66
Salt intrusion at a submarine spring in a fringing reef lagoon Sabrina M. Parra and Arnoldo Valle-Levinson, Engineering School of Sustainable Infrastructure and Environment, University of Florida, Gainesville, Florida, USA Ismael Mariño-Tapia, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional, Merida, Yucatan, Mexico Cecilia Enriquez, Universidad Nacional Autónoma de México, Sisal , Yucatán, México Corresponding author: S. M. Parra, Engineering School of Sustainable Infrastructure and Environment, University of Florida, Weil Hall 365, Gainesville, Florida 32611, United States ( [email protected]) Key Points: Salt intrusion was observed during high syzygy tides Salinities and temperatures suggest jet is connected to a stratified cave A wave setup of 5 cm significantly reduced jet velocities and turbulence Keywords 1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

 · Web viewSalt intrusion at a submarine spring in a fringing reef lagoon Sabrina M. Parra and Arnoldo Valle-Levinson, Engineering School of Sustainable Infrastructure and Environment,

  • Upload
    vodat

  • View
    214

  • Download
    1

Embed Size (px)

Citation preview

Salt intrusion at a submarine spring in a fringing reef lagoon

Sabrina M. Parra and Arnoldo Valle-Levinson, Engineering School of Sustainable Infrastructure

and Environment, University of Florida, Gainesville, Florida, USA

Ismael Mariño-Tapia, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico

Nacional, Merida, Yucatan, Mexico

Cecilia Enriquez, Universidad Nacional Autónoma de México, Sisal, Yucatán, México

Corresponding author: S. M. Parra, Engineering School of Sustainable Infrastructure and

Environment, University of Florida, Weil Hall 365, Gainesville, Florida 32611, United States

([email protected])

Key Points:

Salt intrusion was observed during high syzygy tides

Salinities and temperatures suggest jet is connected to a stratified cave

A wave setup of 5 cm significantly reduced jet velocities and turbulence

Keywords

Salt intrusion, submarine groundwater discharge, turbulent kinetic energy, karst spring, wave

setup

Index terms: 4528 (Fronts and jets), 4568 (Turbulence), 4217 (Coastal processes), 4560 (surface

waves and tides), 4546 (nearshore processes),

Running header: Salt intrusion at a submarine spring

6 experts for review: Patricia BeddowsMatthew ReidenbachAmatzia GeninFiona WhitakerSamantha Smith

1

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20212223242526

Acknowledgments

We thank Edgar Escalante, Francisco Ruiz and Roberto Iglesias from the Puerto Morelos

station of the Instituto de Ciencias de Mar y Limnología of Universidad Nacional Autónoma de

México for providing bathymetric and meteorological data and for the support received during

fieldwork, and Emanuel Sanchez for the support in field work. SMP gratefully acknowledges

support from the National Science Foundation’s Bridge to the Doctorate program. This research

was funded by National Science Foundation project OCE-0825876 and Consejo Nacional de

Ciencia y Tecnología, México project #84847. We would like to thank Aslak Grinsted for

allowing us the use of his MATLAB wavelet package.

2

27

28

29

30

31

32

33

34

35

Abstract (246 words)

Variations in discharge and turbulent kinetic energy (TKE) were studied at a point-source

submarine groundwater discharge (SGD), within a fringing reef lagoon, from quadrature (neap)

to syzygy (spring) tides. The principal factors affecting discharge and TKE variations were tides

and waves. Field data indicated discharge and TKE varied with high and low tides, and with

quadrature and syzygy. Maximum discharge and TKE values were observed during low tides

when the hydrostatic pressure over the jet was minimal, while the lowest discharge and TKE

values were found at high tides. Syzygy tides produced consistent saltwater intrusion during high

tides, while quadrature tides produced the greatest TKE values. In general, as the discharge

intensified during low tides, jet temperatures decreased while jet salinities increased. Decreasing

temperatures suggest that waters originated further within the aquifer. Increasing salinities of the

same discharge suggested mixing of aquifer and seawater. Therefore, it is proposed that the jet

conduit is connected to a chamber that is stratified with seawater below brackish water. The

greatest subtidal discharge occurred during quadrature tides. Syzygy produced low subtidal

discharge driven by flow reversals (flow into the aquifer) observed throughout syzygy high tides

in conjunction with the peak wave setup (>5 cm) observed during a storm. While tides were the

primary driving force of the discharge, waves played a non-negligible role. Wave effects on the

discharge were most evident during syzygy high tides combined with a storm, when the subtidal

spring discharge was weakest and salt intrusion developed.

3

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

1 Introduction

Groundwater reserves in coastal regions of the world are being threatened by sea level

rise, pollution, and ever-increasing demands for freshwater. Of special concern are coastal

aquifers because they are hydraulically connected to coastal ocean waters through the seabed.

These connections pose a fragile balance between the aquifer’s hydraulic pressure head and that

of the sea. Higher aquifer water pressure heads relative to the sea level drive submarine

groundwater discharges (SGDs) through bed sediments [Moore, 1996]. These SGDs can be

either slow diffuse sources (on the order of cm day-1) or fast point sources (on the order of m s-1)

of fresh or brackish waters [Zektser and Everett, 2000]. Diffuse (seepage) sources are associated

with low permeability bed sediments such as sandy soils {e.g.[Uchiyama et al., 2000], while

point sources (submarine springs/jets) often occur in karst topography [e.g. Valle-Levinson et al.,

[2011] and on volcanic terrain through lava tubes [e.g. Zektser and Everett, [2000];[Peterson et

al., 2009] ]. Diffuse SGDs have been studied broadly and are relevant to coastal ecosystems

because of their contribution of dissolved nutrients, both natural and anthropogenic [Taniguchi

et al., 2002; Santos et al., 2008; Moore, 2010].

Extensive study of submarine springs, on the other hand, has only occurred recently [e.g.,

[Peterson et al., 2009; Valle-Levinson et al., 2011; Expósito-Díaz et al., 2013]. In karst

topography, SGD can flow through submarine springs that provide a hydraulic connection

between seawater and aquifer water. Karst topography is found in all continents except

Antarctica. It is characterized by limestone or dolomite bedrock, where groundwater flows can

dissolve channels and caves into the predominantly calcium carbonate soil [Fleury et al., 2007].

Coastal karst aquifers also provide necessary freshwater resources to coastal communities around

the world, which are being threatened by saltwater intrusion [e.g. Fitterman, 2014]. Saltwater

4

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

intrusion is defined as “the encroachment of saline water into fresh ground water regions in

coastal aquifer settings” [Werner and Simmons, 2009].

