1. Introduction
The interactions between turbulent winds and ocean waves play essential roles in many important atmosphere–ocean phenomena. They drive the exchange of mass, momentum, and heat between the atmosphere and the ocean, which are key processes needed to validate and improve models of the atmosphere, the upper ocean, and the waves. To understand and parameterize the processes occurring at, above, and below the wavy sea surface, an explicit description of the airflow over surface waves is needed to improve our knowledge of air–sea fluxes for weather and climate predictions.
The problem of flow over surface waves has been an active research topic for almost a century. Many theoretical studies (Jeffreys 1925; Phillips 1957; Miles 1957, 1993; Janssen 1991; Belcher and Hunt 1993; among others) have focused on explaining the mechanism of wind-wave generation and growth. Meanwhile, the quantification of wind-wave energy and momentum transfer and wavefield evolution has relied substantially on field (e.g., Dobson 1971; Elliott 1972; Snyder et al. 1981; Hristov et al. 2003; Donelan et al. 2006; Grare et al. 2013a; Hogstrom et al. 2015) and laboratory measurements (e.g., Hsu and Hsu 1983; Banner and Peirson 1998; Veron et al. 2007; Grare et al. 2013b; Buckley and Veron 2016).
In addition to theoretical and experimental studies, numerical simulations of turbulent winds over waves have greatly contributed to our understanding of these phenomena. Early studies focused on solving the Reynolds-averaged Navier–Stokes (RANS) equations (Townsend 1972; Gent and Taylor 1976; Al-Zanaidi and Hui 1984; Van Duin and Janssen 1992; Mastenbroek et al. 1996). The growth of computing power has made direct numerical simulation (DNS) possible for the study of turbulent winds over waves under idealized conditions (Sullivan et al. 2000; Kihara et al. 2007; Yang and Shen 2009, 2010). More recently, large-eddy simulation (LES) has become a common approach for the study of winds over ocean waves (Sullivan et al. 2008, 2014; Liu et al. 2010; Yang et al. 2013; Wu et al. 2017).
Recently, there has been a growing interest in waves traveling faster than the wind (often referred to as the regime of “wave-driven wind”) as Hanley et al. (2010) have shown that the wind and the waves are often not in equilibrium; that is, local wave conditions are not necessarily tied to the local wind conditions. That is, the wave-driven wind regime is a common feature around the globe. Sullivan et al. (2008), Hogstrom et al. (2009), Smedman et al. (2009), Soloviev and Kudryavtsev (2010), Hogstrom et al. (2015), and Wu et al. (2017) have focused on this wave-driven regime showing the existence of an upward transfer of momentum from the swell to the wind, leading sometimes to the presence of a jet in the lower part of the marine atmospheric boundary layer (MABL). In this regime, the amplitude of the wave-induced fluctuations of the wind components can compare with the amplitude of the background turbulence, playing a nonnegligible role in the production of turbulence, driving the exchange of momentum between the atmosphere and the ocean, and affecting the shape of the vertical profile of the mean wind speed (Hristov and Ruiz-Plancarte 2014). While the recent developments of LES are able to phase-resolve the airflow above 3D wave fields (Sullivan et al. 2014) and offer further promise to better understand the physics of the air–sea interaction, experimental data remain the ultimate test of these numerical simulations to describe real flows.
In the present paper, we present results from a field experiment showing that the wave-induced fluctuations depend both on the spectral wave age
2. The experiment
a. Experimental setup
The measurements described here were collected from the Floating Instrument Platform (R/P FLIP) moored approximately at the center of the triangle formed by the islands of San Nicholas, Santa Catalina, and San Clemente off the coast of Southern California (33°13.202′N, 118°58.767′W), in water 1160 m deep from 7 to 22 November 2013, during the ONR Southern California 2013 (SoCal2013) experiment. A telescopic mast was deployed at the end of the port board boom approximately 18 m from the 9-m-diameter hull of R/P FLIP (Figs. 1 and 2). The telescopic mast consists of four sections that slide into each other. Each section is approximately 3 m long. While the top section (section 4 in Fig. 2) was tightly coupled to the boom, the extension of each section underneath the boom (sections 1, 2, and 3, section 1 being the lowest) was adjustable. Each section of the mast was equipped with an ensemble of instruments deployed to measure the properties of the airflow above the ocean. Sections 1, 2, 3, and 4 have a 2-, 3-, 4-, and 5-in. square hollow section, respectively. A set of rigging lines connected sections 1, 2, and 3 to the hull of R/P FLIP to constrain the horizontal displacement of the mast.
