack analysis in accordance with ASTM D6913/ASTM D6913M-17 (ASTM 2017b).
Laboratory characterization of erodibility was performed on fifteen 25-cm push cores collected at various stations throughout the week using the JET setup at North Carolina State University. The JET was developed as a field and laboratory tool for testing the erosion behavior of fine-grained soils (Hanson 1990). The laboratory tests were conducted using the mini-JET device, as shown in Fig. 3. The device was connected to a constant pressure head tank fed by a recirculating water system, ultimately creating a submerged jet of water through the mini-JET that impinged vertically on the sediment sample. Pressure head settings ranging between 1.2 and 2 m were used in this study. Before testing, the elevation of the soil surface was recorded using the device’s point gauge (Fig. 3). During testing, the jet of water was run for given periods of time to simulate erosion, measuring the scour depth after each period. After three consecutive and approximately equal scour depth measurements, the time period was increased, and the testing procedure was repeated. The time periods used herein were 5, 15, 30 s, 1, 2, and 5 min. Testing ended after all time periods were completed.
Fig. 3. Laboratory setup of the jet erosion test.
The erodibility parameters that are derived from the JET are the critical shear stress (τcr) and the erodibility coefficient (kd). The former characterizes the threshold entrainment force, and the latter the detachment rate after the onset of erosion. kd can be described as a stress-normalized erosion rate, whereby at the same magnitude of shear stress, a soil with a higher value of kd will have a higher erosion rate than a soil with a lower value of kd. These parameters are typically obtained by analytically solving the erosion rate equation [Eq. (8)], a linear excess shear model for scour via the jet (Hanson and Cook 1997):
dJdt=kd[τJp2J2τcr]
(8)
where J = measured depth of scour; dJ/dt = measured erosion rate; τ = estimated shear stress from the velocity of the water jet; and Jp = length of the jet core from the nozzle. There are multiple methods for solving this equation to obtain τcr and kd (Hanson and Cook 2004; Simon and Thomas 2010; Daly et al. 2013; Al-Madhhachi et al. 2013; Wahl 2021). The method selected for the present study was linear regression of the raw erosion rate and shear stress data, which does not require iteratively solving Eq. (8) [see Wahl (2021) for a comprehensive comparison of available methods]. This method was selected because it provided the best fit to the data of the four methods tested, the others being the Blaisdell method (Hanson and Cook 2004), the iterative method (Simon and Thomas 2010), and the scour depth method (Daly et al. 2013). Model performance for each of the methods was quantified using a normalized objective function, suggested by Al-Madhhachi et al. (2013) for JET analysis.

Incipient Motion Analysis

In addition to the laboratory erodibility testing, analytical modeling was conducted to estimate the incipient motion criteria for the field samples under the given hydrodynamic forcing condition. This was accomplished by estimating the critical Shields parameter (θcr), which provides a nondimensionalization of the critical shear stress (τcr) for incipient motion (Shields 1936). If the in situ Shields parameter (θ), which, like θcr, is a nondimensional shear stress due to the forcing conditions, is greater than θcr, particle motion will occur. It should be noted that the only sediment property included in the estimates presented here is d50. The first estimate of θcr is from Brownlie (1981), which is a regression fit to the original Shields (1936) data in terms of the dimensionless grain size (D), and is given by the following equations:
θcr=0.22D0.9+0.06exp(17.77D0.9)
(9)
D=[g(Gs1)ν2]1/3d50
(10)
where ν = kinematic viscosity of the fluid. The second estimate is from Soulsby and Whitehouse (1997), which also fits the Shields data with the addition of data points under waves and mixed wave/current regimes and is also a function of the dimensionless grain size:
θcr=0.31+1.2D+0.055(1exp(0.02D))
(11)
Once θcr was computed for each inundated location, the original Shields (1936) formulation was solved to give the critical shear velocity, ucr, the flow velocity that will initiate particle motion, given by the following equation:
ucr=θcr(Gs1)gd50
(12)
Three Nortek Vector acoustic Doppler velocimeters (ADVs) were installed in the swash zone near Transect C (Fig. 1, triangles). This array of sensors performed recording at 8 Hz continuously throughout the week and collected three-dimensional velocity measurements in the swash. It should be noted that only cross-shore directed currents were used from this data set. Additionally, Fig. 1 shows the sensors to be emerged, but this is a historical image, and the sensors were generally in the swash at low tide and submerged at high tide. The critical velocity estimates were then compared to the measured swash flows at the site to determine if the measured flows exceeded the threshold for motion.