Jet and seepage SGD in karst topography depends on the gradient between the

piezometric level and the sea level, and the contrast in densities between the aquifer water and

seawater [Valle-Levinson et al., 2011]. Some combination of these factors is necessary for salt

intrusion to occur. Conditions that affect salt intrusion from the sea include: increased sea levels

caused by storm surges [Hu et al., 2006; Smith et al., 2008], waves and wave setup [Li et al.,

1999; Parra et al., 2014], sea level rise caused by climate change [Werner and Simmons, 2009;

Vera et al., 2012; Gonneea et al., 2013], and tides [Taniguchi, 2002; Valle-Levinson et al., 2011;

Expósito-Díaz et al., 2013]. Factors from land include: seasonal variations in groundwater

recharge rates [Vera et al., 2012; Carretero et al., 2013], groundwater pumping rates from

coastal communities’ consumption [Santos et al., 2009; Ferguson and Gleeson, 2012; Carretero

et al., 2013] and the composition of sediments and porosity of coastal regions [Bakhtyar et al.,

2012]. Valle-Levinson et al., [2011] presented a few instances of salt intrusion into a point-

source SGD occurring only during the highest tides (high tide during syzygy tides). Vera et al.,

[2012] explained a more complex relationship between spring discharge and saltwater intrusion,

suggesting that lower frequency (subtidal) processes are the most influential to salt intrusion.

Additionally, groundwater flows were dominated by tides when the tidal amplitude was greatest,

while barometric pressure changes were more influential than tides during small tidal amplitudes

and relatively calm seas.

While tidal variations in submarine springs have been investigated comprehensively, the

effect of waves through wave pumping have only been quantified for seepage SGD, where wave

pumping refers to the pressure oscillations generated by pressure gradients beneath wave crests

5

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

and troughs [McLachlan and Turner, 1994]. Lab experiments showed that wind waves can

“enhance pore water and solute exchange” of diffuse SGD [Precht and Huettel, 2003]. Similar

results were obtained by numerical model results [Bakhtyar et al., 2012; Li et al., 1999]. Wave

setup has been studied in numerical models [Li et al., 1999] and with field observations [Parra

et al., 2014]. Parra et al., [2014] found that wave setup within a fringing reef lagoon acted to

suppress the discharge from a submarine spring. However, the effects of shorter frequency

waves, like wind-waves, on point-source SGD have not been studied with field observations.

Submarine groundwater discharges are known to provide nutrients and pollutant

concentrations to coastal waters and ecosystems, especially in coastal karst aquifers [Hamner

and Wolanski, 1988; Null et al., 2014; Li et al., 1999]. Ecosystems that are especially vulnerable

to and dependent on SGD-derived chemicals are coral reef ecosystems that exist in low nutrient

environments [Hamner and Wolanski, 1988]. The nutrient concentrations and subsequent

nutrient uptake by reef communities is known to depend on turbulent mixing associated with

waves and currents [Patterson et al., 1991]. However, increased pollutant transport originating

from SGD has also been shown to impact the health of coral reef and beds of seagrass worldwide

[e.g. [Metcalfe et al., 2011; Null et al., 2014; Chérubin et al., 2008]. Turbulence emanating from

submarine springs, therefore, could play an important role in nutrient and pollutant mixing, and

advection within coral reef ecosystems. Mixing can act to reduce pollutant and nutrient

concentrations to admissible levels. Understanding the levels of turbulence from point-source

SGD is needed before determining its role on the ecosystems. However, the turbulence

associated with submarine springs has been scantly studied [Parra et al., 2014; Expósito-Díaz et

al., 2013]. Expósito-Díaz et al., [2013] estimated the vertical structure of turbulent properties of

a point-source SGD in a tropical estuary in the Yucatan Peninsula. Results showed maximum

6

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

zonal Reynolds stress components, turbulent kinetic energy (TKE) production, and vertical eddy

viscosity at low tides when the greatest discharge rates were observed. Similar results showing

maximum TKE and discharge at low tides were also observed by Parra et al., [2014] at a

submarine spring located within a fringing coral reef lagoon of the Yucatan Peninsula. However,

the two previous studies have only collected 4 and 3 days of data, thus only considering tidal

variations of the discharge. Variations in turbulence at a point-source SGD from quadrature to

syzygy tides have yet to be determined.

In addition to tides and turbulence, previous studies have observed variations in seepage

SGD from syzygy through quadrature tides [Kim and Hwang, 2002; Taniguchi, 2002; Vera et

al., 2012]. Semi-monthly fluctuations in seepage SGD were more significant than originally

thought, with greater variations during syzygy or spring tides [Kim and Hwang, 2002].

Additionally, diurnal oscillations in seepage were more significant during syzygy tides than

quadrature or neap tides. Similarly, Taniguchi, [2002] found that seepage SGD varied semi-

monthly, with maximum discharge rates during syzygy tides and minimum rates during

quadrature tides.

The aim of this study is to describe fortnightly variations at a point source SGD, and to

understand the conditions of tides and waves that favor salt intrusion into the aquifer. This study

extends previous investigations by describing fortnightly variations of TKE and by determining

the additional influence of waves and wave setup on the spring discharge.

2 Methods

2.1 Study site

This study was carried out in the Puerto Morelos coral reef lagoon, located 30 km south

of Cancun on the northeast coast of the Yucatan Peninsula (Figure 1). The Yucatan Peninsula’s

7

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

watershed topography is karstic and composed of highly permeable limestone. Because of its

high permeability, rainfall quickly seeps into the aquifer. This explains the absence of surface

rivers in the region [Merino et al., 1990]. Because of the high permeability in the aquifer, the

water table has hydraulic gradients ranging from 7 to 100 mm km-1 and is known to be controlled

by sea level variations and recharge from precipitation [González-Herrera et al., 2002; Beddows

et al., 2007]. The peninsula is peppered with an abundant number of sinkholes (locally known as

cenotes), which make the unconfined aquifer visible from the surface [Vera et al., 2012]. Aquifer

waters flow within a massive network of underground caves and conduits carved through the

limestone. Eventually, aquifer waters flow into the ocean through submarine spring and seepage

discharges all along the Yucatan coast, driven by a hydraulic pressure gradient between the

aquifer and the sea. The Yucatan coastal SGDs are dominated by spring discharges, estimated at

between 78-99 % of total SGD. The Puerto Morelos lagoon has an estimated spring SGD

contribution of 78.5 %, with the remaining 21.5 % coming from seepage at the beach face [Null

et al., 2014]. Therefore, submarine springs are of great importance to the region.

Two distinct weather periods are identified in the Yucatan Peninsula region: dry and wet

(Figure 2). The dry period extends from March through June and the wet period is generally

from July through October. Between December and February, brief (2-5 days) cold fronts

characterized by northerly winds with light rains, locally known as nortes, are common. The

average rainfall is approximately 1,300 mm year-1. The time series of this study was collected

during the wet period, indicating that large aquifer recharge rates (estimated at ~25,300 hm3 year-

1) occurred throughout the watershed during this time period [Vera et al., 2012].

The Puerto Morelos lagoon has a mean water depth of ~3.5 m with a maximum depth of

8 m. It is enclosed by fringing coral reefs that are disrupted by two major and one minor inlet.

8

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

The northern inlet, locally known as Boca Grande, is 1200 m wide with a depth of 6 m, while

the central inlet, locally known as Boca Chica, is 300 m wide with a depth of 6 m. The

southern inlet is a navigation channel 400 m wide with a dredged depth of 8 m [Coronado et al.,

2007]. This lagoon is influenced by dominant semidiurnal microtides (≤0.3 m tidal range) and

persistent trade winds that drive onshore wind-waves. The reef region includes shallow

submerged coral banks which can be uncovered during low tides although it is a microtidal

region. Under normal wave conditions it is influenced by incident significant wave heights of 0.8

± 0.4 m [Coronado et al., 2007].

