Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-19T23:37:18.347Z Has data issue: false hasContentIssue false

What can long-term in situ monitoring data tell us about our coastlines?

Published online by Cambridge University Press:  06 January 2023

Masayuki Banno*
Affiliation:
Port and Airport Research Institute, Coastal and Estuarine Environment Department, Yokosuka 239-0826, Japan
*
Author for correspondence: Masayuki Banno, Email: banno-m@p.mpat.go.jp
Rights & Permissions [Opens in a new window]

Abstract

Long-term in situ monitoring of beach morphology is indispensable for capturing the processes of foreshore morphological changes, and thus many beach monitoring campaigns have been conducted globally. Here, we review the various foreshore beach processes attributable to cross-shore sediment transport, which have been elucidated through long-term beach monitoring. Historical in situ beach monitoring has revealed many daily–annual-scale cyclic foreshore beach morphological changes and shoreline changes; however, many shorter- and longer-term processes remain unresolved, for example, the short-term response to tidal fluctuations and the long-term response to sea level rise. The cost per area surveyed of state-of-the-art equipment will gradually decrease over time, and the accuracy, resolution, and volume of information obtained from the monitoring methods, which are still in the early stages of development, will improve as research progresses. Continued long-term monitoring and acquisition of previously unmeasured monitoring data through the development of monitoring methods are expected to help elucidate unresolved beach processes.

Topics structure

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

Impact statement

Beaches play an important role in disaster prevention by reducing the wave energy reaching land and protecting hinterlands from wave overtopping and other hazards. Using various methods, scientists have monitored beach morphology to investigate how and by what mechanism beach morphology changes. For example, long-term in situ monitoring data of beach morphology, obtained using methods such as leveling and satellite-based positioning, have provided insight into beach morphological processes on various time scales. Temporary erosion caused by storms and cyclic morphological changes caused by seasonal wave fluctuations have been well studied; however, certain processes, such as the effects of tidal and wave interactions on beach morphological change and the response to sea level rise, remain unresolved. High-quality multisite monitoring data acquired by new surveying methods, such as drones and LiDAR (light detection and ranging), will provide clues that will help elucidate these processes.

Introduction

Sandy beach morphology is changed continuously by waves, currents, and winds. The variability of beach morphology is one of the important processes in geophysics, and knowledge of the underlying processes is important for coastal disaster prevention. The foreshore, that is, the swash zone, is the area that changes most dynamically in response to the wave regime. Erosion of the foreshore can lead to large-scale shoreline retreat; therefore, foreshore morphology should be monitored carefully from the perspective of coastal management. To capture the processes underlying foreshore morphological change, in situ monitoring data of beach morphology are indispensable, and thus many beach monitoring campaigns have been performed worldwide.

Monitoring of beach morphology, including not only the foreshore but also the upper shoreface (i.e., the surf zone and shallower than the wave base), is mandatory in investigating foreshore morphological changes. Especially for long-term beach processes such as shoreline change due to sea level rise (SLR) and sediment transport caused by extreme storms once every few decades, the sediment transport and deposition processes between the lower and upper shoreface cannot be ignored (Anthony and Aagaard, Reference Anthony and Aagaard2020; Harley et al., Reference Harley, Masselink, Ruiz de Alegría-Arzaburu, Valiente and Scott2022). Most beach monitoring campaigns focus on the backshore to upper shoreface zone, while the lower shoreface is rarely monitored adequately.

Campaigns for monitoring beach morphology can be divided broadly into two categories. The first is short-term and high-frequency or high-resolution monitoring, which mainly focuses on the morphological changes caused by cross-shore sediment transport. We call such campaigns concentrated monitoring. The monitoring usually involves a single or a limited number of cross-shore transects, and the beach profile is measured to reveal the changes that occur over a period of a few days to a few weeks during storm and calm wave conditions. Short-term monitoring campaigns often target foreshore morphological changes that are substantial even over short periods. Generally, in such monitoring campaigns, beach profiles are measured using leveling, RTK-GNSS (real-time kinematic global navigation satellite system) (Harley et al., Reference Harley, Turner, Short and Ranasinghe2011a), and more recently, LiDAR techniques (Almeida et al., Reference Almeida, Masselink, Russell, Davidson, Poate, McCall, Blenkinsopp and Turner2013; O’Dea et al., Reference O’Dea, Brodie and Hartzell2019; Phillips et al., Reference Phillips, Blenkinsopp, Splinter, Harley and Turner2019).

The second category is long-term and low-frequency or low-resolution monitoring, which focuses on morphological changes caused by both cross-shore and longshore sediment transport. The monitoring usually involves obtaining a shoreline position of the entire coast and extensive three-dimensional bathymetry. Morphological changes due to beach rotation (Klein et al., Reference Klein, Benedet and Schumacher2002) and quantitative imbalance of longshore sediment transport around jetties, which can occur coastwide on monthly–multiyear scales, can be obtained. In a situation where external factors such as the wave regime or the sediment supply from a river could change on annual–decadal scales (Warrick et al., Reference Warrick, Bountry, East, Magirl, Randle, Gelfenbaum, Ritchie, Pess, Leung and Duda2015; Nguyen and Takewaka, Reference Nguyen and Takewaka2022), the object of monitoring might be to determine the long-term morphological changes due to cross-shore and longshore sediment transport. In some cases, such as in the Netherlands, a national strategy has promoted continued long-term annual monitoring of beach profiles on the national scale (i.e., the Jarkus program; e.g., Southgate, Reference Southgate2011). Satellite imagery and aerial photos are often used to observe long-term shoreline changes (Crowell et al., Reference Crowell, Leatherman and Buckley1991; Luijendijk et al., Reference Luijendijk, Hagenaars, Ranasinghe, Baart, Donchyts and Aarninkhof2018), and camera measurements such as Argus are also used to analyze short-term variations (Holman et al., Reference Holman, Sallenger, Lippmann and Haines1993; Harley and Kinsela, Reference Harley and Kinsela2022). Single-beam sonar and more recently narrow multibeam sonar mounted on boats are generally used to produce precise bathymetric maps over large areas (Barnard et al., Reference Barnard, Erikson and Kvitek2011), although the frequency of such monitoring is often as low as once per year.

On the basis of the above, it would be pertinent to ask whether there are any long-term high-frequency observations that could fill the gap between the two types of monitoring campaign. The answer to that question would be yes. Long-term and high-frequency monitoring is also important. Even though wave conditions might not change over long periods of time, beaches often show long-period fluctuations in morphology, suggesting that the nonlinearity of short-period beach responses could drive long-term beach changes. However, the nonlinearity is so small that limited short-term high-frequency observations cannot fully reveal the underlying processes. Here, we summarize the relation between the monitoring cost per unit area, and the accuracy and resolution (or volume of information) of each monitoring method (Figure 1). The latest technologies such as airborne LiDAR and multibeam sonar combined with GNSS can provide large volumes of high-quality data with high accuracy; however, they are unsuitable for long-term high-frequency monitoring campaigns owing to their high operational costs. Conversely, satellite imagery and Argus photos are low cost and are suitable for long-term observations. However, the information obtained from such images is often limited to shoreline position, and the temporal or spatial resolution is not always adequate. Positioned between the two in Figure 1, RTK-GNSS, leveling, and single-beam sonar attached to a jet ski are techniques unsuitable for monitoring of large areas, but they can be used to obtain high-accuracy data over a limited area. These methods can be used for long-term high-frequency observations if certain costs are borne. Recently, automatic observations of the zone from dunes to the foreshore have been obtained by fixed LiDAR and photogrammetry using UAVs (unmanned aerial vehicles) (Turner et al., Reference Turner, Harley, Short, Simmons, Bracs, Phillips and Splinter2016), which represents a recent major development in beach monitoring with reduced costs and improved accuracy, although challenges remain with regard to surveying underwater areas. There is no doubt that long-term, high-frequency, and high-resolution in situ monitoring of beach morphology can be realized only at great cost. Consequently, long-term beach profile monitoring campaigns are very limited worldwide owing to the constraints of human and financial resources.

Figure 1. Relationship between cost and information of beach survey methods.