Results

Fig. 3 shows the measured significant wave height from a bottom-mounted acoustic wave and current sensor (AWAC) over the course of the week (US Army Corp of Engineers 2022). During a storm event on October 11, significant wave heights were > 2 m and gradually calmed over the next 5 days, reaching heights of <1 m by October 15 [Fig. 4(a)]. The peak period mostly fluctuated between 7 and 10 s with a decreasing trend toward the end of the week [Fig. 4(b)].
Fig. 4. Significant (a) wave height; (b) and peak period versus time over the course of the survey week.
Fig. 5 displays morphological change over the course of the survey from ground-based LiDAR measurements (US Army Corp of Engineers 2022). Transect A underwent the most morphologic change over the course of the week, with deposition occurring from October 11 to October 13, with Station A4 seeing 60 cm of deposition, followed by slight erosion from October 13 to October 15 [Fig. 5(a)]. Transect B remained relatively stable, with erosion/deposition magnitudes on the order of <20 cm [Fig. 5(b)]. Transect C also remained stable in the upper reaches, with a notable deposition area on the order of 40 cm from October 13 to October 15 [Fig. 5(c)]. The diamonds on each transect represent the highest water level seen on the transect for that particular survey interval using total water level data available at the site (US Army Corp of Engineers 2022). Water levels were measured hourly, and the data points presented in Fig. 5 are the highest water levels observed in the 24 h between surveys. Any stations below the maximum water level were subjected to hydrodynamic forcing, and thus, only these stations were considered for further analysis. A possible contributor to the variation between transects in close proximity (−200 m spacing) is longshore geomorphodynamic features, mostly the presence of beach cusps (see Fig. 1 for a historical qualitative example) on the order of 5 m in the alongshore direction and the position of the transect in relation to the cusp. The presence of cusps was also apparent in the LiDAR scans taken during the survey period. Beach cusps are a common occurrence at this site, typically forming in the days following storm events, with a mean spacing of 12.4–38.8 m (Holland 1998). While longshore processes were not within the scope of this work, it is acknowledged that these processes affect morphology.
Fig. 5. Elevation data over time for (a) Transect A; (b) Transect B; and (c) Transect C, along with station locations. Diamonds indicate the maximum water level observed during the intrasurvey period, with the color of each diamond corresponding to the date. The cross-shore distance is measured relative to the local dune toe.
Fig. 6(a) depicts a histogram of the sample median grain sizes (d50) for all samples taken over the course of the week. Samples were taken around low tide on all days. All of the samples were classified as poorly graded sand (SP) based on grain size analysis and the Unified Soil Classification System (USCS; ASTM D2487/D2487-17, ASTM 2017a). A majority of the samples had a d50 classifying as fine sand via USCS, delineated by particles passing a No. 40 sieve (0.43 mm) and retained on a No. 200 sieve (0.074 mm) (dashed lines, Fig. 6) (ASTM D2487/D2487-17, ASTM 2017a). Fig. 6(b) shows the cross-shore variation of d50. On the x-axis, zero indicates the dune toe, with the swash zone located near the 25–30 m station, depending on the low tide level on a given day. As seen from the figure, d50 exhibited little variation in the cross shore until the lower intertidal zone where grain size and variability increased. Fig. 6(c) shows examples of grain size distributions in the swash zone of Transect C throughout the week, highlighting the variability of particle sizes.
Fig. 6. (a) Histogram of sample d50; (b) cross-shore variation of d50; and (c) representation grain size distributions. Dashed lines indicate the grain size limits of fine sand. Black dots in the middle box and whisker plots indicate outliers (1.5IQR). The cross-shore distance is measured relative to the local dune toe.