Puerto Morelos lagoon circulation is predominantly influenced by wind forcing and

thermohaline circulations caused by SGD and solar heat fluxes during periods of minimal wave

activity [Enriquez et al., 2012], and wind-waves (wave setup) during periods of energetic wave

conditions [Coronado et al., 2007]. Wave-driven circulation produces outflowing currents >0.5

m s-1 through the inlets, which ensures flushing and a well-mixed water column in the shallow

lagoon [Coronado et al., 2007].

The lagoon is punctuated on its bottom by numerous (10 to 15) buoyant jet discharges,

some of which can feature distinguishable surface expressions during low tides. The submarine

spring discharge of interest in this study is the Pargos spring, named after the large number of

snappers that used to congregate at this location. The spring has estimated discharge rates of

approximately 0.4 m3 s-1. It is located 300 m from shore and 1,200 m west of the central inlet

(Figure 1). The bathymetry at the spring displays an opening to the spring vent that is ~2.5 m in

diameter and extends 2 m below the lagoon floor (Figure 3). The spring vent is ~1.2 m diameter

conduit located at the bottom of the opening at an approximate depth of 7 m.

9

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

2.2 Data collection

In order to study the effects of the variability of tides from quadrature to syzygy and the

influence of wind-waves on the submarine spring discharge, data were collected between 19-30

July 2011, with four instruments fixed on the bed of the lagoon. Instruments were moored at

three locations: the forereef, the middle of the lagoon and at the mouth of the jet associated with

Pargos spring (Figure 1). Hydrographic recorders were moored at the spring and in mid-lagoon,

and single-point acoustic velocimeters were deployed at the spring and the forereef.

A 6 MHz Nortek Vector ADV was moored at the forereef at 2 m depth. The ADV

recorded three-dimensional velocity bottom measurements and pressure in hourly bursts of 2048

measurements at 2 Hz (~17.1 min). The water pressure measurements were used to determine

significant wave heights and corresponding spectral energy at the forereef (Figure 4).

A 6 MHz Nortek Vector acoustic Doppler velocimeter (ADV) and a Schlumberger

conductivity, temperature and depth Diver recorder (CTD) were moored at the mouth of Pargos

spring (Figure 1, 3) and gathered data simultaneously. The ADV was moored in the center of the

vent and oriented downward, while the CT was located ~25 cm above the sampling volume of

the ADV (to avoid disturbing the flow field). The ADV recorded three-dimensional velocity

point measurements and pressure in hourly bursts of 8192 measurements at 8 Hz (~17.1 min).

The water pressure bursts were used to calculate the spectral energy of the signal at the spring

(Figure 5). The three-dimensional velocity measurements were used to estimate TKE at the jet

via velocity anomalies from burst averages (Figure 6). The CTD logged point measurements of

conductivity, temperature and depth from the jet, every 10 minutes (Figure 6). Salinity was

estimated using the CTD data following the thermodynamics equation of seawater (TEOS-10)

[Millero et al., 2008].

10

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

In addition, a Sea-Bird SBE-37 conductivity and temperature recorder (CT) was deployed

on the bottom, in the middle of the lagoon (Figure 1). The CT logged measurements of

conductivity, temperature and depth every 10 minutes (Figure 6). Salinity was estimated using

the CT data following the thermodynamic equation of seawater (TEOS-10) [Millero et al., 2008].

Furthermore, hourly wind speeds and direction (Figure 4) were obtained from a meteorological

station located on shore at the Universidad Nacional Autónoma de México (UNAM) Academic

Unit.

2.3 Data processing

Incident waves at the forereef were derived from water pressure measurements (2 Hz)

obtained with an ADV using spectral analysis techniques. Spectral analysis was performed on

the water pressure data for each burst using a fast Fourier transform (FFT). In order to reduce

noise, Hanning windowing was applied to each burst. The water pressure spectra were used to

estimate significant wave heights as:

H s=4 √m0 (1)

where m0 is the zeroth moment of the spectrum based on the energy between 0.05 and 0.5 Hz. In

addition, the peak wave period, T p, was calculated based on the spectral peaks.

In this study turbulence observations were used to quantify the effects of tides and waves

on the energy of the spring discharge. Turbulent kinetic energy, k , was estimated using the

normal components of the Reynolds stress tensor as follows [Pope, 2000; Monismith, 2010]:

k=⟨ui

' 2 ⟩2

=⟨u' 2 ⟩+⟨ v' 2 ⟩+⟨ w '2 ⟩

2(2)

where the prime ( ' ) denotes velocity anomalies from the temporal mean. For example, for the x-

component the anomalies are given by u'=u−⟨U ⟩, where the brackets denote the burst mean.

11

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

The velocities u, v and w represent the instantaneous time series of the east, north and vertical

velocity components, respectively. TKE was calculated to seek relationships between water

surface variations, η, over the spring and spring discharge variations. Wave-turbulence

decomposition in post-processing was not performed as recommended by [Bricker and

Monismith, 2007]. They found that wave bias in turbulent velocity data was related to

uncertainties in instrument orientation (pitch, roll and compass). They recommended performing

wave-turbulence decomposition on data from instruments that do not possess compass, and

pitch/roll sensors. Therefore, data obtained from a velocimeter equipped with reliable compass

and tilt/roll sensors, like the Nortek Vector, do not require wave-turbulence decomposition

because the data produced more reliable power spectra than any method of interpolation or

filtering of the wave frequencies.

Further relationships between water surface variations and spring discharge were further

explored using wavelet analysis techniques. The continuous wavelet transform was used for

signal processing of both stationary and non-stationary datasets [Kang and Lin, 2007]. It is

similar to spectral analysis, in that it describes the contributions of oscillations at different

frequencies. While spectral analysis only considers the total contributions of an assumed-to-be-

stationary dataset, wavelet analysis is “able to determine both the dominant modes of variability

and how those modes vary in time” [Torrence and Compo, 1998]. In order to do this, wavelet

analysis decomposes a non-stationary time series into both time and frequency space. This

decomposition allows the study of both global and local processes, and therefore permits the

analysis of intermittent and aperiodic processes [Baas, 2006]. The continuous wavelet transform

was used to identify statistically significant variations in oscillations at different frequencies of

significant wave height, sea level, jet discharge and TKE throughout the time series. In this

12

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

study, a Morlet wavelet (with dimensionless frequency of 6) was used to examine the behavior in

the time-frequency domain of the variables [Torrence and Compo, 1998; Grinsted et al., 2004].

Squared coherences between water level variations and spring velocities were obtained

for a range of 0.01 to 1 Hz. The squared-coherence, γ122 ( f k ), determines the relationship between

two data sets using the cross spectral,G12 ( f k ), and auto-spectral densities,G11 ( f k ) and G22 ( f k ), as

follows [Emery and Thomson, 2001]:

γ122 ( f k )=

|G12 ( f k )|2

G11 ( f k ) G22 ( f k ) (3)

where f k are the frequencies of the squared-coherence. Considering the innate random noise

found and in order to reliably estimate the squared-coherence between two datasets, only the

coherence amplitudes that were above a 99 % confidence level were considered significant. To

reduce noise of the squared-coherence, a Hanning window was applied to the spectral densities.