Representative beach monitoring campaigns that have obtained relatively high-frequency (i.e., more than bimonthly) beach profiles for more than 15 years have been conducted at Narrabeen, Australia (Short and Trenbanis, Reference Short and Trenbanis2004; Harley et al., Reference Harley, Turner, Short and Ranasinghe2011a; Turner et al., Reference Turner, Harley and Drummond2016), Moruya, Australia (Thom and Hall, Reference Thom and Hall1991; McLean and Shen, Reference McLean and Shen2006; Tamura et al., Reference Tamura, Oliver, Cunningham and Woodroffe2019), Hasaki, Japan (Banno et al., Reference Banno, Nakamura, Kosako, Nakagawa, Yanagishima and Kuriyama2020), Omotehama, Japan (Kato et al., Reference Kato, Okabe and Sawahara2013), Duck, USA (Lippmann and Holman, Reference Lippmann and Holman1990; Larson and Kraus, Reference Larson and Kraus1994; Nicholls et al., Reference Nicholls, Birkemeier and Lee1998), Torrry Pines, USA (Ludka et al., Reference Ludka, Guza, O’Reilly, Merrifield, Flick, Bak, Hesser, Bucciarelli, Olfe, Woodward, Boyd, Smith, Okihiro, Grenzeback, Parry and Boyd2019), Ocean Beach, USA (Yates et al., Reference Yates, Guza, O’Reilly, Hansen and Barnard2011; Splinter et al., Reference Splinter, Turner, Davidson, Barnard, Castelle and Oltman-Shay2014), Truc Vert, France (Castelle et al., Reference Castelle, Bujan, Marieu and Ferreira2020), and Porsmilin, France (Bertin et al., Reference Bertin, Floc’h, Le Dantec, Jaud, Cancouët, Franzetti, Cuq, Prunier, Ammann, Augereau, Lamarche, Belleney, Rouan, David, Deschamps, Delacourt and Suanez2022). In these long-term in situ monitoring campaigns, beach profiles along one or several limited transects are measured using techniques such as GNSS. Although the depths and elevations of the offshore and landward boundaries vary among campaigns, most of the monitoring campaigns at least cover the foreshore area. Long-term campaigns that have been conducted bimonthly (or more frequently), which include various beach morphological changes such as abrupt large erosion during periods of high waves and gradual accretion during periods of low waves, provide very valuable information. Based on such monitoring data, numerical models have been developed to reproduce the morphological changes of the foreshore, and associated beach morphological change processes have been investigated. In this paper, we review the various foreshore beach processes that operate on multiple time scales due to cross-shore sediment transport that have been elucidated through long-term beach monitoring, and we discuss the scientific knowledge that we can expect to obtain in the future through further development of techniques for monitoring beach morphology and continued long-term monitoring campaigns.

In studies on foreshore morphological changes, cross-shore position with a specific elevation is often used as a proxy for beach morphology (Boak and Turner, Reference Boak and Turner2005). It is called the shoreline position and is defined with a specific elevation, such as the height of mean high-water springs (MHWS), which retreats during erosion and advances during accretion. When the morphological change of an entire foreshore is targeted, multiple reference heights are used for shoreline positions. Thus, shoreline changes and foreshore morphological changes are almost synonymous in terms of the definition.

Cyclic foreshore beach processes revealed by observations

The usefulness of long-term high-frequency monitoring data is highlighted in periodicity analyses of short-term (less than 1 year) foreshore morphological changes. Spectral analysis revealed cyclic shoreline changes with 1-year and 6-month cycles at Hasaki, Japan (Banno and Kuriyama, Reference Banno and Kuriyama2020), which are caused by cross-shore sediment transport related to seasonal fluctuations in waves (Eichentopf et al., Reference Eichentopf, Alsina, Christou, Kuriyama and Karunarathna2020). Conversely, data recorded at Narrabeen, where high-frequency morphological changes have also been monitored, showed no clear seasonal variation in the mean shoreline position (Lazarus et al., Reference Lazarus, Harley, Blenkinsopp and Turner2019). This is probably attributable to the fact that the incident wave energy at Narrabeen does not show clear cyclic variation, such as a 1-year cycle, and because of the relatively large influence of morphological changes induced by longshore sediment transport such as beach rotation (Harley et al., Reference Harley, Turner, Short and Ranasinghe2011b; Harley et al., Reference Harley, Turner and Short2015) associated with seasonal changes in the direction of incoming waves. In the case of substantial impacts such as longshore sediment transport and sediment budget changes, the importance of long-term high-frequency monitoring data increases because large volumes of data are required for statistical extraction of the morphological changes caused by cross-shore sediment transport.

A recent study using a large volume of monitoring data revealed a foreshore beach process previously masked by variations in other factors. The process is affected by tidal fluctuations. The study revealed that a large tidal range during spring tides and king tides (Flick, Reference Flick2016) makes erosion of the upper swash zone more likely, even within the same wave regime (Banno and Kuriyama, Reference Banno and Kuriyama2020). Several monitoring studies showed that beach elevation of a few meters above the mean water level was minimal a few days after the spring tide (LaFond, Reference LaFond1939; Aubrey et al., Reference Aubrey, Inman and Nordstrom1976; Clarke et al., Reference Clarke, Eliot and Frew1984). However, the effect of the primary underlying beach process could not be distinguished from other effects, such as waves, because of limited available data and thus full explanation of the beach process was not proposed. In relation to this process induced by tidal fluctuations, sediment transport and the morphological changes even between a single tide, that is, between rising and falling tides, might be affected by infiltration and exfiltration of water on the beach face (Duncan, Reference Duncan1964; Clarke and Eliot, Reference Clarke and Eliot1987; Butt et al., Reference Butt, Russell and Turner2001; Masselink and Li, Reference Masselink and Li2001; Coco et al., Reference Coco, Burnet, Werner and Elgar2004). However, beach processes on the hourly scale have not yet been clarified because of lack of long-term morphological change data monitored at intervals of less than 1 day, for example, 6 h, which is the time scale of ebb and flood tides.

Daily–annual-scale foreshore beach processes revealed by observations

For time scales of less than 1 year, as shown by the seasonal cycle in the spectra of beach monitoring data, it has long been known that a shoreline will advance and retreat seasonally owing to seasonal wave fluctuations (e.g., Shepard, Reference Shepard1950). When waves are calm, sediment is transported onshore by sheet flow and accreted over a period of days. Conversely, when waves are large during storms, erosion can occur in just a few days (sometimes only a few hours) with sediment transported offshore via suspension. On many beaches, wave characteristics vary seasonally, and the foreshore accretes with a well-developed berm during seasons with frequent periods of calm waves and retreats with a relatively gradual slope during seasons with frequent periods of high waves.

Equilibrium-based shoreline change models are used widely to estimate short–long-term accretion and erosion, such as that over daily–multiyear periods including seasonal variation (Miller and Dean, Reference Miller and Dean2004; Yates et al., Reference Yates, Guza and O’Reilly2009, Reference Yates, Guza, O’Reilly, Hansen and Barnard2011; Davidson et al., Reference Davidson, Splinter and Turner2013; Turki et al., Reference Turki, Medina, Coco and Gonzalez2013; Castelle et al., Reference Castelle, Marieu, Bujan, Ferreira, Parisot, Capo, Sénéchal and Chouzenoux2014; Splinter et al., Reference Splinter, Turner, Davidson, Barnard, Castelle and Oltman-Shay2014; Banno et al., Reference Banno, Kuriyama and Hashimoto2015; Jara et al., Reference Jara, González and Medina2015; Dean and Houston, Reference Dean and Houston2016; Vitousek et al., Reference Vitousek, Barnard, Limber, Erikson and Cole2017; Lemos et al., Reference Lemos, Floc’h, Yates, Le Dantec, Marieu, Hamon, Cuq, Suanez and Delacourt2018; Chataigner et al., Reference Chataigner, Yates, Le Dantec, Suanez, Floch, Bouvard, Leary, Petton and Cailler2020; D’Anna et al., Reference D’Anna, Idier, Castelle, Le Cozannet, Rohmer and Robinet2020, Reference D’Anna, Castelle, Idier, Rohmer, Le Cozannet, Thieblemont and Bricheno2021a,Reference D’Anna, Idier, Castelle, Vitousek and Le Cozannetb; Montaño et al., Reference Montaño, Coco, Antolínez, Beuzen, Bryan, Cagigal, Castelle, Davidson, Goldstein, Ibaceta, Idier, Ludka, Masoud-Ansari, Méndez, Murray, Plant, Ratliff, Robinet, Rueda, Sénéchal, Simmons, Splinter, Stephens, Townend, Vitousek and Vos2020). A beach will approach a theoretical equilibrium profile, on which no further morphological change is caused, when subjected to the same wave regime for an extended period. Equilibrium-based shoreline change models are based on the concept of the equilibrium profile determined by waves. In such numerical models, shoreline change is determined by the imbalance of the actual position of the shoreline relative to the equilibrium shoreline position. The advantage of such models is that beaches where substantial erosion has already occurred will not experience further notable erosion (Eichentopf et al., Reference Eichentopf, Alsina, Christou, Kuriyama and Karunarathna2020). Consequently, the long-term robustness of equilibrium-based shoreline change models is high because simulation results tend not to diverge unrealistically but fluctuate around the equilibrium shoreline position; thus, such models are often used for long-term hindcasts and forecasts.