Table 2 shows the results of the incipient motion analysis. Using grain sizes from the 51 stations that were shown to be inundated during the surveys, ucr was calculated via the two methods presented in Section “Incipient Motion Analysis,” and the range of threshold velocities is presented. Additionally, ucr was computed from measurements of τcr from the JET samples, and the range of these values is also presented. ADV data from sensors in the swash zone were available during the survey period to characterize the in situ flows. Typical values of flow velocity, u, ranged from 60 to 80 cm/s on October 11, decreasing to 30 cm/s on October 15, corresponding to the decrease in wave heights over the course of the week (Fig. 4). Maximum flows were on the order of 300 cm/s. As seen from Table 2, the measured flows greatly exceeded the threshold for motion for the entirety of the survey period.
Table 2. Comparison of estimated critical velocities to measured swash flows
Threshold velocity, ucrRange of values (cm/s)Measured swash flows
utypical (cm/s)umax (cm/s)
ucr (Brownlie 1981)1.36–2.5230–80300
ucr (Soulsby and Whitehouse 1997)1.33–2.40
ucr (JET samples)2.01–3.73
Fig. 7 shows the relationship between firmness factor, qsbc, and void ratio. In the plot, the shading of each data point differs with regard to the measured moisture content. Firmness factor [Fig. 7(a)] and qsbc [Fig. 7(b)] both correlate negatively with the void ratio. Additionally, samples with higher densities (low e), qsbc, and FF consistently had higher water contents. While this may appear counterintuitive, this can be explained by the location of these samples, which were found in and near the swash zone and, thus, were near or at full saturation (Fig. 7).
Fig. 7. (a) Firmness factor versus void ratio; and (b) qsbc versus void ratio, with warmer colors of the data points indicating higher moisture contents and vice versa.
Changes in sediment properties were then compared to morphologic change. Fig. 8 depicts changes in the bed levels at inundated stations as a function of corresponding changes in void ratio [Fig. 8(a)], firmness factor [Fig. 8(b)], and qsbc [Fig. 8(c)]. Changes in properties and bed levels were computed over the 24 h interval between surveys. Of the 25 inundated stations where LiDAR data were available along with sediment and strength properties to make comparisons, 22 stations saw zero change or deposition, while only 3 stations appeared erosional. As seen in the figure, there are no readily apparent trends between the soil properties and morphologic changes. However, it can be observed that: (1) all sites where void ratio decreased the bed level increased, but negative changes in void ratio were overall smaller than the observed positive changes; and (2) an increase in bed level was more often accompanied by a decrease in beach surface strength or hardness (i.e., decrease in firmness factor and qsbc), while an increase in firmness factor or qsbc was mostly associated with little change in bed elevation.
Fig. 8. (a) Changes in the void ratio; (b) firmness factor; and (c) qsbc with respect to corresponding changes in the bed level at inundated stations.
Finally, sediment packing and strength properties, e, γtotal, FF, and qsbc were compared to the detachment rate coefficient, kd, from the JET experiments (Fig. 9). As mentioned in the “Laboratory Testing” section, kd was determined through linear regression of the raw erosion rate/shear stress data. Each data point represents a location where measurements were taken for the PFFP, a 25-cm JET sample, and a 10-cm density sample simultaneously. Scattering was low, with R2 ranging from 0.80 to 0.99, with 10 of the 13 samples tested having an R2 value > 0.97. For the directly measured sediment properties, increasing void ratio and decreasing total unit weight correlated well with increasing kd [Figs. 9(a and b)]. Total unit weight was included here to capture the effects of water content. Both properties were similarly well-correlated with kd, with R2 = 0.87 for e and R2 = 0.84 for γtotal. Penetrometer measurements, FF and qsbc, were both negatively correlated with kd [Figs. 9(c and d)]. Although these relationships were generally less well-defined than the sediment properties, qsbc provided a stronger correlation than FF, with R2 = 0.67 and R2 = 0.43, respectively. These four properties were also compared to τc from the JET experiments, but no significant correlation was observed.
Fig. 9. Detachment rate coefficient from the JET versus (a) void ratio; (b) total unit weight; (c) firmness factor; and (d) qsbc.