3 Results

3.1 Winds and waves

During the measurement period, winds exhibited weak, moderate and storm conditions.

Weak winds were predominantly westward at ≤5 m s-1 between 19-23 July 2011 (Figure 4a).

Moderate winds were southwestward on 23-27 and 29-30 July 2011 with increased speeds of ~7

m s-1. Finally, north-northwestward winds with speeds up to 10 m s-1 occurred during a passing

storm on 28 July.

The lagoon’s circulation is known to be dominated by wind-waves (Coronado et al.

2007). The smallest significant wave heights observed at the forereef (~0.4 m) were recorded

during weak winds (Figure 4b). Increased significant wave heights (0.4-0.7 m) were observed as

winds strengthened during moderate conditions. Lastly, significant wave heights reached their

13

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

peak of up to 1 m on 28 July 2011 during the storm. On the other hand, peak wave periods

ranged mainly between 6 and 8 s (seconds) throughout the time series (Figure 4b).

In addition to looking at the wave heights and periods, wave energy at different

frequencies was also determined with a spectral analysis. Wave energy presented minimum

energy during the first four days and maximum energy on 28 July (Figure 4c). During weak

waves, energy was concentrated at frequencies between 0.1 and 0.3 Hz, with minimal activity at

other frequencies. As wave heights increased between 23-26 July, wave energy also increased

with a broader frequency range (0.09-0.35 Hz). Intermittent energy was also detected at lower

frequencies (<0.08 Hz), related to infragravity waves present in the lagoon and forereef (Fig 4c,

5b). Between 28-30 July as the storm passed, wave energy was at its maximum with the greatest

frequency range (0.08-0.4 Hz). Concurrently, the largest infragravity wave energy was observed.

Waves observed at the forereef (Figure 4c) had a very similar spectral signature to those at the jet

(Figure 5b) but were ~4 times greater in magnitude. The spectra of the water surface bursts at the

jet displayed waves between 0.1 and 0.3 Hz, corresponding to wind waves. However, no distinct

period of oscillation greater than the infragravity waves was observed for significant wave

heights at Pargos throughout the time series (Figure 7a).

The measurement period extended from quadrature or neap tides on 23 July (<0.1 m

range) to syzygy or spring tides on 29-30 July (~0.35 m range; Figure 5a). Throughout the time

series, semidiurnal tides dominated the water level variability. Clear tidal variations were also

observed in the jet’s velocities, with magnitudes ranging between -0.25 and 0.33 m s-1 (Figure

5c), producing sinusoidal semidiurnal oscillations. These oscillations featured maximum

velocities throughout low tides and sudden drops in velocity at high tides, with the exception of

neap tides on day 22. The observed periodicity of the jet followed both syzygy-quadrature and

14

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

high-low tidal cycles. Peak discharge was greater during low syzygy tides than at low quadrature

tides, while the range of jet velocities was smaller during quadrature than syzygy. Quadrature

produced velocities ranging between 0.05 and 0.25 m s-1, with increasing semidiurnal oscillations

as the tidal range increased from quadrature to syzygy. Starting on 25 July and through the end

of the time series high tides produced negative jet velocities that signaled flow reversals or salt

intrusion. Here, lagoon waters entered through the spring into the aquifer system. On the other

hand, during the same time period low tides produced jet velocities >0.3 m s-1. The semidiurnal

oscillations of the jet discharge also appeared in the burst spectral analysis, with relatively high

values throughout the range of spectral frequencies at low tides when the discharge was the

greatest (Figure5d). Conversely, when discharge was low at high tides, the spectral signal

decreased at higher frequencies (>0.1 Hz). The squared-coherence analysis between the water

surface and spring velocities (Figure 5e) showed statistically significant coherence (within the 99

% confidence limit depicted as the black contour line) during periods of negative vertical

(downward) velocities (high tides on 26-29 July). Statistically significant coherences between

water surface and spring velocities on 26-29 July developed at the wind-wave and infragravity

wave frequencies, especially during the storm event of 28 July, revealing that wave oscillations

were observed within the jet discharge. This is known as wave pumping.

To observe the effects of waves on the lagoon circulation and spring discharge, the wave

setup was calculated as the difference between the demeaned water levels at Pargos spring and

the forereef (Figure 6b). Because the circulation in the lagoon is known to be dominated by

waves, it was expected that wave setup depended on significant wave heights. This indeed was

the case, especially during high wave events and could be explained with the following

relationship:

15

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

Δ η=0.25 H s , jet (4)

where Δ η is the wave setup between Pargos and the forereef and H s is the significant wave

height measured at the jet (r2=0.51). A greater agreement was observed after 22 July (r2=0.80),

suggesting that setup was correlated with waves only when significant wave heights became >0.4

m. However, when waves at the forereef were < 0.4 m, setup was small and dominated by other

factors.

3.2 Jet turbulent kinetic energy (TKE) and velocities

Semidiurnal variations were also evident in the jet’s turbulent kinetic energy with an

inverse relationship to tides (Figure 6c). Semidiurnal oscillations in TKE occurred throughout

most of the time series. Quadrature tides (22-24 July) allowed development of the highest values

of TKE per unit mass (up to 0.04 m2 s-2), when jet velocities remained ~0.2 m s-1 and significant

wave heights at the forereef were at their lowest (~0.4 m). Subsequent TKE maxima steadily

decreased after their peak on 23 July and as syzygy tides increased their influence on the

discharge by damping TKE development.

3.3 Jet temperature and salinity

While jet velocities and TKE oscillated inversely to the semidiurnal tides, lagoon bottom

temperatures (away from the spring’s influence) showed a diurnal pattern consistent with

atmospheric heat fluxes (Figure 6d). Between 27-29 July 2011, however, lagoon bottom

temperatures did not show diurnal oscillations, but a cooling trend most likely caused by cloud

cover from the storm. On the other hand, lagoon bottom salinities (away from the spring’s

influence) were relatively constant, starting out near 36 g kg-1 and steadily decreasing to 35.3 g

kg-1 toward the end of the time series. This could be attributed to any or a combination of the

16

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

following: stronger water exchange between the lagoon and the sea, decreased evaporation from

increased cloud cover, lower air temperatures, and precipitation caused by the storm (Figure 6e).

During this time of year, aquifer water temperatures are lower than ambient lagoon

temperatures, therefore jet temperatures acted as a tracer of aquifer water discharging into the

lagoon. In contrast to the ambient lagoon temperature variations, jet discharge temperatures were

clearly semidiurnal ranging from 27 to 32 °C (Figure 6d). Semidiurnal oscillations in jet

temperatures differed from quadrature to syzygy tides, with a clear transition in variability

occurring on 24 July. Throughout, high tides coincided with decreased jet flows and relatively

high water temperatures at the jet. Conversely, low tides produced increased jet discharge and

low water temperatures.