Equilibrium-based shoreline change models have been used to reproduce shoreline changes over periods of several years relatively well (e.g., Yates et al., Reference Yates, Guza and O’Reilly2009; Splinter et al., Reference Splinter, Turner, Davidson, Barnard, Castelle and Oltman-Shay2014). The temporal resolution of the estimated shoreline changes depends largely on the monitoring data used to calibrate the model; therefore, in many cases, such simulations are limited to reproduction of only seasonal change. When using high-frequency beach monitoring data for calibration, such as the daily data obtained at Hasaki and the biweekly data acquired at Narrabeen, short-term erosion over a period of a few days can be estimated (e.g., Davidson et al., Reference Davidson, Splinter and Turner2013; Splinter et al., Reference Splinter, Turner, Davidson, Barnard, Castelle and Oltman-Shay2014; Banno et al., Reference Banno, Kuriyama and Hashimoto2015).

Even when using high-frequency shoreline change data for calibration, the estimated temporal resolution might not always be high depending on the tuning of the model parameters (Montaño et al., Reference Montaño, Coco, Antolínez, Beuzen, Bryan, Cagigal, Castelle, Davidson, Goldstein, Ibaceta, Idier, Ludka, Masoud-Ansari, Méndez, Murray, Plant, Ratliff, Robinet, Rueda, Sénéchal, Simmons, Splinter, Stephens, Townend, Vitousek and Vos2020). Parameter tuning can be classified into two methods: minimizing the error with respect to the observed shoreline position and minimizing the error with respect to the amount of shoreline change. In most studies, the former method is adopted because it is most likely to reduce the overall error; however, it tends to reproduce trends without reproducing fine-scale high-resolution temporal variation. The latter method is more likely to reproduce short-term variations, but it does not account for shoreline position errors; consequently, the resulting output of estimated shoreline positions might differ substantially from the measured shoreline positions. We have not found any studies that even discuss the above calibration methods, even though selection of the method adopted depends on the purpose for which the model will be used.

Annual–decadal-scale foreshore beach processes revealed by observations

Beach morphological changes over long time scales are often discussed in the context of global climate change. Using long-term beach monitoring data, coastal responses to atmospheric and oceanic variations such as the El Niño–Southern Oscillation (ENSO), which varies on the scale of a few years, have been investigated (Barnard et al., Reference Barnard, Short, Harley, Splinter, Vitousek, Turner, Allan, Banno, Bryan, Doria, Hansen, Kato, Kuriyama, Randall-Goodwin, Ruggiero, Walker and Heathfield2015). Beach morphological monitoring data along 48 coasts of the Pacific Rim revealed different responses to the impact of ENSO in different locations. For example, erosion is greater along the California coast (e.g., Ocean Beach) and Japanese coast (e.g., Hasaki) facing the Pacific Ocean in winter when El Niño events occur, while erosion is greater in Australia (e.g., Narrabeen) in winter when La Niña events occur. It has been suggested that the impact on erosion is caused by waves and sea level variations affected by ENSO. Relationships between atmospheric and oceanic variations and waves have been indicated by the high correlation of atmospheric pressure patterns and climate indices with the variations of waves (Castelle et al., Reference Castelle, Dodet, Masselink and Scott2017; Kishimoto et al., Reference Kishimoto, Shimura, Mori and Mase2017), which consequently also influence shoreline variations (Kuriyama et al., Reference Kuriyama, Banno and Suzuki2012; Robinet et al., Reference Robinet, Castelle, Idier, Cozannet, Déqué and Charles2016). Studies at Hasaki (Kuriyama et al., Reference Kuriyama, Banno and Suzuki2012) and Truc Vert (Robinet et al., Reference Robinet, Castelle, Idier, Cozannet, Déqué and Charles2016) showed that 45% and 70% of long-term shoreline variability can be explained by teleconnections of large-scale atmospheric and oceanic variability, respectively. Another study reported that the 18.6-year nodal tidal cycle causes shoreline retreat (Gratiot et al., Reference Gratiot, Anthony, Gardel, Gaucherel, Proisy and Wells2008); however, the response of beach morphological change to long-term sea level change with such periodicity has not yet been fully investigated using a sufficient volume of long-term monitoring data.

The Bruun rule (Bruun, Reference Bruun1962), which determines the probable amount of future shoreline retreat associated with a shift of the equilibrium beach profile with SLR, is widely used. In terms of the concept, a rising sea level is expected to cause foreshore erosion and offshore sediment transport near the foreshore. It has been highlighted that this method enforces many assumptions (Cooper and Pilkey, Reference Cooper and Pilkey2004) and is susceptible to the setting of parameters such as closure depth (Udo et al., Reference Udo, Ranasinghe and Takeda2020) and that it might not always match actual morphological changes (Ranasinghe et al., Reference Ranasinghe, Callaghan and Stive2012). In historical beach monitoring data, the effect of SLR per unit time on cross-shore sediment transport is very small in comparison with the morphological changes caused by wave variations and longshore sediment transport. Therefore, the validity of the Bruun rule and the details of actual coastal response to SLR remain unknown. The effects of SLR on beach morphological changes are currently being investigated in laboratory experiments that simulate SLR as close as possible to actual scale, while controlling all other external forces (Atkinson et al., Reference Atkinson, Baldock, Birrien, Callaghan, Nielsen, Beuzen, Turner, Blenkinsopp and Ranasinghe2018; Atkinson and Baldock, Reference Atkinson and Baldock2020), and in analysis of shoreline changes exposed to hypothetical SLR where land subsidence is important (Nguyen and Takewaka, Reference Nguyen and Takewaka2020). The current solution is to couple the Bruun rule with an equilibrium-based shoreline change model. Long-term shoreline predictions have been simulated using this approach by incorporating the shoreline retreat predicted by the Bruun rule (Vitousek et al., Reference Vitousek, Barnard, Limber, Erikson and Cole2017; D’Anna et al., Reference D’Anna, Idier, Castelle, Le Cozannet, Rohmer and Robinet2020, Reference D’Anna, Castelle, Idier, Rohmer, Le Cozannet, Thieblemont and Bricheno2021a), and by introducing the Bruun rule directly into the term representing the equilibrium shoreline position in the model (Banno et al., Reference Banno, Kuriyama and Hashimoto2015; D’Anna et al., Reference D’Anna, Idier, Castelle, Vitousek and Le Cozannet2021b).