Discussion

The modeled values of ucr compared to the measured flows, u, show that motion would be initiated under the hydrodynamic forcing for all inundated stations sampled (Table 2). The range of threshold velocity, ucr, from the JET was larger than the two grain-size-based methods, but it still predicted motion for all samples. This result is consistent with theory and the observations because sand is the easiest particle size to move on the Shields curve and geomorphological change is observed throughout the survey period (Shields 1936). Having the estimates of ucr from the JET is a useful addition to the analysis because this test method captures the in situ properties and variability in each physical sample more thoroughly than grain-size-based estimates. The measured flows are consistent with the literature because swash bore velocities have been observed to travel at up to 3 m/s, matching the maximum velocities observed in this study (Nielsen 1992; Puleo et al. 2000). While the presented methodology provides a useful proof of initiation of motion, and thus, sediment mobility at this sandy beach, it offers little information on the actual geomorphodynamics since erosion, deposition, and fairly unchanged conditions have been observed at the different locations throughout the short survey period.
The results presented in Fig. 8 indicate that changes in the void ratio and PFFP strength properties exhibit little to no direct relationship with morphologic change. The relatively limited number of data points available for comparison (i.e., 25) compared to the total number of soil measurements in the data set (i.e., 118) presents one of the challenges when developing more reliable relationships. This can be attributed to a variety of factors. First, not all stations were exposed to hydrodynamic forcing because they were above the high-water mark between survey intervals (Fig. 4), and thus, these stations were removed from the analysis, leaving 52 stations remaining. Not all hourly LiDAR coverage was available from the survey period because the high traffic of people and equipment on the beach rendered many of the scans unusable, requiring the available data to be averaged over the course of each day. Further, the limited available data were significantly biased toward deposition over erosion (22 depositional sites to 3 erosional sites). Results shown in Fig. 8 suggest that decreases in void ratio tended to predict deposition and that deposition was accompanied by a decrease in beach surface strength. However, these results must be considered biased by limited data points being associated with erosion and possible differences in the time since we are looking at net deposition or erosion over a certain period during which the specific site may have undergone different phases of erosion and deposition and variability in moisture contents affecting the strength measurements, among other possible impacts. Some of these factors may be addressed through changes in the survey strategy. Data collection could be improved in future work by collecting denser spatial measurements at locations below the high-water mark, increasing the frequency of sediment property measurements to capture more variability, increasing the temporal density of bed level measurements, specifically near survey times, and collecting a more representative data set encompassing both depositional and erosional behaviors. A further complicating factor in making direct comparisons between sediment properties and morphodynamics, as evidenced by Fig. 8, is the inherent variability and complexity of coastal and nearshore processes. Sediment properties are highly variable on spatiotemporal scales (wave-by-wave), and comparing changes in sediment properties from one snapshot in time to morphologic change over a 24-h period likely does not capture the variability at the necessary resolution required for making definitive associations between sediment properties and morphologic changes. Even though the three transects were relatively closely spaced, the differences in morphology and their evolution over the course of the survey (Fig. 4) highlight the complexity of sediment transport processes and their relation to local hydrodynamic forcing, even on a relatively small spatial scale. Differences in local grain size within survey periods or between data points could further complicate feedback between wave energy, local morphology, and sediment properties. Of the 112 sediment samples collected, 56 contained at least 20% medium sand (0.43–2 mm), 10 contained at least 5% coarse sand (2–4.75 mm), and 7 contained gravel size particles (>4.75 mm), highlighting the variability in the grain size distributions not captured by d50. This variability is highlighted in Fig. 6(c), where a series of representative grain size distributions from the swash zone of Transect C (C6 and C7) show a large range of d50 and also a wide range of particle sizes present in the samples. This variability in grain sizes has been documented at this site and identified as a potential driver of localized morphodynamics (Gallagher et al. 2016). Thus, while the authors conclude that available data are not complete enough to draw direct associations between sediment properties and morphologic changes, this data set offers an initial attempt and highlights pathways and challenges for future studies.
The results of JET testing on the in situ samples provided useful insights. Void ratio, total unit weight, and strength properties from the PFFP were related to kd from the JET. Fig. 9 shows clear relationships between the detachment rate coefficient, kd, and void ratio/total unit weight [Figs. 9(a and b)]. Additionally, kd exhibited a negative correlation with FF and qsbc, albeit a weaker relationship than with the sediment properties. Lower kd indicates less erodible sediment (in terms of erosion rate), and higher kd indicates more erodible sediment (Hanson 1991). Denser, stronger sediments are less erodible than loose, weaker sediments (Fig. 9). While this is intuitively and theoretically expected, this data set represents a rare direct comparison made between in situ soil properties and direct measurements of erodibility, and the analysis offers a quantitative measure of how erodibility is affected by the void ratio, total unit weight, and sediment strength based on field measurements. It should be noted that the value of kd depends on the chosen solution methodology, in this case, a linear regression of the erosion rate versus shear stress data, which is supported by the results of Wahl (2021). The outputs of this model will be used for comparison amongst results at the study site and not in relation to the magnitude of erodibility parameters compared to other sediments. While this study was conducted over a narrow range of grain sizes, void ratios, and total unit weights, specifically for beach sands, the results highlight the variability in erodibility present even over a small range of properties, which is often overlooked in studies relating soil properties over a larger range of soil types (NASEM 2019).
Fig. 9 shows a weaker relationship between FF or qsbc and kd, with the trend showing a tendency for lower FF (within the sandy sediments) to indicate less erodible sediment and vice versa. This result can be understood more clearly when considered in the context of Fig. 7. In Fig. 7, there is a clear relationship between FF, qsbc, and void ratio, with the general trend being increasing density with increasing FF and qsbc. However, there is a significant spread in the data. As seen from the color gradient, the result is also highly sensitive to changes in water content, with an increase in water content corresponding to an increase in FF and qsbc. However, this is likely due to increasing water content causing increased viscous effects due to high-strain rate, artificially increasing amax (Martin et al. 2009). The modified strain rate factor [Eq. (3)] represents a first step toward accounting for these effects, but the relationship breaks down at water contents > 25% by volume due to the aforementioned viscous effects, which could explain why the spread in the data was greater at these high water contents (Fig. 7; Brilli 2023). Improvements in understanding partially saturated strain-rate effects on PFFP measurements, specifically at high water contents, could help reduce the spread of these data. If void ratios and unit weights of beach sands can be quantitatively obtained from PFFP measurements, it would follow that PFFP data could be used to estimate erodibility and predict morphodynamics as discussed previously in this section. Overall, the results show relationships between geotechnical sediment properties and erodibility and support the hypothesis that these properties should be included in future models of beach evolution.