Similarly to jet temperatures, jet salinities showed semidiurnal oscillations with low

values during periods of high jet velocities (low tides). Spring salinities were generally lower

than ambient lagoon bottom salinities and also served as a conservative tracer of aquifer waters.

At Pargos spring, salinities ranged between 28 and 36 g kg-1 with clear differences between high

and low tides, and between quadrature and syzygy tides. Spring discharges were brackish,

indicating a subterranean estuary where mixing of recirculated seawater was driven by ‘tidal

pumping’ [Moore, 1999]. Tidal pumping, as defined by Santos et al., [2009], is separated into 3

components: (1) hydraulic gradient between the aquifer and the sea driving outflow during low

tides, (2) seawater recirculation or saltwater intrusion at high tide, and (3) current-driven

recirculation outside the spring. The first two components can be grouped as discharge driven by

oscillations in the water level, with the smallest sea level range at quadrature and greatest range

at syzygy tides.

17

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

4 Discussion

Jet discharge was dominated by tides, with semidiurnal oscillations in jet velocities that

varied from quadrature to syzygy (Figure 7b, c). Maximum jet velocities occurred during low

tides as the water pressure over the jet was least. Minimum jet velocities developed during high

tides, consistent with previous investigations [Parra et al., 2014; Expósito-Díaz et al., 2013;

Valle-Levinson et al., 2011]. Similarly, low tides displayed peaks in TKE while high tides

damped it (Figure 7d). However, a marked difference was observed between quadrature and

syzygy tides. Quadrature tides displayed less variability in discharge and greater TKE than

syzygy tides, which could be attributed to the smaller tidal range in quadrature compared to

syzygy. Smaller tidal ranges in quadrature produced smaller ranges in jet velocities possibly

caused by the fast response of the aquifer to changes in water level. Therefore, an increase in

tidal range produced a subsequent increase in the jet velocity range.

4.1 Salt Intrusion

During high syzygy tides, reversals in jet velocity driven by tidal pumping were observed

from 25 July on to the end of the observation period (Figure 6c). These reversals were associated

with lagoon water flowing into the aquifer. Additionally, salinity and temperature signals at the

jet indicated warm and salty lagoon water entered the spring (concurrent with downward jet

velocities), which also confirmed the occurrence of salt intrusion at high tides (Figure 6d, e).

Instances of salt intrusion have also been observed at submarine springs discharging at the

northern Yucatan Peninsula coast [Valle-Levinson et al., 2011; Vera et al., 2012]. Valle-

Levinson et al., [2011] reported salt intrusion only during the highest water levels, therefore

attributing these cases to increased water pressure over the spring. However, the results in Puerto

Morelos lagoon suggested that the observed saltwater intrusion was caused by a more complex

relationship between tidal pumping and tidal amplitudes as explained below.

18

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

Flow reversals into the aquifer began on midday 25 July and continued through the end

of the time series (Figure 6c; dashed green line in Figure 6a). These flow reversals depended on

the hydraulic pressure gradient and density differences between the sea and aquifer [Valle-

Levinson et al., 2011]. This relationship can be approximated by using a modified form of the

Bernoulli equation that takes into account an energy balance of the system and includes frictional

losses, assuming the flow within the aquifer was comparable to pipe flow. The friction term

should be included because the flow within the aquifer’s conduits and cave formations is

considered turbulent [Bauer-Gottwein et al., 2011; Beddows et al., 2007]. The jet discharge, U2,

could be approximated as follows:

g ρ1 h1+ρ1 U1

2

2=g ρ2h2+

ρ2U 22

2+ρ jet hL (5)

This equation represents the energy (per unit volume) balance between a cenote (sinkholes) on

land (left hand side) and the submarine spring (right hand side). Friction losses within the

aquifer, hL (m2 s-2), could be defined using the Darcy-Weisbach equation [Okiishi et al., 2006] :

hL=fLU2

2

2 D(6)

where f is the friction factor estimated at 0.016, L is the distance between the cenote and spring

estimated at 2 km and D is the diameter of the spring opening estimated at 1 m. Assuming U 1 at

the cenote is negligible, an inland water level of h1 = 6 cm above sea level, a freshwater density

of ρ1 = 1000 kg m-3, lagoon water density of ρ2 = 1025 kg m-3, and using the sea level

oscillations over the jet as h2, then jet velocities are reasonably predicted by

U 22=

2 g (ρ1 h1−ρ2 h2)

ρ2(1+ fLD

)(7)

19

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

as observed in Figure 8b (r2=0.75). Negative velocities were assumed when the right hand side of

the equation became negative. The resulting hydraulic gradient of 30 mm km-1 (¿h1/ L) is within

ranges of 7-10 mm km-1 and 50-100 mm km-1 provided by Beddows et al., [2007] and González-

Herrera et al., [2002], respectively. Jet velocities for hydraulic gradients of 5 and 100 mm km-1

are also plotted, showing the expected discharge during dry and wet periods, respectively. The

dry period jet velocity estimates (red) show average outflow of 0.37 throughout with no periods

of salt intrusion. The wet period jet velocity estimates (green) show predominant salt intrusion

during most high tides, except at the peak of quadrature tides on day 22. This relationship for jet

velocities suggests that density differences, hydraulic gradient and friction losses play an

important role in the dynamics of the system. It also could be used to predict when jet flows

should reverse direction. However, the approximation underestimated the maximum intrusion jet

velocities during syzygy and overestimated them before the peak of neap tides. This suggested

other factors were affecting jet velocities. These factors include a dynamic aquifer water table or

density gradients (equation assumed these to be static), and unknown dynamics within the

aquifer caves and conduits, among others.

Although instantaneous jet velocities during low tides were stronger at syzygy than

quadrature, subtidal velocities were greater in quadrature than syzygy (dashed red line of Figure

6c). This is relevant because previous studies have observed greater SGD seepage discharge rates

in syzygy than in quadrature [Kim and Hwang, 2002; Taniguchi, 2002]. The different reaction

observed at Pargos spring could be attributed to the fact that point-source SGDs have faster

response times to changes in water surface than seepage SGD. Faster responses are caused by the

high hydraulic conductivity (describes water movement through soils) of the Yucatan aquifer

system [Valle-Levinson et al., 2011]. The response time of Pargos spring velocities, U , to

20

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

changes in the water surface, η, was 22 min as calculated with a cross-correlation of the two

data sets (r2=0.65). Similarly, Vera et al., [2012] measured the long-term variability of a coastal

karst aquifer in the northern Yucatan Peninsula and found high correlation and phase agreement

between sea level variations and spring discharge. Furthermore, lag times between coastal karst

aquifers and sea level have been studied, specifically during the passage of hurricanes [Vera et

al., 2012]. These authors observed reversals in the flow at a submarine spring in the northern

Yucatan coast during Hurricane Ike. Salt intrusion lasted 2 days as observed at a cenote or

sinkhole just 150 m from shore, and remained for 26 days at a well found 10 km inland. In

contrast, [Smith et al., 2008] observed the effects of the passage of two hurricanes on seepage

SGD on the Florida coast. The hurricanes produced a reversal in the hydraulic gradient that took

80-130 d to dissipate. The large difference between the two locations could be attributed, again,

to the greater hydraulic conductivity of the Yucatan’s aquifer system when compared to

Florida’s aquifer, even though both are of karstic origin.