Experimental setup on R/P FLIP. Details of the telescopic mast are presented in Fig. 2. The following instruments were mounted on the booms of the R/P FLIP: IMU/GPS (SPAN-CPT, Hemisphere), laser wave gauges (ILM-500), net radiometer (CNR1), and wind lidar (Windcube).
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
(left) The telescopic mast with (right) details of the different sections. The following instruments were mounted on the mast: sonic anemometers (CSAT3 and Gill R3-50), temperature–humidity sensors (HC2S3), IMU (AHRS400, MTi300), hygrometer CO2 sensor (LICOR7500), and laser wave gauges (ILM-500).
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
The top section (section 4) of the mast, lengthened by a 2-m vertical pole (with a 2-in. square cross section), was equipped at its highest point with a 3D sonic anemometer (Campbell CSAT3) and with an open-path infrared hygrometer CO2 sensor (LICOR 7500). These two instruments were mounted on a horizontal pole (with a 2-in. cross section) perpendicular both to the mast and the boom, 1 m away from the mast. At the intersection between the mast and the horizontal pole were mounted a 6-degrees-of-freedom inertial measurement unit [IMU; MEMSIC Attitude Heading Reference System (AHRS) 400] and a relative humidity–temperature sensor (Campbell HC2S3). All these instruments were located about 3 m above the walkway of the boom. As for section 4, each moving section 1–3 of the mast was equipped, at its base, with a 3D sonic anemometer CSAT3, an IMU, and a humidity–temperature sensor (HC2S3). The main differences between the moving sections and section 4 are that the moving sections are not equipped with a hygrometer CO2 sensor, the IMU model is different (Xsens MTi 300), and the IMU is mounted directly behind the anemometer. In addition to these instruments, the lowest section of the mast was equipped with a 3D sonic anemometer (Gill R3-50) mounted upside down directly below the lowest point of the mast. Hence, the measuring volumes of the Gill and the lowest CSAT3 were separated by 0.85 m in the vertical direction and 1 m in the horizontal direction. Aside the base of the Gill, a single-point laser altimeter (MDL ILM-500) was mounted to monitor the distance between the bottom of the mast and the sea surface. This laser altimeter was also used as a wave gauge. This setup ensured that all four CSAT3 anemometers were strictly one above the other.
The total extension of the mast was controlled by an electrical winch, while the relative extension of each section was adjusted with a set of rigging lines, manually controlled from the hull of R/P FLIP. The extension of the mast was set to bring the lowest instruments as close as possible to the mean sea level (MSL) while preserving the sensors’ integrity and distributing the instruments along the vertical axis. Thus, while the highest anemometer remained at 14.8 m (±0.2 m) above MSL, the lowest anemometer was positioned at an adjustable height, as low as 2.6 m from the MSL (see Fig. 3b).
Environmental conditions of the marine atmospheric boundary layer: (a)
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
Reference values of humidity and temperature were provided by Campbell HC2S3 sensors. The hygrometer CO2 instrument (LICOR 7500), which is also equipped with a pressure sensor, was used to measure the fast fluctuations of the water vapor density and the absolute pressure of the air. Each sonic anemometer measured the three components of the wind.
The temperature–humidity sensors, the hygrometer CO2 sensor, and the sonic anemometers were sampled at 20 Hz. The laser altimeter is internally sampled at 12 kHz and was averaged down to 100 Hz. The MEMSIC IMU was sampled at 20 Hz, while the Xsens IMUs were sampled at 50 Hz.
In addition to the instruments mounted on the mast, an array of five laser wave gauges (MDL ILM-500, sampled at 100 Hz) located on the starboard, port, and face booms of R/P FLIP were deployed to measure the directional spectrum of the sea surface displacement. A Global Navigation Satellite System (GNSS) and Inertial Navigation System (INS; NovAtel SPAN-CPT) was deployed to measure the motion of R/P FLIP. This device combines GNSS and INS solutions to deliver position, velocity, and altitude angles with very high accuracy (the accuracy in absolute position is only a few centimeters). The combined solution was sampled at 100 Hz.