Long-term shoreline changes are also expected to occur in response to long-term changes in wave climate. Therefore, it is important to determine whether we reproduce long-term shoreline change on annual–decadal scales using equilibrium-based shoreline change models. A recent study using multiple equilibrium-based shoreline change models hindcasted shoreline change over 15 years and predicted blind shoreline change over 3 years for Tairua beach, New Zealand (Montaño et al., Reference Montaño, Coco, Antolínez, Beuzen, Bryan, Cagigal, Castelle, Davidson, Goldstein, Ibaceta, Idier, Ludka, Masoud-Ansari, Méndez, Murray, Plant, Ratliff, Robinet, Rueda, Sénéchal, Simmons, Splinter, Stephens, Townend, Vitousek and Vos2020). Long-term hindcasts (past 22 years) and scenario projections (future 88 years) of shoreline change have been calculated for Hasaki, Japan (Banno and Kuriyama, Reference Banno and Kuriyama2014; Banno et al., Reference Banno, Kuriyama and Hashimoto2015). In both studies, the observed long-term shoreline variability (type of long-term trend) could not be reproduced adequately, leading to deterioration of the overall reproduction accuracy (i.e., R 2 and RMSE (root-mean-square error)). Whether this observed long-term shoreline variation is caused by wave climate change or by other processes remains a matter of debate, but current shoreline change models are limited in terms of their skill in reproducing and predicting long-term shoreline variation. In forecasting applications, equilibrium-based shoreline change models use unique parameters determined from observational data. Thus, the simulated shoreline position tends to be near the mean shoreline position, and reproducibility and prediction skill are diminished when forced with nonstationary wave conditions over time (Ibaceta et al., Reference Ibaceta, Splinter, Harley and Turner2020), such as wave climate changes due to global warming. It has also been suggested that beach response could be affected in the long-term by substantial erosion due to extreme storms (Kuriyama and Yanagishima, Reference Kuriyama and Yanagishima2018). Although foreshore morphological change is also affected by the effects of wave breaking by longshore bars (Kuriyama and Banno, Reference Kuriyama and Banno2016), which move and develop on a multiyear scale (Ruessink et al., Reference Ruessink, Wijnberg, Holman, Kuriyama and van Enckevort2003, Reference Ruessink, Kuriyama, Reniers, Roelvink and Walstra2007), these effects are not introduced into equilibrium-based shoreline change models. In practice, there is a dilemma regarding incorporation of the effects of longshore bars into shoreline change models because prediction of longshore bars has yet to be fully accomplished. Although process-based models such as XBeach (Roelvink et al., Reference Roelvink, Reniers, van Dongeren, van Thiel de Vries, McCall and Lescinski2009) and DELFT3D (Lesser et al., Reference Lesser, Roelvink, van Kester and Stelling2004), which include calculations of wave deformation and dissipation, are effective in simulating detailed coastal processes, producing long-term forecasts over periods of more than a few months remains challenging (Hanson et al., Reference Hanson, Aarninkhof, Capobianco, Jim’enez, Larson, Nicholls, Plant, Southgate, Steetzel, Stive and de Vriend2003). Nevertheless, application of process-based models to long-term forecasting has been studied (Davidson, Reference Davidson2021). An approach to long-term forecasting in response to wave fluctuations and SLR, which combines a process-based model and a probabilistic method, has also been studied (Ranasinghe et al., Reference Ranasinghe, Callaghan and Stive2012; Dastgheib et al., Reference Dastgheib, Martinez, Udo and Ranasinghe2022). Currently, a reasonable simple solution is to account for the uncertainties in hindcasts and forecasts by Monte Carlo simulations (Banno and Kuriyama, Reference Banno and Kuriyama2014) or by ensembles of multiple models (Montaño et al., Reference Montaño, Coco, Antolínez, Beuzen, Bryan, Cagigal, Castelle, Davidson, Goldstein, Ibaceta, Idier, Ludka, Masoud-Ansari, Méndez, Murray, Plant, Ratliff, Robinet, Rueda, Sénéchal, Simmons, Splinter, Stephens, Townend, Vitousek and Vos2020). Original wave monitoring and prediction data also contain uncertainty, and the uncertainty of wave data should be considered, for example, using Monte Carlo simulations (Chataigner et al., Reference Chataigner, Yates, Le Dantec, Harley, Splinter and Goutal2022) or ensembles of wave conditions (Vitousek et al., Reference Vitousek, Cagigal, Montaño, Rueda, Mendez, Coco and Barnard2021).

As described above, reasonably adequate knowledge of foreshore morphological changes on time scales of a few days to a few years has been obtained from long-term beach monitoring data. However, we need further monitoring data to investigate foreshore morphological changes on time scales shorter than a day and longer than several years.

Monitoring future

The development of monitoring methods will provide new insights into beach morphological changes and processes. The cost per surveyed area of state-of-the-art equipment such as LiDAR will gradually decrease through technological innovation and more efficient manufacturing, and the accuracy, resolution, and volume of information obtained from monitoring method using satellite imagery and Argus photos will improve as research progresses. In recent years, miniaturized LiDAR scanners that can be mounted on UAVs have reduced the cost of using large airborne platforms. Efficient collection of beach morphological data by combining sonar with autonomous underwater vehicles and other remotely operated vehicles is also expected to be adopted for long-term monitoring in the future. Additionally, estimation of bathymetry using satellite techniques, such as synthetic aperture radar imagery (Pleskachevsky et al., Reference Pleskachevsky, Lehner, Heege and Mott2011; Capo et al., Reference Capo, Lubac, Marieu, Robinet, Bru and Bonneton2014) and inverse estimation of bathymetry from video recordings of waves (Matsuba and Sato, Reference Matsuba and Sato2018), have recently been investigated, and the resolution and accuracy of such approaches will improve over coming years. It is expected that long-term ultra-high-frequency monitoring data will become more readily available in the future following these developments, and we anticipate that easier acquisition of such data will allow enhanced monitoring of beach morphology.

For example, if beach profiles were observed over a long period at intervals of several hours using automatic instruments such as LiDAR with a green laser (e.g., Pastol, Reference Pastol2011), it might be possible to quantify the effects of tides on beach morphological changes during flood and ebb tides, as suggested by Phillips et al. (Reference Phillips, Blenkinsopp, Splinter, Harley and Turner2019) and Banno and Kuriyama (Reference Banno and Kuriyama2020). Moreover, XBeach (Roelvink et al., Reference Roelvink, Reniers, van Dongeren, van Thiel de Vries, McCall and Lescinski2009), which is used mainly to simulate short-term morphological changes, could be used to consider the effects of wave run-up infiltration and groundwater effluent on the morphological changes of sandy beaches, expanding on previous work that used XBeach to study similar effects on gravel beaches (XBeach-G; McCall et al., Reference McCall, Masselink, Poate, Roelvink and Almeida2015). Elucidation of detailed processes through high-frequency beach monitoring could greatly advance these modeling efforts, potentially improving not only the reproducibility of short-term morphological changes, but also the reproducibility of long-term morphological changes that represent the integration of the former.

Considering the risk of coastal disasters, temporary erosion following a single storm event should also be estimated in addition to estimation of seasonal shoreline variations. Therefore, it is necessary to increase the temporal resolution of shoreline changes estimated by models using monitoring data with the highest possible temporal resolution. Moreover, if long-term observation data with high temporal resolution were available for many beaches, the practicality of research on shoreline prediction would be greatly enhanced. One of the future developments of equilibrium-based shoreline change models, through application with as much beach monitoring data as possible, will be to accumulate knowledge to permit generalization of the model parameters that are currently estimated site-specifically for each beach where monitoring data are available. For example, the relationship between sediment grain size and model parameters might also lead to generalization of model parameters (Yates et al., Reference Yates, Guza, O’Reilly, Hansen and Barnard2011; Splinter et al., Reference Splinter, Turner, Davidson, Barnard, Castelle and Oltman-Shay2014), which would allow prediction of short-term erosion on beaches where monitoring data are not necessarily abundant.

Together with improved data quality (increase in data acquisition frequency, resolution, and volume of information) and quantity (increase in the number of beaches monitored), it is also essential that long-term morphological data continue to be obtained. We need long-term monitoring data of beach morphological change on beaches where SLR is significant to distinguish the effects of SLR on morphological changes from complex beach morphological changes caused by various other factors such as waves. We will experience a clearer and more extreme beach response to ongoing dynamic SLR and changes in wave climate in coming decades. With the monitoring data expected to be obtained in the future, we will be able to rapidly advance verification of the Bruun rule, which has currently been discussed only as a concept, and elucidation of long-term factors and processes of beach fluctuation that are not yet fully understood.