Moreover, subtidal jet velocities were positive (outflow) throughout with the exception of

midday 28 July 2011 (Figure 6c), which coincided with the peak wave setup (Figure 6b, ~5 cm)

generated by the storm. This was noteworthy because it showed the great vulnerability of the

system to small changes in water level. An increase of only 5 cm in the water level dropped the

subtidal velocities from 8 cm s-1 to zero. Wave setup increased hydrostatic pressure over the

spring, which appeared to inhibit both the discharge and turbulence.

4.2 Turbulent Kinetic Energy (TKE)

Semidiurnal and fortnightly oscillations were also observed in the TKE time series

(Figure 7d). Peaks in the jet’s TKE became more intense as quadrature tides approached,

reaching maximum TKE on 23 July. Thereafter, TKE maxima progressively decreased as tidal

21

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

amplitudes increased, pointing to an inverse relationship between the TKE trend and tidal range.

The lowest values of TKE (0.001 m2 s-2) were observed during syzygy tides, suggesting that tidal

pumping during syzygy tides inhibited TKE growth from the jet discharge. Furthermore, wave-

induced setup could have inhibited TKE at the jet, as the greatest significant wave height

coincided with the peak of syzygy tides and the minimum TKE values. A setup of only 5 cm

reduced subtidal TKE by half. Therefore, TKE were primarily driven by semidiurnal tides, with

the tidal range and wave setup acting as a secondary drivers. An exponential equation to explain

the relationship between TKE and sea level variations was proposed by Parra et al., [2014] for a

3 day time series. However, their relationship does not explain the TKE pattern observed here

(r2=0.09) and even the best fit between TKE and water level was not reliable (r2=0.23). The

exponential relationship previously proposed would suggest greater TKE produced during

syzygy when the tidal range was greatest. Yet, TKE decreased with the approach of syzygy. This

suggests that TKE observed at the spring opening could be comprised of turbulence that was

locally-generated and advected from within the aquifer.

Flows in the conduits and caves of the Yucatan aquifer system are known to be

dominated by turbulence [Bauer-Gottwein et al., 2011; Beddows et al., 2007]. Generally,

turbulence is expected to be proportional to mean velocities. However, the maximum TKE at the

jet did not occur at maximum jet velocities (Figure 6c). Consequently, it is postulated that

turbulence intensity was driven by other factors. It could be assumed that the TKE observed at

the jet included two components: turbulence generated locally by jet velocities and turbulence

advected from within the aquifer system. Using this assumption, it can be expected that extended

periods of outflow would produce increasing levels of turbulence within the aquifer. In other

words, we could expect increased TKE when subtidal flows were greatest, and decreased TKE

22

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

when subtidal flows were least. This relationship was observed when comparing subtidal TKE

with subtidal jet velocities (Figure 6c). Additionally, maximum TKE on 22 July also coincided

with the minimum jet velocity range. The small jet velocity range produced steady flows within

the aquifer that allowed for the greatest growth of TKE. This explanation is analogous to the

system having “memory” of what previously happened. The “memory” of the aquifer is to

discharge water toward the ocean. The aquifer had the “memory” of previous cycles when

subtidal outflows were maximum (during quadrature), producing the subsequent buildup of

TKE. However, jet flows during syzygy were greatly influenced by pulses of inflow and by

minimum subtidal velocities. During syzygy the system appeared to have lost “memory” of

previous cycles as represented by weakened subtidal outflows and subtidal turbulence. Inflow

pulses effectively disrupted the “memory” of the aquifer. Turbulence within the aquifer is

relevant as it could increase the rate of chemical reactions between saltwater and the karstic

substrate [Clemens et al., 1997; Kaufmann and Braun, 1999; Moore, 1999]. These reactions

could alter the concentrations of nutrients that discharge into the lagoon and affect the coral reefs

and related ecosystems. Future efforts should consider these responses. Furthermore, increased

turbulence in the spring discharge could produce increase rates of mixing between jet and lagoon

waters that could dilute pollutants and increase the exchange rates between reefs and seawater

[Reidenbach et al., 2006].

4.3 Waves

Wind-waves are known to drive circulation within the Puerto Morelos lagoon [Enriquez

et al., 2012; Coronado et al., 2007]. Spectral energy of wind-generated waves observed at the

forereef and over the spring was mostly concentrated within the 0.1-0.3 Hz frequency range

throughout the time series (Figure 4c). Infragravity wave energy (<0.1 Hz) increased during

23

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

periods of higher significant wave heights (between 23-26, 28-30 July), and were of greater

relative energy over Pargos spring (Figure 5b) than at the forereef. Therefore, a link was

apparent between increased wave heights and infragravity wave enhancement observed at the

forereef. The presence of infragravity waves in the Puerto Morelos lagoon has been reported

numerically [Torres-Freyermuth et al., 2012]. According to them, these low-frequency pulses

were controlled by offshore wave energy and related to “nonlinear energy transfer from higher

frequencies.” Therefore, greater activity within the infragravity range was to be expected with

increased wave energy from wind forcing, as observed.

On the other hand, the spectral energy of the jet discharge (Figure 5d) was different from

that of the water level (Figure 5b). Jet velocity spectral energy varied with a semidiurnal pattern.

During quadrature tides, elevated spectral energy was observed at all frequencies (e.g., 22 July).

Syzygy tides, however, showed low spectral energy at the highest frequencies during high tides

(e.g., day 28), when the discharge reversed directions. Additionally, a band of high spectral

energy appeared within the wind-wave frequency range (around 0.1 Hz). This indicated that

wind-wave effects could have influenced the jet discharge during syzygy at high tides. Wind-

wave influence on jet discharge was corroborated with the squared-coherence between sea level

and jet discharge at Pargos spring (Figure 5e). Significant squared-coherence (>99 % confidence

limit, bold black contour line) was observed during the second half of the time series within the

0.01-0.3 Hz frequency range that included both wind-waves and infragravity waves. This

showed that waves indeed modulated the discharge through wave pumping. Such observation

can be explained by the fact that the periods of high coherence at the wind-wave and infragravity

frequencies occurred at high tides, when jet velocities had reversed (salt-intrusion). In these

cases, wave influence would be observed as the jet discharge became lagoon water flowing into

24

508

509

510

511

512

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

the aquifer. When the jet discharge reversed direction during high syzygy tides and 5 cm wave

setup, wind-waves and infragravity waves played a role in modulating the discharge through

wave pumping.