A temperature chain (PME T-Chain; 17 thermistors sampled at 0.8 Hz) was deployed from the starboard boom of R/P FLIP to monitor the temperature of the first 50 m of the water column. Finally, a wind lidar system (Leosphere Windcube WLS7) was deployed on the starboard boom to measure the vertical profile of the wind every 20 m from 50 up to 400 m above the MSL. The four-beam lidar provided a triangulated measure of the wind components every 4 s.
b. Data postprocessing and analysis
The sonic velocities and the air–sea interface displacements were corrected to account for the motion of R/P FLIP. Those corrections were computed combining solutions from both the IMUs and the SPAN-CPT. The combined solution was used to correct the wind components following Edson et al. (1998). The sea surface elevations measured by the laser wave gauges were corrected for the vertical displacements of the instruments according to the solutions provided by the SPAN-CPT instrument.






Using the same dataset, Grare et al. (2016) showed wind speed discrepancies between measurements from the CSAT3 and the Gill R3-50. They found that differences in the mean wind speed could reach up to 4%, a few degrees in the mean wind direction, up to 5% for the standard deviation of the vertical wind component
Finally, 30-min records when the sonic anemometers were located in the wake of the hull of R/P FLIP or in the wake of the mast were discarded from the analysis. The wake from the hull of R/P FLIP was assumed to be contained within a ±45° segment about the port boom. Similarly, the wake from the mast was assumed to be contained within a ±45° segment about the direction the CSAT3s were pointing to.







c. Environmental conditions
Figure 3 shows the variation of statistical parameters of the marine atmospheric boundary layer (MABL) during the SoCal2013 experiment. The horizontal axis represents the date in days from 11 to 21 November 2013 using the coordinated universal time (UTC). Data points are 30-min averages. The data presented hereinafter do not include measurements performed prior to 11 November 2013 because all the instruments described in the previous section were not fully operational until then. Figure 3a shows the mean wind speed at 10 m
Figure 3b shows the heights of the sonic anemometers during the course of the experiment. These heights are relative to the MSL. The mast was retracted when the waves could potentially hit and damage the lowest anemometer (e.g., on 15 and 16 November 2013).













(a) Mean wind speed as a function of time and height measured by both the sonic anemometers on the telescopic mast (2.5 < z < 15 m) and the lidar (60 < z < 500 m). (b) Vertical profiles of the normalized mean wind speed
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
For each 30-min record, the omnidirectional frequency spectrum
Example of the omnidirectional and directional spectra of the wave field from 30-min records starting at 0230 UTC 11 Nov 2013. (a) Omnidirectional spectra from measurements of the laser wave gauge located at the bottom of the telescopic mast (red) and by the azimuthal integration of the directional spectrum (black). (b) Directional spectrum from the array of laser wave gauges. (c) Smoothed directional spectrum used to partition the wave field. For the directional spectra, the direction refers to the direction the waves move to and the frequencies correspond to concentric circles logarithmically spaced. The colored boundaries delimit the main partitions of the wave field: southern swell (purple), northern swell (black), local swell (green), and wind waves (orange). The solid colored dots correspond to the mean frequency and direction of each partition.
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
For the directional spectrum (Figs. 5b,c), the frequency scale corresponds to concentric circles that are logarithmically spaced, and the azimuthal direction corresponds to the direction of propagation of the waves. The colored contours delimit the different partitions of the wave field. This example emphasizes the utility of the directional spectrum, as in the omnidirectional spectrum only three peaks are distinguishable, while four wave partitions are clearly visible in the directional spectrum. The partitioning of the wave field was performed running a script based on the methods described in Hanson and Phillips (2001) and Portilla et al. (2009). As suggested in Portilla et al. (2009), we applied to the directional spectra a 2D noise-removal filter before partitioning the spectrum using a “watershed” algorithm.1 In Fig. 5b, the raw directional spectrum is plotted while its filtered form is plotted in Fig. 5c.
The example presented is characteristic of the conditions encountered during the experiment. The wave field was formed by the superposition of four main components:
Wind waves (delimited by the orange contour) directly associated with local winds, with frequencies generally higher than 0.1 Hz.
Local north–northwest swell (delimited by the green contour), residual of wind waves generated upwind of our location [say, O(100) km], which propagated to our position after the wind died. These waves were passing by for short periods of time (less than 2 days) with frequencies ranging between 0.08 and 0.15 Hz.