Conclusions

This paper broadly summarized the findings of foreshore morphological change through historical long-term and high-frequency monitoring of beach profiles. Morphological changes on the time scale of a few days to a few years can now be reproduced by equilibrium-based shoreline change models using waves as the driving force, which has been achieved thorough the availability of monitoring data and extensive study in recent years. However, because of lack of beach monitoring, we still do not have sufficient knowledge of the morphological changes that occur over periods shorter than a day, which might be affected by tides, and those that occur over periods of a decade, which are closely related to SLR and wave climate changes. Recently, data acquired during several long-term beach monitoring campaigns, for example, that at Narrabeen (Turner et al., Reference Turner, Harley, Short, Simmons, Bracs, Phillips and Splinter2016), have been released to the public under the concept of open data (Ludka et al., Reference Ludka, Guza, O’Reilly, Merrifield, Flick, Bak, Hesser, Bucciarelli, Olfe, Woodward, Boyd, Smith, Okihiro, Grenzeback, Parry and Boyd2019; Castelle et al., Reference Castelle, Bujan, Marieu and Ferreira2020; Bertin et al., Reference Bertin, Floc’h, Le Dantec, Jaud, Cancouët, Franzetti, Cuq, Prunier, Ammann, Augereau, Lamarche, Belleney, Rouan, David, Deschamps, Delacourt and Suanez2022). It is expected that more and more beach monitoring data will become available in the future, and that the use of open data will greatly advance our understanding of foreshore beach processes. Long-term and high-frequency monitoring of beach morphological changes is also expected to be conducted for more beaches following the development of monitoring methods. As larger volumes of monitoring data become available for use in future studies, we expect not only improvement of physics-based models, but also marked development of statistical models, such as deep neural networks, which have progressed remarkably in recent years (Goldstein et al., Reference Goldstein, Coco and Plant2019). Through various studies using monitoring data, we expect to gain more comprehensive understanding of beach processes on various time scales, resulting in development of prediction technology for beach morphological changes with high accuracy and high certainty that will help in the effort to maintain long-term beach stability and durability.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/cft.2022.9.

Acknowledgments

We are thankful to Dr. Yoshiaki Kuriyama and Dr. Quang Hao Nguyen for their comments on an early version of this work. We thank James Buxton, from Edanz for editing a draft of this manuscript.

Author contributions

M.B. wrote this review article and prepared the figure.

Financial support

This work was supported by JSPS KAKENHI (Grant Numbers 21 K04285, 22F22057, and 22H01595) and the SENTAN Program (Grant Number JPMXD0722678534) of the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan.

Competing interest

The author declares none.