Conclusions

The goal of this study is to measure geotechnical sediment properties in a sandy intertidal beach environment and investigate the relationships between those properties, erodibility, and local geomorphodynamics. These relationships are tested using a field data set collected at the USACE FRF in Duck, North Carolina, in October 2021 and JET erosion testing of obtained soil samples. Measurements include moisture content, grain size, density, and sediment strength via PFFP deployments. Morphological changes are recorded from terrestrial LiDAR, and hydrodynamics are measured using an offshore bottom-mounted AWAC and ADVs measuring in the swash zone. Theoretical sediment mobility modeling and laboratory erodibility testing are also performed. It should be highlighted that variations in sediment properties are tested within mostly one sediment type, fine sand with d50 ranging from 0.3 to 0.43 mm (with some outliers in the range of 0.9–1 mm) and total unit weights and void ratios ranging from 14 to 22 kN/m3 and from 0.39 to 0.96, respectively, exploring small-scale variations in the void ratio/unit weight and associated strength and the relationship with erodibility and geomorphodynamic changes. The key takeaways are listed as follows:
1.
The Shields approach using just grain size compared to measured swash flows shows sediment mobility for all stations but is too simplistic for the limited range in sediments to provide further information on geomorphodynamics, which ranged from erosion to stable morphology to deposition.
2.
Void ratio is positively correlated with detachment coefficient, kd, from the JET, indicating that denser sediments are less erodible than looser sediments. Thus, the void ratio correlates to the amount of sediment mobilized by the hydrodynamic forcing condition, with denser areas mobilizing less sediment and vice versa.
3.
Portable free-fall penetrometer sediment strength estimates are correlated with total unit weight, void ratio, water content, and detachment coefficient. There is a significant spread in the data, which could be mitigated by focusing future work on understanding strain-rate behavior in partially saturated emerged environments with particular emphasis on zones of high water content.
Overall, this paper presents a rare field data set to compare in situ soil properties to direct measurements of erodibility, specifically in the context of coastal sediment transport, and the results show that sediment properties, namely, void ratio, change with and affect sediment erodibility.

Data Availability Statement

All data, models, or codes generated or used during the study are available from the corresponding author upon reasonable request. Data are also available from an online repository in accordance with funder data retention policies (Brilli et al. 2023).

Acknowledgments

The authors thank Julie Paprocki, Matthew Florence, Saurav Shrestha, Jonathan Moore, and the scientists and staff at the USACE FRF for their support in data collection efforts. Special thanks go to Dr. Britt Raubenhiemer for providing ADV data. These data were collected as part of the During Nearshore Event Experiment (DUNEX), which was facilitated by the U.S. Coastal Research Program (USCRP). The authors thank USCRP for supporting this effort through funding for logistics and coordination. This material is based upon work supported by the National Science Foundation under Grant CMMI-1751463. Any opinions, findings, conclusions, or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation.