Old (remote) north–northwest swell (delimited by the black contour) generated by storms in the North Pacific Ocean, passing by our location for up to four days, with the longest waves (down to ~0.06 Hz) arriving before the shorter ones (up to ~0.1 Hz).
Swell from the Southern Hemisphere (delimited by the purple contour) generated by remote storms in the Southern Hemisphere, with frequencies in the range of 0.07–0.1 Hz.
For each 30-min record and for each wave component defined by an ensemble of pairs (
- The mean frequency2
- The mean direction
Figure 6a shows the omnidirectional spectrogram of the surface displacement. The
Characteristics of the wave field during the experiment. (a) Spectrogram of the surface displacement. (b) Significant wave height
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1




The mean frequency
These two bottom panels show that the wave field was quite complex during this experiment, with a superposition of local wind waves, young and old swells from the north–northwest, and long swell from the south where the relative energy contribution of each wave component varies strongly over time.
The plots presented in Figs. 3 and 6 show dropouts in the data (e.g., on 17 November 2013). These dropouts either were the result of the loss of GPS signal from the SPAN-CPT inertial unit because of temporary technical issues or were the result of periods of low wind speeds when the laser wave gauges had low returns from the glassy sea surface. Cases with a percentage of returns less than 12.5% were excluded from our analysis. This 12.5% threshold was selected because we found that below this threshold the temporal and spatial distributions of the returns were too scattered to properly reconstruct the wave profile.
3. Wave-induced fluctuations of the wind
a. Coherence and phase
Figures 7 and 8 show 11 days of the squared coherence and phase shift between the waves and the wind measured by each anemometer. The left panels show the coupling with the horizontal fluctuations
The squared coherences
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
The phase shift
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
Figure 8 shows that the phase between the waves and the wind is relatively constant below the critical layers, the horizontal component being out of phase (about 180°) with the waves while the vertical component is in quadrature (about 90°). For most of the data, the phase above the critical layers is very noisy and appears random. This is a direct effect of the low level of coherence square
b. Selection of the wave-coherent fluctuations
To study the variation of the wave-induced velocities with wind forcing and height, we have selected fluctuations of the wind coherent with the waves as done in Hare et al. (1997) and Grare et al. (2013a), that is, according to the level of squared coherence between the surface elevation and the velocities. To minimize the effect of spurious noise and to ensure reliable wave-induced signals, only the frequencies for which the squared coherence
c. Double dependence and parameterization























As the wave field was complex in our experiment with waves propagating in different directions, each wave component is likely to generate its own wave-induced field of velocity in its direction of propagation. To account for the three-dimensional properties of the wave-induced airflow, we regrouped the streamwise










(a)–(j) The amplitude of the dimensionless wave-induced velocities (left)
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
The double dependence of the wave-induced velocities on both the wave age and the dimensionless height, is presented in the Fig. 10, where the variations of
The bin-averaged amplitude of the (left) horizontal and (right) vertical dimensionless wave-induced velocities, (a),(b) as a function of the spectral wave age
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
Figures 10a and 10b show that the amplitudes of the wave-induced velocities decrease for wave ages close to 1. Figures 10c and 10d show that wave-induced velocities decay with normalized height


















Best-fit coefficients of proposed wave-induced velocity parameterizations.