References

Almeida, LP, Masselink, G, Russell, P, Davidson, M, Poate, T, McCall, R, Blenkinsopp, C and Turner, I (2013) Observations of the swash zone on a gravel beach during a storm using a laser-scanner (Lidar). Journal of Coastal Research 65(10065), 636641.CrossRefGoogle Scholar
Anthony, EJ and Aagaard, T (2020) The lower shoreface: Morphodynamics and sediment connectivity with the upper shoreface and beach. Earth-Science Reviews 210, 103334.CrossRefGoogle Scholar
Atkinson, AL and Baldock, TE (2020) Laboratory investigation of nourishment options to mitigate sea level rise induced erosion. Coastal Engineering 161, 103769.CrossRefGoogle Scholar
Atkinson, AL, Baldock, TE, Birrien, F, Callaghan, DP, Nielsen, P, Beuzen, T, Turner, IL, Blenkinsopp, CE and Ranasinghe, R (2018) Laboratory investigation of the Bruun rule and beach response to sea level rise. Coastal Engineering 136, 183202.CrossRefGoogle Scholar
Aubrey, DG, Inman, DL and Nordstrom, CE (1976) Beach profiles at Torrey Pines, California. Coastal Engineering Proceedings 15, 12971311.Google Scholar
Banno, M and Kuriyama, Y (2014) Prediction of future shoreline change with sea-level rise and wave climate change at Hasaki, Japan. Coastal Engineering Proceedings, 34, sediment.56.CrossRefGoogle Scholar
Banno, M and Kuriyama, Y (2020) Supermoon drives beach morphological changes in the swash zone. Geophysical Research Letters 47, e2020GL089745.CrossRefGoogle Scholar
Banno, M, Kuriyama, Y and Hashimoto, N (2015) Equilibrium-based foreshore beach profile change model for long-term data. In The Proceedings of the Coastal Sediments 2015. Singapore: World Scientific. https://doi.org/10.1142/9789814689977_0235.Google Scholar
Banno, M, Nakamura, S, Kosako, T, Nakagawa, Y, Yanagishima, S and Kuriyama, Y (2020) Long-term observations of beach variability at Hasaki, Japan. Journal of Marine Science and Engineering 8, 871.CrossRefGoogle Scholar
Barnard, PL, Erikson, LH and Kvitek, RG (2011) Small-scale sediment transport patterns and bedform morphodynamics: New insights from high-resolution multibeam bathymetry. Geo-Marine Letters 31, 227236.CrossRefGoogle Scholar
Barnard, PL, Short, AD, Harley, MD, Splinter, KD, Vitousek, S, Turner, IL, Allan, J, Banno, M, Bryan, KR, Doria, A, Hansen, JE, Kato, S, Kuriyama, Y, Randall-Goodwin, E, Ruggiero, P, Walker, IJ and Heathfield, DK (2015) Coastal vulnerability across the Pacific dominated by El Niño/southern oscillation. Nature Geoscience 8, 801807.CrossRefGoogle Scholar
Bertin, S, Floc’h, F, Le Dantec, N, Jaud, M, Cancouët, R, Franzetti, M, Cuq, V, Prunier, C, Ammann, J, Augereau, E, Lamarche, S, Belleney, D, Rouan, M, David, L, Deschamps, A, Delacourt, C and Suanez, S (2022) A long-term dataset of topography and nearshore bathymetry at the macrotidal pocket beach of Porsmilin, France. Scientific Data 9, 79.CrossRefGoogle ScholarPubMed
Boak, EH and Turner, IL (2005) Shoreline definition and detection: A review. Journal of Coastal Research 214, 688703.CrossRefGoogle Scholar
Bruun, P (1962) Sea-level rise as a cause of shore erosion. Journal of the Waterways and Harbors Division 88, 117130.CrossRefGoogle Scholar
Butt, T, Russell, P and Turner, I (2001) The influence of swash infiltration-exfiltration on beach face sediment transport: Onshore or offshore? Coastal Engineering 42(1), 3552.CrossRefGoogle Scholar
Capo, S, Lubac, B, Marieu, V, Robinet, A, Bru, D and Bonneton, P (2014) Assessment of the decadal morphodynamic evolution of a mixed energy inlet using ocean color remote sensing. Ocean Dynamics 64, 15171530.CrossRefGoogle Scholar
Castelle, B, Bujan, S, Marieu, V and Ferreira, S (2020) 16 years of topographic surveys of rip-channelled high-energy meso-macrotidal sandy beach. Scientific Data 7, 410.CrossRefGoogle ScholarPubMed
Castelle, B, Dodet, G, Masselink, G and Scott, T (2017) A new climate index controlling winter wave activity along the Atlantic coast of Europe: The West Europe pressure anomaly. Geophysical Research Letters 44(3), 13841392.CrossRefGoogle Scholar
Castelle, B, Marieu, V, Bujan, S, Ferreira, S, Parisot, JP, Capo, S, Sénéchal, N and Chouzenoux, T (2014) Equilibrium shoreline modelling of a high-energy meso-macrotidal multiple-barred beach. Marine Geology 347, 8594.CrossRefGoogle Scholar
Chataigner, T, Yates, ML, Le Dantec, N, Harley, MD, Splinter, KD and Goutal, N (2022) Sensitivity of a one-line longshore shoreline change model to the mean wave direction. Coastal Engineering 172, 104025.CrossRefGoogle Scholar
Chataigner, T, Yates, M, Le Dantec, N, Suanez, S, Floch, F, Bouvard, G, Leary, M, Petton, C, and Cailler, N (2020) Equilibrium modeling of current and future beach evolution: Vougot beach, France. Coastal Engineering Proceedings, 36v, sediment.17.CrossRefGoogle Scholar
Clarke, DJ and Eliot, IG (1987) Ground water-level changes in a coastal dune, sea-level fluctuations and shoreline movement on a sandy beach. Marine Geology 77, 319326.CrossRefGoogle Scholar
Clarke, DJ, Eliot, IG and Frew, JR (1984) Variation in subaerial beach sediment volume on a small sandy beach over a monthly lunar tidal cycle. Marine Geology 58, 319344.CrossRefGoogle Scholar
Coco, G, Burnet, TK, Werner, BT and Elgar, S (2004) The role of tides in beach cusp development. Journal of Geophysical Research: Oceans 109, C04011.CrossRefGoogle Scholar
Cooper, JAG and Pilkey, OH (2004) Sea-level rise and shoreline retreat: Time to abandon the Bruun rule. Global and Planetary Change 43, 157171.CrossRefGoogle Scholar
Crowell, M, Leatherman, SP and Buckley, MK (1991) Historical shoreline change: Error analysis and mapping accuracy. Journal of Coastal Research 7(3), 839852.Google Scholar
D’Anna, M, Castelle, B, Idier, D, Rohmer, J, Le Cozannet, G, Thieblemont, R and Bricheno, L (2021a) Uncertainties in shoreline projections to 2100 at Truc Vert beach (France): Role of sea-level rise and equilibrium model assumptions. Journal of Geophysical Research: Earth Surface 126, e2021JF006160.Google Scholar
D’Anna, M, Idier, D, Castelle, B, Le Cozannet, G, Rohmer, J and Robinet, A (2020) Impact of model free parameters and sea-level rise uncertainties on 20-years shoreline hindcast: The case of Truc Vert beach (SW France). Earth Surface Processes and Landform 45, 18951907.CrossRefGoogle Scholar
D’Anna, M, Idier, D, Castelle, B, Vitousek, S and Le Cozannet, G (2021b) Reinterpreting the Bruun rule in the context of equilibrium shoreline models. Journal of Marine Science and Engineering 9(9), 974.CrossRefGoogle Scholar
Dastgheib, A, Martinez, C, Udo, K and Ranasinghe, R (2022) Climate change driven shoreline change at Hasaki Beach Japan: A novel application of the probabilistic coastline recession (PCR) model. Coastal Engineering 172, 104079.CrossRefGoogle Scholar
Davidson, M (2021) Forecasting coastal evolution on time-scales of days to decades. Coastal Engineering 168, 103928.CrossRefGoogle Scholar
Davidson, MA, Splinter, KD and Turner, IL (2013) A simple equilibrium model for predicting shoreline change. Coastal Engineering 73, 191202.CrossRefGoogle Scholar
Dean, RG and Houston, JR (2016) Determining shoreline response to sea level rise. Coastal Engineering 114, 18.CrossRefGoogle Scholar
Duncan, US (1964) The effect of water table and tide cycle on swash-backwash sediment distribution and beach profile development. Marine Geology 2, 117130.CrossRefGoogle Scholar
Eichentopf, S, Alsina, JM, Christou, M, Kuriyama, Y and Karunarathna, H (2020) Storm sequencing and beach profile variability at Hasaki, Japan. Marine Geology 424, 106153.CrossRefGoogle Scholar
Flick, RE (2016) California tides, sea level, and waves—Winter 2015–2016. Shore and Beach 84, 2530.Google Scholar
Goldstein, E, Coco, G and Plant, NG (2019) A review of machine learning applications to coastal sediment transport and morphodynamics. Earth-Science Reviews 194, 97108.CrossRefGoogle Scholar
Gratiot, N, Anthony, EJ, Gardel, A, Gaucherel, C, Proisy, C and Wells, JT (2008) Significant contribution of the 18.6 year tidal cycle to regional coastal changes. Nature Geoscience 1(3), 169172.CrossRefGoogle Scholar
Hanson, H, Aarninkhof, S, Capobianco, M, Jim’enez, JA, Larson, M, Nicholls, RJ, Plant, NG, Southgate, HN, Steetzel, HJ, Stive, MJF and de Vriend, HJ (2003) Modelling of coastal evolution on yearly to decadal time scales. Journal of Coastal Research 19, 790811.Google Scholar
Harley, MD and Kinsela, MA (2022) CoastSnap: A global citizen science program to monitor changing coastlines. Continental Shelf Research 245, 104796.CrossRefGoogle Scholar
Harley, MD, Masselink, G, Ruiz de Alegría-Arzaburu, A, Valiente, NG and Scott, T (2022) Single extreme storm sequence can offset decades of shoreline retreat projected to result from sea-level rise. Communications Earth & Environment 3, 112.CrossRefGoogle Scholar
Harley, MD, Turner, IL and Short, AD (2015) New insights into embayed beach rotation: The importance of wave exposure and cross-shore processes. Journal of Geophysical Research Earth Surface 120, 14701484.CrossRefGoogle Scholar
Harley, MD, Turner, IL, Short, AD and Ranasinghe, R (2011a) Assessment and integration of conventional, RTK-GPS and image-derived beach survey methods for daily to decadal coastal monitoring. Coastal Engineering 58, 194205.CrossRefGoogle Scholar
Harley, MD, Turner, IL, Short, AD and Ranasinghe, R (2011b) A reevaluation of coastal embayment rotation: The dominance of cross-shore versus alongshore sediment transport processes, Collaroy-Narrabeen Beach, southeast Australia. Journal of Geophysical Research: Earth Surface 116, F04033.CrossRefGoogle Scholar
Holman, RA, Sallenger, AH, Lippmann, TC and Haines, JW (1993) The application of video image processing to the study of nearshore processes. Oceanography 6(3), 7885.CrossRefGoogle Scholar
Ibaceta, R, Splinter, KD, Harley, MD and Turner, IL (2020) Enhanced coastal shoreline modeling using an ensemble Kalman filter to include nonstationarity in future wave climates. Geophysical Research Letters 47, e2020GL090724.CrossRefGoogle Scholar