References

Akal, T., and R. D. Stoll. 1995. “Expendable penetrometer for rapid assessment of seafloor parameters.” In Proc., Oceans Conf. Record. New York: IEEE.
Albatal, A., and N. Stark. 2017. “Rapid sediment mapping and in situ geotechnical characterization in challenging aquatic areas.” Limnol. Oceanogr. Methods 15 (8): 690–705. https://doi.org/10.1002/lom3.10192.
Albatal, A., N. Stark, and B. Castellanos. 2020. “Estimating in situ relative density and friction angle of nearshore sand from portable free-fall penetrometer tests.” Can. Geotech. J. 57 (1): 17–31. https://doi.org/10.1139/cgj-2018-0267.
Albatal, A., H. Wadman, N. Stark, C. Bilici, and J. McNinch. 2019. “Investigation of spatial and short-term temporal nearshore sandy sediment strength using a portable free fall penetrometer.” Coastal Eng. 143: 21–37. https://doi.org/10.1016/j.coastaleng.2018.10.013.
Al-Madhhachi, A. T., G. J. Hanson, G. A. Fox, A. K. Tyagi, and R. Bulut. 2013. “Measuring soil erodibility using a laboratory ‘mini’ jet.” Trans. ASABE 56 (3): 901–910. https://doi.org/10.13031/trans.56.9742.
ASTM. 2017a. Standard practice for classification of soils for engineering purposes (unified soil classification system). ASTM D2487/D2487-17. West Conshohocken, PA: ASTM.
ASTM. 2017b. Standard test methods for particle-size distribution (gradation) of soils using sieve analysis. ASTM D6913/ASTM D6913M-17. West Conshohocken, PA: ASTM.
Bender, M. A., T. R. Knutson, R. E. Tuleya, J. J. Sirutis, G. A. Vecchi, S. T. Garner, and I. M. Held. 2010. “Modeled impact of anthropogenic warming on the frequency of intense Atlantic hurricanes.” Science 327: 5964. https://doi.org/10.1126/science.1180568.
Briaud, J.-L. 2013. Geotechnical engineering: Unsaturated and saturated soils. Hoboken, NJ: Wiley.
Briaud, J. L., F. C. K. Ting, H. C. Chen, Y. Cao, S. W. Han, and K. W. Kwak. 2001. “Erosion function apparatus for scour rate predictions.” J. Geotech. Geoenviron. Eng. 127 (2): 105–113. https://doi.org/10.1061/(asce)1090-0241(2001)127:2(105).
Brilli, N. C. 2023. “Influence of geotechnical properties on sediment dynamics, erodibility, and geomorphodynamics in coastal environments based on field measurements.” Ph.D. thesis, Dept. of Civil and Environmental Engineering, Virginia Tech.
Brilli, N. C., and N. Stark. 2020. “Variations in sediment strength across a sandy peninsula.” In Proc., Geo-Congress 2020, 769–788. Reston, VA: ASCE.
Brilli, N. C., N. Stark, and C. Castro-Bolinaga. 2023. Data associated with influence of geotechnical properties on sediment dynamics, erodibility, and geomorphodynamics in coastal environments based on field measurements. Blacksburg, VA: Virginia Tech Univ. Libraries.
Brilli, N. C., N. Stark, B. Raubenhiemer, S. Elgar, and B. Korka. 2022. “A simple laboratory calibration method for mitigating seawater effects on soil moisture sensors.” In Proc., Int. Con. on Coastal Engineering 2022. Reston, VA: ASCE.
Brownlie, W. B. 1981. “Prediction of flow depth and sediment discharge in open channels.” Ph.D. dissertation, California Institute of Technology.
Cao, Z., R. Day, and S. Egashira. 2002. “Coupled and decoupled numerical modeling of flow and morphological evolution in alluvial rivers.” J. Hydraul. Eng. 128 (3): 306–321. https://doi.org/10.1061/(asce)0733-9429(2002)128:3(306).
Cheng, W., S. Chen, J. Zhu, X. Zhong, J. Hu, and J. Guo. 2023. “Identification of the sediment movement mechanism via grain size and shape: A case study of a beach in eastern Hainan Island in south China.” Water 15 (20): 3637. https://doi.org/10.3390/w15203637.
Chow, S. H., B. Bienen, and M. F. Randolph. 2018. “Rapid penetration of piezocones in sand.” In Proc., 4th Int. Symp. on Cone Penetration Testing. Boca Raton, FL: CRC Press.
Dalrymple, R. A. 1992. “Prediction of storm/normal beach profiles.” J. Waterw. Port Coastal Ocean Eng. 118 (2): 193–200. https://doi.org/10.1061/(asce)0733-950x(1992)118:2(193).
Daly, E. R., G. A. Fox, A. T. Al-Madhhachi, and R. B. Miller. 2013. “A scour depth approach for deriving erodibility parameters from jet erosion tests.” Trans. ASABE 56 (6): 1343–1351. https://doi.org/10.13031/trans.56.10350.
Davidson-Arnott, R. G., K. MacQuarrie, and T. Aagaard. 2005. “The effect of wind gusts, moisture content and fetch length on sand transport on a beach.” Geomorphology 68 (1–2): 115–129. https://doi.org/10.1016/j.geomorph.2004.04.008.
Dayal, U., and J. H. Allen. 1973. “Instrumented impact cone penetrometer.” Can. Geotech. J. 10 (3): 397–409. https://doi.org/10.1139/t73-034.