4.4 Jet Temperatures and Salinities

In addition to the effects of tides and waves on the jet discharge, tidal effects were

observed on the jet’s temperature and salinity (Figure 6d, e). Quadrature tides showed starkly

different temperature and salinity oscillations than syzygy tides. At the height of quadrature tides

throughout 22 July, jet temperatures and salinities were practically constant (~27.1 °C, ~29 g kg-

1). The rest of quadrature tides produced triangle-like peaks with sharp increases toward high

tides and more gradual decreases toward low tides. A similar triangular pattern was observed at

this jet in July 2010 [Parra et al., 2014]. In contrast, syzygy tides produced greater temperature

and salinity oscillations between high and low tides (up to 3 °C, 10 g kg-1) than quadrature tides

(≤1 °C, <4 g kg-1). Both temperature and salinity oscillations during syzygy tides exhibited a

rectangular shape with sharp increases as high tides approached, and sudden drops as low tides

were reached. During syzygy high tides, temperatures and salinities at the jet were nearly

equivalent to those observed at the lagoon bottom, suggesting seawater intrusion. Concurrently,

jet velocities were negative, showing flow of water from the lagoon into the aquifer. A similar

rectangular pattern was observed by Expósito-Díaz et al., [2013], although their jet temperatures

were warmer than the ambient during their measurement period (autumn in a mangrove lagoon

some 160 km south of Puerto Morelos) and flow reversals were not observed. Low tides

produced smoothly decreasing jet temperatures. Because the water temperature within the

aquifer remains ~25° C year-round, the decreasing jet temperatures indicated that discharging

spring waters originated from further inland within the aquifer. Inside the Yucatan aquifer

25

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

system, a trend of decreasing temperatures from the coast to inland has been previously observed

[Beddows et al., 2007]. However, low tides also displayed increasing jet salinities, which point to

increased mixing of aquifer and sea waters throughout the progression of low tides. These

seemingly contradicting patterns suggested that the jet conduit is connected to a stratified cave

within the aquifer with cool and salty water below cool and fresher waters (Figure 9), as

previously suggested by Parra et al. [2014]. As low tides approached, the first waters to

discharge were from the cool and fresher water layer on top (Figure 9a), corroborated by the

salinity minima observed at the start of outflow at every low tide after 25 July (Figure 6e). Then

as discharge continued, mixing between the two cave-layers increased, producing higher salinity

discharge (Figure 9b). Furthermore, because discharging waters originated from further within

the aquifer as low tide progressed, discharge temperatures also gradually decreased throughout

low tides. Therefore, this aquifer cave system has a vertical salinity gradient and a horizontal

temperature gradient. The Yucatan Peninsula aquifer system is karstic and highly hydraulically

conductive. Cave divers have mapped many miles of submerged caves and tunnels and have

discovered an abundance of stratified caves throughout [Beddows et al., 2007]. Additionally,

unpublished field data show a cavity in the aquifer of 30-50 m height that appears to connect to

Pargos spring (M. Rebolledo-Vieyra pers. comm.) and divers at Pargos have observed the jet

conduit spanning ~10 m, at which point the conduit opened into a large cave. Further fieldwork

is necessary in this particular area to produce more detailed topo-bathymetric and hydrographic

charts of the aquifer cave systems.

A stratified water column within the cave could be expected despite the turbulence

observed in the jet discharge and the inflow of lagoon waters. By considering conservation of

26

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

mass, Q=U1A1, the velocities created by the salt intrusion at the jet into the aquifer cave could be

ascertained.

U 1( π D12

4 )=U 2( π D22

4 ) (8)

U 1 D12=U 2 D2

2 (9)

The observed inflow of lagoon waters at the jet were of 0.2 m s-1 magnitude through an

opening of ~1 m diameter, into a cave of assumed diameter of 30 m would produce a velocity of

0.2 mm s-1 within the cave. These low cave velocities would allow the formation of a stratified

water column, as have been observed throughout the area [Beddows et al., 2007].

5 Conclusion

The jet discharge was primarily modulated by semidiurnal tides and explained by a

relationship including density gradients and hydraulic gradients between the aquifer and jet, and

frictional losses throughout the system. A wave setup of 5 cm within the lagoon reduced subtidal

jet velocities by 8 cm s-1 and subtidal TKE by half. The jet salinity and temperature readings

suggested that the jet conduit was connected to a stratified cave with fresher cool waters

overlaying saltier cool waters (Figure 9).

This expansion in our understanding of the connections between coastal karstic aquifers

and sea level modulations has particular implications for water resource protection. Not only

must we be concerned with the overuse of groundwater reserves, but we should consider how

rising sea levels can also adversely affect this valuable resource. However, rising seas in

conjunction with storm events and syzygy tides can potentially pollute coastal aquifers within a

shorter time scale than previously expected. Additionally, the dynamics of these spring

discharges also have implications for the coastal ecosystems of the area. Therefore, further

27

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

studies are needed to understand these systems through longer temporal scales and including

remote forcing such as variations outside the lagoon in the Yucatan Current.

REFERENCES

Baas, A. C. (2006), Wavelet power spectra of aeolian sand transport by boundary layer

turbulence, Geophys. Res. Lett., 33(5).

Bakhtyar, R., D. A. Barry, and A. Brovelli (2012), Numerical experiments on interactions

between wave motion and variable-density coastal aquifers, Coast. Eng., 60, 95-108.

Bauer-Gottwein, P., B. R. Gondwe, G. Charvet, L. E. Marín, M. Rebolledo-Vieyra, and G.

Merediz-Alonso (2011), Review: The Yucatán Peninsula karst aquifer, Mexico, Hydrogeol.

J., 19(3), 507-524.

Beddows, P. A., P. L. Smart, F. F. Whitaker, and S. L. Smith (2007), Decoupled fresh–saline

groundwater circulation of a coastal carbonate aquifer: spatial patterns of temperature and

specific electrical conductivity, Journal of Hydrology, 346(1), 18-32.

Bricker, J. D. and S. G. Monismith (2007), Spectral wave-turbulence decomposition, J. Atmos.

Ocean. Technol., 24(8), 1479-1487.

Carretero, S., J. Rapaglia, H. Bokuniewicz, and E. Kruse (2013), Impact of sea-level rise on

saltwater intrusion length into the coastal aquifer, Partido de La Costa, Argentina, Cont.

Shelf Res., 61, 62-70.

Chérubin, L., C. Kuchinke, and C. Paris (2008), Ocean circulation and terrestrial runoff

dynamics in the Mesoamerican region from spectral optimization of SeaWiFS data and a

high resolution simulation, Coral Reefs, 27(3), 503-519.

28

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

Clemens, T., D. Hueckinghaus, M. Sauter, R. Liedl, and G. Teutsch (1997), Modelling the

genesis of karst aquifer systems using a coupled reactive network model, IAHS

PUBLICATION(241), 3-10.

Coronado, C., J. Candela, R. Iglesias-Prieto, J. Sheinbaum, M. López, and F. Ocampo-Torres

(2007), On the circulation in the Puerto Morelos fringing reef lagoon, Coral Reefs, 26(1),

149-163.

Emery, W. J. and R. E. Thomson (2001), - Data Analysis Methods in Physical Oceanography,

1st ed., Elsevier.

Enriquez, C., I. Mariño-Tapia, A. Valle-Levinson, A. Torres-Freyermuth, and R. Silva-Casarín

(2012), Mechanisms driving the circulation in a fringing reef lagoon: a numerical study,

Book of Abstracts of the Physics of Estuaries and Coastal Seas Symposium.