Jara, MS, González, M and Medina, R (2015) Shoreline evolution model from a dynamic equilibrium beach profile. Coastal Engineering 99, 114.CrossRefGoogle Scholar
Kato, S, Okabe, T and Sawahara, N (2013) Influence of long-period variation on shoreline change. In Proceedings of the 7th International Conference on Asian and Pacific Coasts (APAC2013), pp. 9195. Hasanuddin University Press.Google Scholar
Kishimoto, R, Shimura, T, Mori, N and Mase, H (2017) Statistical modeling of global mean wave height considering principal component analysis of sea level pressures and its application to future wave height projection. Hydrological Research Letters 11(1), 5157.CrossRefGoogle Scholar
Klein, AHDF, Benedet, L and Schumacher, DF (2002) Short-term beach rotation processes in distinct headland bay beach systems. Journal of Coastal Research 18(3), 442458.Google Scholar
Kuriyama, Y and Banno, M (2016) Shoreline change caused by the increase in wave transmission over a submerged breakwater due to sea level rise and land subsidence. Coastal Engineering 112, 916.CrossRefGoogle Scholar
Kuriyama, Y, Banno, M and Suzuki, T (2012) Linkages among interannual variations of shoreline, wave and climate at Hasaki, Japan. Geophysical Research Letters 39, L06604.CrossRefGoogle Scholar
Kuriyama, Y and Yanagishima, S (2018) Regime shifts in the multi-annual evolution of a sandy beach profile. Earth Surface Processes and Landforms 43, 31333141.CrossRefGoogle Scholar
LaFond, EC (1939) Sand movement near the beach in relation to tides and waves. Proceeding of 6th Pacific Science Congress 8, 795799.Google Scholar
Larson, M and Kraus, NC (1994) Temporal and spatial scales of beach profile change, Duck, North Carolina. Marine Geology 117, 7594.CrossRefGoogle Scholar
Lazarus, ED, Harley, MD, Blenkinsopp, CE and Turner, IL (2019) Environmental signal shredding on sandy coastlines. Earth Surface Dynamics 7(1), 7786.CrossRefGoogle Scholar
Lemos, C, Floc’h, F, Yates, M, Le Dantec, N, Marieu, V, Hamon, K, Cuq, V, Suanez, S and Delacourt, C (2018) Equilibrium modeling of the beach profile on a macrotidal embayed low tide terrace beach. Ocean Dynamics 68, 12071220.CrossRefGoogle Scholar
Lesser, GR, Roelvink, JA, van Kester, JATM and Stelling, GS (2004) Development and validation of a three-dimensional morphological model. Coastal Engineering 51, 883915.CrossRefGoogle Scholar
Lippmann, TC and Holman, RA (1990) The spatial and temporal variability of sandbar morphology. Journal of Geophysical Research: Oceans 95, 1157511590.CrossRefGoogle Scholar
Ludka, BC, Guza, RT, O’Reilly, WC, Merrifield, MA, Flick, RE, Bak, AS, Hesser, T, Bucciarelli, R, Olfe, C, Woodward, B, Boyd, W, Smith, K, Okihiro, M, Grenzeback, R, Parry, L and Boyd, G (2019) Sixteen years of bathymetry and waves at San Diego beaches. Scientific Data 6, 161.CrossRefGoogle Scholar
Luijendijk, A, Hagenaars, G, Ranasinghe, R, Baart, F, Donchyts, G and Aarninkhof, S (2018) The state of the World’s beaches. Scientific Reports 8, 6641.CrossRefGoogle ScholarPubMed
Masselink, G and Li, L (2001) The role of swash infiltration in determining the beachface gradient: A numerical study. Marine Geology 176(1–4), 139156.CrossRefGoogle Scholar
Matsuba, Y and Sato, S (2018) Nearshore bathymetry estimation using UAV. Coastal Engineering Journal 60(1), 5159.CrossRefGoogle Scholar
McCall, RT, Masselink, G, Poate, TG, Roelvink, JA and Almeida, LP (2015) Modelling the morphodynamics of gravel beaches during storms with XBeach-G. Coastal Engineering 103, 5266.CrossRefGoogle Scholar
McLean, R and Shen, JS (2006) From foreshore to foredune: Foredune development over the last 30 years at Moruya Beach, New South Wales, Australia. Journal of Coastal Research 22, 2836.CrossRefGoogle Scholar
Miller, JK and Dean, RG (2004) A simple new shoreline change model. Coastal Engineering 51, 531556.CrossRefGoogle Scholar
Montaño, J, Coco, G, Antolínez, JAA, Beuzen, T, Bryan, KR, Cagigal, L, Castelle, B, Davidson, MA, Goldstein, EB, Ibaceta, R, Idier, D, Ludka, BC, Masoud-Ansari, S, Méndez, FJ, Murray, AB, Plant, NG, Ratliff, KM, Robinet, A, Rueda, A, Sénéchal, N, Simmons, JA, Splinter, KD, Stephens, S, Townend, I, Vitousek, S and Vos, K (2020) Blind testing of shoreline evolution models. Scientific Reports 10, 2137.CrossRefGoogle ScholarPubMed
Nguyen, QH and Takewaka, S (2020) Land subsidence and its effects on coastal erosion in the Nam Dinh Coast (Vietnam). Continental Shelf Research 207, 104227.CrossRefGoogle Scholar
Nguyen, QH and Takewaka, S (2022) Historical reconstruction of shoreline evolution at the Nam Dinh Coast, Vietnam. Coastal Engineering Journal. Advance online publication. https://doi.org/10.1080/21664250.2022.2073748CrossRefGoogle Scholar
Nicholls, RJ, Birkemeier, WA and Lee, G (1998) Evaluation of depth of closure using data from Duck, NC, USA. Marine Geology 148, 179201.CrossRefGoogle Scholar
O’Dea, A, Brodie, KL and Hartzell, P (2019) Continuous coastal monitoring with an automated terrestrial lidar scanner. Journal of Marine Science Engineering 7(2), 37.CrossRefGoogle Scholar
Pastol, Y (2011) Use of airborne LIDAR bathymetry for coastal hydrographic surveying: The French experience. Journal of Coastal Research 62, 618.CrossRefGoogle Scholar
Phillips, MS, Blenkinsopp, CE, Splinter, KD, Harley, MD and Turner, IL (2019) Modes of berm and beachface recovery following storm reset: Observations using a continuously scanning lidar. Journal of Geophysical Research: Earth Surface 124, 720736.CrossRefGoogle Scholar
Pleskachevsky, A, Lehner, S, Heege, T and Mott, C (2011) Synergy and fusion of optical and synthetic aperture radar satellite data for underwater topography estimation in coastal areas. Ocean Dynamics 61(12), 20992120.CrossRefGoogle Scholar
Ranasinghe, R, Callaghan, D and Stive, MJF (2012) Estimating coastal recession due to sea level rise: Beyond the Bruun rule. Climate Change 110, 561574.CrossRefGoogle Scholar
Robinet, A, Castelle, B, Idier, D, Cozannet, GL, Déqué, M and Charles, E (2016) Statistical modeling of interannual shoreline change driven by North Atlantic climate variability spanning 2000–2014 in the Bay of Biscay. Geo-Marine Letters 36, 479490.CrossRefGoogle Scholar
Roelvink, D, Reniers, A, van Dongeren, A, van Thiel de Vries, J, McCall, R and Lescinski, J (2009) Modelling storm impacts on beaches, dunes and barrier islands. Coastal Engineering 56, 11331152.CrossRefGoogle Scholar
Ruessink, BG, Kuriyama, Y, Reniers, AJHM, Roelvink, JA and Walstra, DJR (2007) Modeling cross-shore sandbar behavior on the timescale of weeks. Journal of Geophysical Research: Earth Surface 112, F03010.CrossRefGoogle Scholar
Ruessink, BG, Wijnberg, KM, Holman, RA, Kuriyama, Y and van Enckevort, IMJ (2003) Intersite comparison of interannual nearshore bar behavior. Journal of Geophysical Research: Oceans 108, 3249.CrossRefGoogle Scholar
Shepard, FP (1950) Beach cycles in Southern California. U.S. Army Corps of Engineers. Beach Erosion Board Technical Memorandum No. 20.Google Scholar
Short, A and Trenbanis, AC (2004) Decadal scale patterns in beach oscillation and rotation Narrabeen Beach, Australia time series, PCA and wavelet analysis. Journal of Coastal Research 20, 523532.CrossRefGoogle Scholar
Southgate, HN (2011) Data-based yearly forecasting of beach volumes along the Dutch North Sea coast. Coastal Engineering, 58, 749760.CrossRefGoogle Scholar
Splinter, KD, Turner, IL, Davidson, MA, Barnard, P, Castelle, B and Oltman-Shay, J (2014) A generalized equilibrium model for predicting daily to interannual shoreline response. Journal of Geophysical Research: Earth Surface 119, 19361958.CrossRefGoogle Scholar
Tamura, T, Oliver, TSN, Cunningham, AC and Woodroffe, CD (2019) Recurrence of extreme coastal erosion in SE Australia beyond historical timescales inferred from beach ridge morphostratigraphy. Geophysical Research Letters 46, 47054714.CrossRefGoogle Scholar
Thom, B and Hall, W (1991) Behaviour of beach profiles during accretion and erosion dominated periods. Earth Surface Processes and Landforms 16, 113127.CrossRefGoogle Scholar
Turki, I, Medina, R, Coco, G and Gonzalez, M (2013) An equilibrium model to predict shoreline rotation of pocket beaches. Marine Geology 346, 220232.CrossRefGoogle Scholar
Turner, IL, Harley, MD and Drummond, CD (2016) UAVs for coastal surveying. Coastal Engineering 114, 1924.CrossRefGoogle Scholar
Turner, IL, Harley, MD, Short, AD, Simmons, JA, Bracs, MA, Phillips, MS and Splinter, KD (2016) A multi-decade dataset of monthly beach profile surveys and inshore wave forcing at Narrabeen, Australia. Scientific Data 3(1), 113.CrossRefGoogle ScholarPubMed
Udo, K, Ranasinghe, R and Takeda, Y (2020) An assessment of measured and computed depth of closure around Japan. Scientific Reports 10, 2987.CrossRefGoogle ScholarPubMed
Vitousek, S, Barnard, PL, Limber, P, Erikson, L and Cole, B (2017) A model integrating longshore and cross-shore processes for predicting long-term shoreline response to climate change. Journal of Geophysical Research: Earth Surface 122, 782806.CrossRefGoogle Scholar
Vitousek, S, Cagigal, L, Montaño, J, Rueda, A, Mendez, F, Coco, G and Barnard, PL (2021) The application of ensemble wave forcing to quantify uncertainty of shoreline change predictions. Journal of Geophysical Research: Earth Surface 126, e2019JF005506.Google Scholar
Warrick, JA, Bountry, JA, East, AE, Magirl, CS, Randle, TJ, Gelfenbaum, G, Ritchie, AC, Pess, GR, Leung, V and Duda, JJ (2015) Large-scale dam removal on the Elwha River, Washington, USA: Source-to-sink sediment budget and synthesis. Geomorphology 246, 729750.CrossRefGoogle Scholar
Yates, ML, Guza, RT and O’Reilly, WC (2009) Equilibrium shoreline response: Observations and modeling. Journal of Geophysical Research: Oceans 114, C09014.CrossRefGoogle Scholar
Yates, ML, Guza, RT, O’Reilly, WC, Hansen, JE and Barnard, PL (2011) Equilibrium shoreline response of a high wave energy beach. Journal of Geophysical Research 116, C04014.CrossRefGoogle Scholar
Figure 0