Dayal, U., J. H. Allen, and J. M. Jones. 1975. “Use of an impact penetrometer for the evaluation of the in-situ strength of marine sediments.” Mar. Geotechnol. 1 (2): 73–89. https://doi.org/10.1080/10641197509388155.
Dean, R. G., and R. A. Dalrymple. 2004. Coastal processes with engineering applications. Cambridge, UK: Cambridge University Press.
Erikson, L. H., M. Larson, and H. Hanson. 2007. “Laboratory investigation of beach scarp and dune recession due to notching and subsequent failure.” Mar. Geol. 245 (1–4): 1–19. https://doi.org/10.1016/j.margeo.2007.04.006.
Gallagher, E., H. Wadman, J. McNinch, A. Reniers, and M. Koktas. 2016. “A conceptual model for spatial grain size variability on the surface of and within beaches.” J. Mar. Sci. Eng. 4 (2): 37. https://doi.org/10.3390/jmse4020038.
Hanson, G. J. 1990. “Surface erodibility of earthen channels at high stresses. Part II—Developing an in situ testing device.” Trans. ASAE 33 (1): 0132–0137. https://doi.org/10.13031/2013.31306.
Hanson, G. J. 1991. “Development of a jet index to characterize erosion resistance of soils in earthen spillways.” Trans. ASAE 34 (5): 2015–2020. https://doi.org/10.13031/2013.31831.
Hanson, G. J., and K. R. Cook. 1997. “Development of excess shear stress parameters for circular jet testing.” In Proc., American Society of Agricultural Engineers. St. Joseph, MN: American Society of Agricultural Engineers (ASAE).
Hanson, G. J., and K. R. Cook. 2004. “Apparatus, test procedures, and analytical methods to measure soil erodibility in situ.” Appl. Eng. Agric. 20 (4): 455–462. https://doi.org/10.13031/2013.16492.
Holland, K. T. 1998. “Beach cusp formation and spacings at duck, USA.” Cont. Shelf Res. 18 (10): 1081–1098. https://doi.org/10.1016/S0278-4343(98)00024-7.
Hsu, C.-E., K. Serafin, X. Yu, C. Hegermiller, J. C. Warner, and M. Olabarrieta. 2023. “Total water levels along the South Atlantic Bight during three along-shelf propagating tropical cyclones: Relative contributions of storm surge and wave runup.” Nat. Hazards Earth Syst. Sci. 23 (12): 3895–3912. https://doi.org/10.5194/nhess-2023-49.
Jones, B. M., K. M. Hinkel, C. D. Arp, and W. R. Eisner. 2009. “Modern erosion rates and loss of coastal features and sites, Beaufort Sea coastline, Alaska.” Arctic 61 (4): 361–372. https://doi.org/10.14430/arctic44.
Kiptoo, D., N. Stark, G. Massey, C. Wright, and C. T. Friedrichs. 2020. “Strain rate effects in soft estuarine soils using portable free fall penetrometers.” In Proc., Ocean Science Meeting. Washington, DC: American Geophysical Union (AGU).
Knutson, T. R., J. L. McBride, J. Chan, K. Emanuel, G. Holland, C. Landsea, I. Held, J. P. Kossin, A. K. Srivastava, and M. Sugi. 2010. “Tropical cyclones and climate change.” Natl. Geosci. 3: 157–163. https://doi.org/10.1038/ngeo779.
Lucking, G., N. Stark, T. Lippmann, and S. Smyth. 2017. “Variability of in situ sediment strength and pore pressure behavior of tidal estuary surface sediments.” Geo-Mar. Lett. 37 (5): 441–456. https://doi.org/10.1007/s00367-017-0494-6.
Martin, B. E., W. Chen, B. Song, and S. A. Akers. 2009. “Moisture effects on the high strain-rate behavior of sand.” Mech. Mater. 41 (6): 786–798. https://doi.org/10.1016/j.mechmat.2009.01.014.
Masselink, G., A. Kroon, and R. G. D. Davidson-Arnott. 2006. “Morphodynamics of intertidal bars in wave-dominated coastal settings—A review.” Geomorphology 73 (1–2): 33–49. https://doi.org/10.1016/j.geomorph.2005.06.007.
Masselink, G., and A. D. Short. 1993. “The effect of tide range on beach morphodynamics and morphology: A conceptual beach model.” J. Coastal Res. 9 (3): 785–800. http://www.jstor.org/stable/4298129.
Mulukutla, G. K., L. C. Huff, J. S. Melton, K. C. Baldwin, and L. A. Mayer. 2011. “Sediment identification using free fall penetrometer acceleration–time histories.” Mar. Geophys. Res. 32 (3): 397–411. https://doi.org/10.1007/s11001-011-9116-2.
NASEM (National Academy of Sciences Engineering and Medicine). 2019. Relationship between erodibility and properties of soils. Washington, DC: National Academic Press.
Nielsen, P. 1992. Coastal bottom boundary layers and sediment transport. Singapore: World Scientific Publishing, Advanced Series on Ocean Engineering.
O’Dea, A., K. L. Brodie, and P. Hartzell. 2019. “Continuous coastal monitoring with an automated terrestrial Lidar scanner.” J. Mar. Sci. Eng. 7 (2): 37. https://doi.org/10.3390/jmse7020037.
Puleo, J. A., R. A. Beach, R. A. Holman, and J. S. Allen. 2000. “Swash zone sediment suspension and transport and the importance of bore-generated turbulence.” J. Geophys. Res.: Oceans 105 (C7): 17021–17044. https://doi.org/10.1029/2000jc900024.