Figure 11 shows the amplitude of the wave-induced fluctuations scaled by the functions
The bin-averaged amplitude of the (left) horizontal and (right) vertical dimensionless wave-induced velocities (a),(b) corrected for their dependence on kz plotted as a function of the spectral wave age
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1





















The bin-averaged amplitude of the (left) horizontal and (right) vertical dimensionless wave-induced velocities (a),(b) corrected for their dependence on kz plotted as a function of the spectral wave age
Citation: Journal of Physical Oceanography 48, 12; 10.1175/JPO-D-18-0121.1
In Eq. (18), we introduced two degrees of freedom as
4. Discussion and conclusions
Field measurements of the airflow above ocean waves were conducted using a vertical array of ultrasonic anemometers mounted on a telescopic mast attached to R/P FLIP. The spectral decomposition of the wind fluctuations led to the quantification of the amplitude and phase of the wave-induced fluctuations of the three components of the wind for frequencies ranging from 0.06 to 0.5 Hz. The telescopic mast permitted adjusting the height of the anemometers to adapt to environmental conditions, yielding measurements of the wind velocities from 2.6 m up to 15 m above the MSL. The range of heights covered by the array of anemometers coupled with the spectral decomposition of the wind fluctuations led to a description of the wave-induced velocities at normalized heights
Measurements at several heights along with a proper normalization of the wave-induced velocities by the orbital velocity
Our results are consistent with the early work from Kitaigorodskii et al. (1983) with data collected on Lake Ontario over a small range of wave ages (
We showed that the dependence on the wave age
First, it is assumed that, at a given frequency, all the waves have the same wave age; in other words, the wind forcing (i.e., the inverse wave age) is the same regardless of the angle between the wind and the wave directions. However, the formulation of the inverse wave age accounting for the misalignment between the wind and the waves is a still controversial topic (Hanley et al. 2010; Hogstrom et al. 2011; Hanley et al. 2011), and further discussion of this subject is beyond the scope of this paper.
Second, for a case where two or more wave fields with the same dominant phase speed c are traveling in different directions, it is assumed that their corresponding horizontal wave-induced velocities do not interact with each other and that the amplitude of the horizontal wave-induced fluctuations coherent with the phase of these wave fields is linearly related to the vertical displacement of the surface resulting from the superposition of these wave fields.
Regarding the two points above, we emphasize that for most of the experiment, the waves were mostly aligned with the wind and that when it was not the case, for example, when a south swell was superposed upon the northwest wind and waves, the wave energy associated with the south swell accounted for less than 20% of the total wave energy. Therefore, we believe that using a spectral wave age that does not account for the misalignment between the wind and the waves, and presenting data in the form of
Acknowledgments
We thank Tom Golfinos and the crew of the R/P FLIP for their tireless support of this research and Captain Bill Gaines for his logistical support of FLIP operations. The measurements would not have been possible without the work of Peter Sutherland, Luc Deike, and Nick Statom during the experiment. The manuscript also benefited from the helpful suggestions of two anonymous reviewers. This research was supported by funding to WKM from the physical oceanography programs at ONR (Grants N00014-14-1-0710 and N00014-12-1-1022) and NSF (Grants OCE-1634289 and OCE-1155403).
REFERENCES
Al-Zanaidi, M. A., and W. H. Hui, 1984: Turbulent airflow over water waves—A numerical study. J. Fluid Mech., 148, 225–246, https://doi.org/10.1017/S0022112084002329.
Banner, M. L., and W. L. Peirson, 1998: Tangential stress beneath wind-driven air-water interfaces. J. Fluid Mech., 364, 115–145, https://doi.org/10.1017/S0022112098001128.
Belcher, S. E., and J. C. R. Hunt, 1993: Turbulent shear flow over slowly moving waves. J. Fluid Mech., 251, 109–148, https://doi.org/10.1017/S0022112093003350.
Bendat, J. S., and A. G. Piersol, 2010: Random Data Analysis and Measurement Procedures. 4th ed. Wiley and Sons, 640 pp.
Buckley, M. P., and F. Veron, 2016: Structure of the airflow above surface waves. J. Phys. Oceanogr., 46, 1377–1397, https://doi.org/10.1175/JPO-D-15-0135.1.
Dobson, F. W., 1971: Measurements of atmospheric pressure on wind-generated sea waves. J. Fluid Mech., 48, 91–127, https://doi.org/10.1017/S0022112071001496.
Donelan, M. A., A. V. Babanin, I. R. Young, and M. L. Banner, 2006: Wave-follower field measurements of the wind-input spectral function. Part II: Parameterization of the wind input. J. Phys. Oceanogr., 36, 1672–1689, https://doi.org/10.1175/JPO2933.1.
Edson, J. B., A. A. Hinton, K. E. Prada, J. E. Hare, and C. W. Fairall, 1998: Direct covariance flux estimates from mobile platforms at sea. J. Atmos. Oceanic Technol., 15, 547–562, https://doi.org/10.1175/1520-0426(1998)015<0547:DCFEFM>2.0.CO;2.
Elliott, J. A., 1972: Microscale pressure fluctuations near waves being generated by the wind. J. Fluid Mech., 54, 427–448, https://doi.org/10.1017/S0022112072000783.