Figure 1. Relationship between cost and information of beach survey methods.

Author comment: What can long-term in situ monitoring data tell us about our coastlines? — R0/PR1

Comments

No accompanying comment.

Review: What can long-term in situ monitoring data tell us about our coastlines? — R0/PR2

Comments

Comments to Author: The central opinion of the paper is to suggest long-term high-frequency monitoring is highly needed. Indeed, in situ monitoring data is very important for understanding coalstine changes and for modelling. But that argument that long-term high frequency data is needed is doubtful. The coasltine changes are results of beach morphological changes that have processes and landforms at different temporal and spatial scales.For example, high frequency observations are not needed for large-scale landforms and their formations. Likewise, small scale bedforms like ripples are formed by short-term wave processes. Currently, there are no evidences showing such small-cale bedforms have important direct effects on coastline changes. Undertanding short term processes can be studied only by short-term high frequency observations. It is not nessesary to have long-term high frequency observations and monitors.

Moreover, a review about monitoring approaches at different scales are interesting. But ,this paper have a large portion talking about beach processes and modelling approaches, which deviate from my expectation from the title. There are already many reviews about modelling approaches and beach processes. Therefore, the author may need to significantly revise the paper and resubmit it.

Review: What can long-term in situ monitoring data tell us about our coastlines? — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: Review: What can long-term in situ monitoring data tell us about our coastlines?

Mitchell Harley, 24/08/2022

This is an interesting discussion piece/short review on the value of long-term foreshore measurements, with suggestions for knowledge gaps. I found the paper a valuable summary of the literature and worthy of publication. While the paper covers many interesting points related to long-term measurements, there does appear to be a few gaps that I think need addressing. The first is related to the role of shoreface measurements. While the paper makes it clear that it is focused on foreshore monitoring, the focus on sea-level rise and the Bruun rule in particular suggests that monitoring of the shoreface should not be neglected in this review. I think this is important as shoreface changes can be crucial to driving long-term shoreline changes, particularly due to sea-level rise (as summarised by Anthony and Aagard, 2020 in their review of the shoreface, also Harley et al., 2022 https://www.nature.com/articles/s43247-022-00437-2). Could the author comment on how shoreface measurements (e.g. the JARKUS monitoring program in the Netherlands) could also help complement this monitoring?

The other knowledge gap appears in the use of autonomous vehicles. Curiously, UAV is not mentioned in the paper (or in Figure 1) but this has been the focus of much recent monitoring advances. How does the author believe UAV fit in to long-term monitoring? Is there a role for new advances such as submerged autonomous vehicles in this?

Some other specific comments I have related to certain points made in the manuscript are below:

1) I do not necessarily agree with the statement that LIDAR “are expensive for monitoring large areas and are unsuitable for long-term high-frequency observation campaigns”. This might be true for Airborne LIDAR, but not fixed LIDAR systems (e.g. Phillips et al., 2019, O’Dea et al., 2019) that can easily obtain high-frequency (5 Hz) data continuously over many years. It is my experience that these systems cost less than ARGUS, so I think they should be added into Figure 1, perhaps separating LIDAR into Airborne and Fixed?

2) Regarding seasonal variation at Narrabeen, the data does reveal seasonal beach rotation (refer Harley et al., 2015), which is attributed to the seasonal variability in wave direction between summer and winter.

3) I do not necessarily agree with the statement regarding “the long-term robustness of equilibrium-based shoreline change models is very high because the simulation results do not diverge”. In fact, these models do diverge quite considerably when forced with nonstationary wave conditions, which is a challenge that is currently being addressed (refer Ibaceta et al, 2020 https://agupubs.onlinelibrary.wiley.com/doi/full/10.1029/2020GL090724)

4) “hydrographic experiments” - do you mean “laboratory experiments”? The references suggest that is the case

5) “it might be possible to quantify the effects of tides on beach morphological changes during flood and ebb tides”. I agree that this is something that is needed - it has already been demonstrated at the local scale in Phillips et al (2019) in Figure 6 where there is a clear link between the berm crest elevation and cycles of spring/neap tides

6) “have been released to the public under the concept of open data” as a comment, it would be great to see the Hasaki dataset released as open data as well!

As I reiterate, I found this manuscript a useful contribution and would be happy to see it published following consideration of the above points/suggestions.

Recommendation: What can long-term in situ monitoring data tell us about our coastlines? — R0/PR4

Comments

No accompanying comment.

Decision: What can long-term in situ monitoring data tell us about our coastlines? — R0/PR5

Comments

No accompanying comment.

Author comment: What can long-term in situ monitoring data tell us about our coastlines? — R1/PR6

Comments

No accompanying comment.

Review: What can long-term in situ monitoring data tell us about our coastlines? — R1/PR7

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: I thank the Author's careful consideration of the points raised in my review. I have read the revised manuscript and believe that the manuscript has been strengthened in the new version. My recommendation is that the manuscript is worthy of publishing. I do have however a couple of minor edits of the new additions that the author may wish to consider. These are considered minor and for that reason I have selected "Accept" (and do not believe I need to see the manuscript again). These are listed below:

Line 43: “while the lower shoreface is not always monitored adequately”. “Not always” I think is an understatement, I am not aware of many/any adequate lower shoreface monitoring program. I therefore personally think this could be strengethed to “the lower shoreface is rarely monitored adequately”

Line 79: does “long periods” here refer to long wave periods, or periods of time? This could be clarified

Line 146: “wave energies” should not be plural here. Suggest “incident wave energy”

Recommendation: What can long-term in situ monitoring data tell us about our coastlines? — R1/PR8

Comments

Comments to Author: Dear Dr Banno.

Many thanks for submitting your revised manuscript, along with the point by point response to the reviewer 1.

The manuscript is much improved and has been re-reviewed and deemed acceptable. I agree that this manuscript is now worthy of publication. I have ticked 'minor revision' because reviewer 1 raises a few minor suggestions, which I think are worth implementing (pasted below). This will not take long, but will improve the manuscript. Please implement these suggestions before we transition the paper to the next stage of publication. Well done, this is an interesting and useful contribution to the journal.

Reviewer 1 additional minor suggestions:

I thank the Author's careful consideration of the points raised in my review. I have read the revised manuscript and believe that the manuscript has been strengthened in the new version. My recommendation is that the manuscript is worthy of publishing. I do have however a couple of minor edits of the new additions that the author may wish to consider. These are considered minor and for that reason I have selected "Accept" (and do not believe I need to see the manuscript again). These are listed below:

Line 43: “while the lower shoreface is not always monitored adequately”. “Not always” I think is an understatement, I am not aware of many/any adequate lower shoreface monitoring program. I therefore personally think this could be strengethed to “the lower shoreface is rarely monitored adequately”

Line 79: does “long periods” here refer to long wave periods, or periods of time? This could be clarified

Line 146: “wave energies” should not be plural here. Suggest “incident wave energy”

Decision: What can long-term in situ monitoring data tell us about our coastlines? — R1/PR9

Comments

No accompanying comment.

Author comment: What can long-term in situ monitoring data tell us about our coastlines? — R2/PR10

Comments

No accompanying comment.

Recommendation: What can long-term in situ monitoring data tell us about our coastlines? — R2/PR11

Comments

No accompanying comment.

Decision: What can long-term in situ monitoring data tell us about our coastlines? — R2/PR12

Comments

No accompanying comment.