Reeve, B., N. Stark, and P. Mewis. 2018. “Cross-shore variation in sediment strength at a sandy beach.” Coastal Engineering Proceedings 36: 83. https://doi.org/10.9753/icce.v36.sediment.83.
Sallenger, J. 2000. “Storm impact scale for barrier islands.” J. Coastal Res. 16 (3): 890–895. https://www.jstor.org/stable/4300099.
Sassa, S., and Y. Watabe. 2007. “Role of suction dynamics in evolution of intertidal sandy flats: Field evidence, experiments, and theoretical model.” J. Geophys. Res.: Earth Surf. 112 (1). https://doi.org/10.1029/2006JF000575.
Shields, I. A. 1936. “Application of similarity principles and turbulence research to bed-load movement.” Ph.D. dissertation, Prussian Research Institute for Hydraulic Engineering and Shipbuilding.
Simon, A., and R. E. Thomas. 2010. “Comparison and experiences with field techniques to measure critical shear stress and erodibility of cohesive deposits.” In Proc., 2nd Joint Federal Interagency Conf. Las Vegas, NV: Advisory Committee on Water Information (ACWI).
Soulsby, R. L., and R. J. S. Whitehouse. 1997. “Threshold of sediment motion in coastal environments.” In Proc., Pacific Coasts and Ports 1997 Conf. Christchurch, New Zealand: University of Canterbury.
Stark, N., G. Coco, K. R. Bryan, and A. Kopf. 2012. “In-situ geotechnical characterization of mixed-grain-size bedforms using a dynamic penetrometer.” J. Sediment. Res. 82 (7): 540–544. https://doi.org/10.2110/jsr.2012.45.
Stark, N., A. E. Hay, and G. Trowse. 2014. “Cost-effective geotechnical and sedimentological early site assessment for ocean renewable energies.” In Proc., 2014 Oceans—St. John’s, OCEANS 2014. New York: IEEE.
Stark, N., and A. Kopf. 2011. “Detection and quantification of sediment remobilization processes using a dynamic penetrometer.” In Proc., OCEANS’11—MTS/IEEE Kona, Program Book. New York: IEEE.
Stark, N., A. Kopf, H. Hanff, S. Stegmann, and R. Wilkens. 2009. “Geotechnical investigations of sandy seafloors using dynamic penetrometers.” In Proc., MTS/IEEE Biloxi—Marine Technology for Our Future: Global and Local Challenges, OCEANS 2009. New York: IEEE.
Stephan, S., N. Kaul, and H. Villinger. 2015. “Validation of impact penetrometer data by cone penetration testing and shallow seismic data within the regional geology of the Southern North Sea.” Geo-Mar. Lett. 35 (3): 203–219. https://doi.org/10.1007/s00367-015-0401-y.
Stoll, R. D., Y.-F. Sun, and I. Bitte. 2007. “Seafloor properties from penetrometer tests.” IEEE J. Oceanic Eng. 32 (1): 57–63. https://doi.org/10.1109/JOE.2007.890943.
True, D. G. 1976. “Undrained vertical penetration into ocean bottom soils.” Ph.D. thesis, Dept. of Civil Engineering, Univ. of California-Berkeley.
US Army Corp of Engineers. 2022. “FRF Data.” Coastal Hydraulics Lab (CHL) Data Server. Accessed October 22, 2022. https://chldata.erdc.dren.mil/thredds/catalog/frf/catalog.html.
van der Lugt, M. A., E. Quataret, A. van Dongeren, M. van Ormondt, and C. R. Sherwood. 2019. “Morphodynamic modeling of the response of two barrier islands to Atlantic hurricane forcing.” Estuarine Coastal Shelf Sci. 229: 106404. https://doi.org/10.1016/j.ecss.2019.106404.
Wahl, T. L. 2021. “Methods for analyzing submerged jet erosion test data to model scour of cohesive soils.” Trans. ASABE 64 (3): 785–799. https://doi.org/10.13031/TRANS.14212.
Zambrano-Cruzatty, L., A. Yerro, and N. Stark. 2019. “Influence of shear strength and moisture content on Aeolian sand erosion.” In Proc., Geo-Congress 2019. Reston, VA: ASCE.
Zhang, K., B. C. Douglas, and S. P. Leatherman. 2004. “Global warming and coastal erosion.” Clim. Change 64 (1–2): 41–58. https://doi.org/10.1023/B:CLIM.0000024690.32682.48.

Information & Authors

Information

Published In

Go to Journal of Waterway, Port, Coastal, and Ocean Engineering
Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 150Issue 4July 2024

History

Received: Mar 27, 2023
Accepted: Jan 18, 2024
Published online: Apr 5, 2024
Published in print: Jul 1, 2024
Discussion open until: Sep 5, 2024

ASCE Technical Topics:

Authors

Affiliations

Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg, VA 24061 (corresponding author). ORCID: https://orcid.org/0000-0003-3050-3103. Email: [email protected]
Nina Stark, Ph.D., M.ASCE [email protected]
Engineering School of Sustainable Infrastructure and Environment, Univ. of Florida, Gainesville, FL 326113. Email: [email protected]
Celso Castro-Bolinaga, Ph.D., A.M.ASCE https://orcid.org/0000-0002-5990-8609 [email protected]
Dept. of Biological and Agricultural Engineering, North Carolina State Univ., Raleigh, NC 27606. ORCID: https://orcid.org/0000-0002-5990-8609. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share