Gent, P. R., and P. A. Taylor, 1976: A numerical model of the air flow above water waves. J. Fluid Mech., 77, 105–128, https://doi.org/10.1017/S0022112076001158.
Grare, L., L. Lenain, and W. K. Melville, 2013a: Wave-coherent airflow and critical layers over ocean waves. J. Phys. Oceanogr., 43, 2156–2172, https://doi.org/10.1175/JPO-D-13-056.1.
Grare, L., W. L. Peirson, H. Branger, J. W. Walker, J.-P. Giovanangeli, and V. Makin, 2013b: Growth and dissipation of wind-forced, deep-water waves. J. Fluid Mech., 722, 5–50, https://doi.org/10.1017/jfm.2013.88.
Grare, L., L. Lenain, and W. K. Melville, 2016: The influence of wind direction on Campbell scientific CSAT3 and Gill R3-50 sonic anemometer measurements. J. Atmos. Oceanic Technol., 33, 2477–2497, https://doi.org/10.1175/JTECH-D-16-0055.1.
Hanley, K. E., S. E. Belcher, and P. P. Sullivan, 2010: A global climatology of wind–wave interaction. J. Phys. Oceanogr., 40, 1263–1282, https://doi.org/10.1175/2010JPO4377.1.
Hanley, K. E., S. E. Belcher, and P. P. Sullivan, 2011: Reply. J. Phys. Oceanogr., 41, 1814–1817, https://doi.org/10.1175/JPO-D-11-051.1.
Hanson, J. L., and O. M. Phillips, 2001: Automated analysis of ocean surface directional wave spectra. J. Atmos. Oceanic Technol., 18, 277–293, https://doi.org/10.1175/1520-0426(2001)018<0277:AAOOSD>2.0.CO;2.
Hare, J. E., T. Hara, J. B. Edson, and J. M. Wilczak, 1997: A similarity analysis of the structure of airflow over surface waves. J. Phys. Oceanogr., 27, 1018–1037, https://doi.org/10.1175/1520-0485(1997)027<1018:ASAOTS>2.0.CO;2.
Hogstrom, U., A. Smedman, E. Sahlee, W. M. Drennan, K. K. Kahma, H. Pettersson, and F. Zhang, 2009: The atmospheric boundary layer during swell: A field study and interpretation of the turbulent kinetic energy budget for high wave ages. J. Atmos. Sci., 66, 2764–2779, https://doi.org/10.1175/2009JAS2973.1.
Hogstrom, U., A. Smedman, A. Semedo, and A. Rutgersson, 2011: Comments on “A global climatology of wind–wave interaction.” J. Phys. Oceanogr., 41, 1811–1813, https://doi.org/10.1175/JPO-D-10-05015.1.
Hogstrom, U., E. Sahlee, A.-S. Smedman, A. Rutgersson, E. Nilsson, K. K. Kahma, and W. M. Drennan, 2015: Surface stress over the ocean in swell-dominated conditions during moderate winds. J. Atmos. Sci., 72, 4777–4795, https://doi.org/10.1175/JAS-D-15-0139.1.
Hristov, T., and J. Ruiz-Plancarte, 2014: Dynamic balances in a wavy boundary layer. J. Phys. Oceanogr., 44, 3185–3194, https://doi.org/10.1175/JPO-D-13-0209.1.
Hristov, T., S. D. Miller, and C. A. Friehe, 2003: Dynamical coupling of wind and ocean waves through wave-induced air flow. Nature, 422, 55–58, https://doi.org/10.1038/nature01382.
Hsu, C.-T., and E. Y. Hsu, 1983: On the structure of turbulent flow over a progressive water wave: Theory and experiment in a transformed wave-following coordinate system. Part 2. J. Fluid Mech., 131, 123–153, https://doi.org/10.1017/S0022112083001263.
Hussain, A. K. M. F., and W. C. Reynolds, 1970: The mechanics of an organized wave in turbulent shear flow. J. Fluid Mech., 41, 241–258, https://doi.org/10.1017/S0022112070000605.
Jacobs, S. J., 1987: An asymptotic theory for the turbulent flow over a progressive water wave. J. Fluid Mech., 174, 69–80, https://doi.org/10.1017/S0022112087000041.
Janssen, P. A. E. M., 1991: Quasi-linear theory of wind-wave generation applied to wave forecasting. J. Phys. Oceanogr., 21, 1631–1642, https://doi.org/10.1175/1520-0485(1991)021<1631:QLTOWW>2.0.CO;2.
Jeffreys, H., 1925: On the formation of water waves by wind. Proc. Roy. Soc. London, 107A, 189–206, https://doi.org/10.1098/rspa.1925.0015.
Kihara, N., H. Hanazaki, T. Mizuya, and H. Ueda, 2007: Relationship between airflow at the critical height and momentum transfer to the traveling waves. Phys. Fluids, 19, 015102, https://doi.org/10.1063/1.2409736.
Kitaigorodskii, S. A., M. A. Donelan, J. L. Lumley, and E. A. Terray, 1983: Wave–turbulence interactions in the upper ocean. Part II: Statistical characteristics of wave and turbulent components of the random velocity field in the marine surface layer. J. Phys. Oceanogr., 13, 1988–1999, https://doi.org/10.1175/1520-0485(1983)013<1988:WTIITU>2.0.CO;2.
Kudryavtsev, V. N., and V. K. Makin, 2004: Impact of swell on the marine atmospheric boundary layer. J. Phys. Oceanogr., 34, 934–949, https://doi.org/10.1175/1520-0485(2004)034<0934:IOSOTM>2.0.CO;2.
Kudryavtsev, V. N., V. K. Makin, and J. F. Meirink, 2001: Simplified model of the air flow above waves. Bound.-Layer Meteor., 100, 63–90, https://doi.org/10.1023/A:1018914113697.
Lighthill, M. J., 1957: The fundamental solution for small steady three-dimensional disturbances to a two-dimensional parallel shear flow. J. Fluid Mech., 3, 113–144, https://doi.org/10.1017/S002211205700052X.
Lighthill, M. J., 1962: Physical interpretation of the mathematical theory of wave generation by wind. J. Fluid Mech., 14, 385–398, https://doi.org/10.1017/S0022112062001305.
Liu, Y., D. Yang, X. Guo, and L. Shen, 2010: Numerical study of pressure forcing of wind on dynamically evolving water waves. Phys. Fluids, 22, 041704, https://doi.org/10.1063/1.3414832.
Mastenbroek, C., V. K. Makin, M. H. Garat, and J. P. Giovanangeli, 1996: Experimental evidence of the rapid distortion of turbulence in the air flow over water waves. J. Fluid Mech., 318, 273–302, https://doi.org/10.1017/S0022112096007124.
Miles, J. W., 1957: On the generation of surface waves by shear flows. J. Fluid Mech., 3, 185–204, https://doi.org/10.1017/S0022112057000567.
Miles, J. W., 1993: Surface-wave generation revisited. J. Fluid Mech., 256, 427–441, https://doi.org/10.1017/S0022112093002836.
Monin, A., and A. Obukhov, 1954: Basic laws of turbulent mixing in the surface layer of the atmosphere. Tr. Geofiz. Inst., Akad. Nauk SSSR, 24, 163–187.
Phillips, O. M., 1957: On the generation of waves by turbulent wind. J. Fluid Mech., 2, 417–445, https://doi.org/10.1017/S0022112057000233.
Plant, W. J., 1982: A relationship between wind stress and wave slope. J. Geophys. Res., 87, 1961–1967, https://doi.org/10.1029/JC087iC03p01961.
Portilla, J., F. J. Ocampo-Torres, and J. Monbaliu, 2009: Spectral partitioning and identification of wind sea and swell. J. Atmos. Oceanic Technol., 26, 107–122, https://doi.org/10.1175/2008JTECHO609.1.
Smedman, A., U. Hogstrom, E. Sahlée, W. M. Drennan, K. K. Kahma, H. Pettersson, and F. Zhang, 2009: Observational study of marine atmospheric boundary layer characteristics during swell. J. Atmos. Sci., 66, 2747–2763, https://doi.org/10.1175/2009JAS2952.1.
Smith, J. A., and K. F. Rieder, 1997: Wave induced motion of FLIP. Ocean Eng., 24, 95–110, https://doi.org/10.1016/0029-8018(96)00008-X.
Snyder, R. L., F. W. Dobson, J. A. Elliott, and R. B. Long, 1981: Array measurements of atmospheric pressure fluctuations above surface gravity waves. J. Fluid Mech., 102, 1–59, https://doi.org/10.1017/S0022112081002528.
Soloviev, Y. P., and V. N. Kudryavtsev, 2010: Wind-speed undulations over swell: Field experiment and interpretation. Bound.-Layer Meteor., 136, 341–363, https://doi.org/10.1007/s10546-010-9506-z.
Sullivan, P. P., J. C. McWilliams, and C.-H. Moeng, 2000: Simulation of turbulent flow over idealized water waves. J. Fluid Mech., 404, 47–85, https://doi.org/10.1017/S0022112099006965.
Sullivan, P. P., J. B. Edson, T. Hristov, and J. C. McWilliams, 2008: Large-eddy simulations and observations of atmospheric marine boundary layers above nonequilibrium surface waves. J. Atmos. Sci., 65, 1225–1245, https://doi.org/10.1175/2007JAS2427.1.
Sullivan, P. P., J. C. McWilliams, and E. G. Patton, 2014: Large-eddy simulation of marine atmospheric boundary layers above a spectrum of moving waves. J. Atmos. Sci., 71, 4001–4027, https://doi.org/10.1175/JAS-D-14-0095.1.
Townsend, A. A., 1972: Flow in a deep turbulent boundary layer over a surface distorted by water waves. J. Fluid Mech., 55, 719–735, https://doi.org/10.1017/S0022112072002101.
Van Duin, C. A., and P. A. E. M. Janssen, 1992: An analytic model of the generation of surface gravity waves by turbulent air flow. J. Fluid Mech., 236, 197–215, https://doi.org/10.1017/S0022112092001393.
Veron, F., G. Saxena, and S. K. Misra, 2007: Measurements of the viscous tangential stress in the airflow above wind waves. Geophys. Res. Lett., 34, L19603, https://doi.org/10.1029/2007GL031242.
Veron, F., W. K. Melville, and L. Lenain, 2008: Wave-coherent air–sea heat flux. J. Phys. Oceanogr., 38, 788–802, https://doi.org/10.1175/2007JPO3682.1.
Veron, F., W. K. Melville, and L. Lenain, 2009: Measurements of ocean surface turbulence and wave–turbulence interactions. J. Phys. Oceanogr., 39, 2310–2323, https://doi.org/10.1175/2009JPO4019.1.
WAFO Group, 2000: WAFO. Department of Mathematical Statistics, Centre for Mathematical Sciences, Lund University, http://www.maths.lth.se/matstat/wafo/.
Wilczak, J., S. Oncley, and S. Stage, 2001: Sonic anemometer tilt correction algorithms. Bound.-Layer Meteor., 99, 127–150, https://doi.org/10.1023/A:1018966204465.
Wu, L., A. Rutgersson, and E. Nilsson, 2017: Atmospheric boundary layer turbulence closure scheme for wind-following swell conditions. J. Atmos. Sci., 74, 2363–2382, https://doi.org/10.1175/JAS-D-16-0308.1.
Wu, L., T. Hristov, and A. Rutgersson, 2018: Vertical profiles of wave-coherent momentum flux and velocity variances in the marine atmospheric boundary layer. J. Phys. Oceanogr., 48, 625–641, https://doi.org/10.1175/JPO-D-17-0052.1.
Yang, D., and L. Shen, 2009: Characteristics of coherent vortical structures in turbulent flows over progressive surface waves. Phys. Fluids, 21, 125106, https://doi.org/10.1063/1.3275851.
Yang, D., and L. Shen, 2010: Direct-simulation-based study of turbulent flow over various waving boundaries. J. Fluid Mech., 650, 131–180, https://doi.org/10.1017/S0022112009993557.
Yang, D., C. Meneveau, and L. Shen, 2013: Dynamic modelling of sea-surface roughness for large-eddy simulation of wind over ocean wavefield. J. Fluid Mech., 726, 62–99, https://doi.org/10.1017/jfm.2013.215.
Young, W. R., and C. L. Wolfe, 2014: Generation of surface waves by shear-flow instability. J. Fluid Mech., 739, 276–307, https://doi.org/10.1017/jfm.2013.617.
The “watershed” algorithm, which treats each directional spectrum like a topographic map, with the power spectral density of each point representing its height, finds the lines that run along the tops of ridges and separates adjacent drainage basins that correspond to adjacent partitions of the wave field.
Note that the definition is different from the one used in Hanson and Phillips (2001).