Expósito-Díaz, G., M. A. Monreal‐Gómez, A. Valle‐Levinson, and D. A. Salas‐de‐León (2013),

Tidal variations of turbulence at a spring discharging to a tropical estuary, Geophys. Res.

Lett., 40(5), 898-903.

Ferguson, G. and T. Gleeson (2012), Vulnerability of coastal aquifers to groundwater use and

climate change, Nature Climate Change, 2(5), 342-345.

Fitterman, D. V. (2014), Mapping Saltwater Intrusion in the Biscayne Aquifer, Miami-Dade

County, Florida using Transient Electromagnetic Sounding, Journal of Environmental and

Engineering Geophysics, 19(1), 33-43.

Fleury, P., M. Bakalowicz, and G. de Marsily (2007), Submarine springs and coastal karst

aquifers: a review, Journal of Hydrology, 339(1), 79-92.

Gonneea, M. E., A. E. Mulligan, and M. A. Charette (2013), Climate‐driven sea level anomalies

modulate coastal groundwater dynamics and discharge, Geophys. Res. Lett., 40(11), 2701-

2706.

29

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

González-Herrera, R., I. Sánchez-y-Pinto, and J. Gamboa-Vargas (2002), Groundwater-flow

modeling in the Yucatan karstic aquifer, Mexico, Hydrogeol. J., 10(5), 539-552.

Grinsted, A., J. C. Moore, and S. Jevrejeva (2004), Application of the cross wavelet transform

and wavelet coherence to geophysical time series, Nonlinear processes in geophysics,

11(5/6), 561-566.

Hamner, W. W. and E. E. Wolanski (1988), Hydrodynamic forcing functions and biological

processes on coral reefs: a stus review, Proceedings of the 6th International Coral Reef

Symposium, Townsville, Australia 8-12 August 1988-pages: 1: 103-113.

Hu, C., F. E. Muller‐Karger, and P. W. Swarzenski (2006), Hurricanes, submarine groundwater

discharge, and Florida's red tides, Geophys. Res. Lett., 33(11).

Kang, S. and H. Lin (2007), Wavelet analysis of hydrological and water quality signals in an

agricultural watershed, Journal of Hydrology, 338(1), 1-14.

Kaufmann, G. and J. Braun (1999), Karst aquifer evolution in fractured rocks, Water Resour.

Res., 35(11), 3223-3238.

Kim, G. and D. Hwang (2002), Tidal pumping of groundwater into the coastal ocean revealed

from submarine 222Rn and CH4 monitoring, Geophys. Res. Lett., 29(14), 23-1-23-4.

Li, L., D. Barry, F. Stagnitti, and J. Parlange (1999), Submarine groundwater discharge and

associated chemical input to a coastal sea, Water Resour. Res., 35(11), 3253-3259.

McLachlan, A. and I. Turner (1994), The interstitial environment of sandy beaches, Mar. Ecol.,

15(3‐4), 177-212.

Merino, M., S. Czitrom, E. Jordán, E. Martin, P. Thomé, and O. Moreno (1990), Hydrology and

rain flushing of the Nichupté lagoon system, Cancún, Mexico, Estuar. Coast. Shelf Sci.,

30(3), 223-237.

30

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

Metcalfe, C. D., P. A. Beddows, G. G. Bouchot, T. L. Metcalfe, H. Li, and H. Van Lavieren

(2011), Contaminants in the coastal karst aquifer system along the Caribbean coast of the

Yucatan Peninsula, Mexico, Environmental pollution, 159(4), 991-997.

Millero, F. J., R. Feistel, D. G. Wright, and T. J. McDougall (2008), The composition of

Standard Seawater and the definition of the Reference-Composition Salinity Scale, Deep

Sea Research Part I: Oceanographic Research Papers, 55(1), 50-72.

Monismith, S. G. (2010), Mixing in estuaries, Contemporary Issues in Estuarine Physics, edited

by: Valle-Levinson, A, Cambridge University Press, Cambridge, 145-185.

Moore, W. S. (1996), Large groundwater inputs to coastal waters revealed by 226Ra

enrichments, Nature, 380(6575), 612-614.

Moore, W. S. (1999), The subterranean estuary: a reaction zone of ground water and sea water,

Mar. Chem., 65(1), 111-125.

Moore, W. S. (2010), The effect of submarine groundwater discharge on the ocean, Annual

review of marine science, 2, 59-88.

Null, K. A., K. L. Knee, E. D. Crook, N. R. de Sieyes, M. Rebolledo-Vieyra, L. Hernández-

Terrones, and A. Paytan (2014), Composition and fluxes of submarine groundwater along

the Caribbean coast of the Yucatan Peninsula, Cont. Shelf Res., 77, 38-50.

Okiishi, M. Y., B. Munson, and D. Young (2006), Fundamentals of Fluid Mechanics, John

Wiley & Sons, Inc.

Parra, S., I. Mariño-Tapia, C. Enriquez, and A. Valle-Levinson (2014), Variations in turbulent

kinetic energy at a point-source submarine groundwater discharge in a reef lagoon, Ocean

Dynamics.

Patterson, M. R., K. P. Sebens, and R. R. Olson (1991), In situ measurements of flow effects on

primary production and dark respiration in reef corals., Limnol. Oceanogr., 36(3), 936-948.

31

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

Peterson, R. N., W. C. Burnett, C. R. Glenn, and A. G. Johnson (2009), Quantification of point-

source groundwater discharges to the ocean from the shoreline of the Big Island, Hawaii,

Limnol. Oceanogr., 54(3), 890-904.

Pope, S. B. (2000), Turbulent Flows, Cambridge university press.

Precht, E. and M. Huettel (2003), Advective pore-water exchange driven by surface gravity

waves and its ecological implications, Limnol. Oceanogr., 48(4), 1674-1684.

Reidenbach, M. A., S. G. Monismith, J. R. Koseff, G. Yahel, and A. Genin (2006), Boundary

layer turbulence and flow structure over a fringing coral reef, Limnol. Oceanogr., 51(5),

1956.

Santos, I. R., W. C. Burnett, J. Chanton, N. Dimova, and R. N. Peterson (2009), Land or ocean?:

Assessing the driving forces of submarine groundwater discharge at a coastal site in the Gulf

of Mexico, Journal of Geophysical Research: Oceans (1978–2012), 114(C4).

Santos, I. R., F. Niencheski, W. Burnett, R. Peterson, J. Chanton, C. F. Andrade, I. B. Milani, A.

Schmidt, and K. Knoeller (2008), Tracing anthropogenically driven groundwater discharge

into a coastal lagoon from southern Brazil, Journal of Hydrology, 353(3), 275-293.

Smith, C. G., J. E. Cable, and J. B. Martin (2008), Episodic high intensity mixing events in a

subterranean estuary: Effects of tropical cyclones, Limnol. Oceanogr., 53(2), 666.

Taniguchi, M. (2002), Tidal effects on submarine groundwater discharge into the ocean,

Geophys. Res. Lett., 29(12), 2-1-2-3.

Taniguchi, M., W. C. Burnett, J. E. Cable, and J. V. Turner (2002), Investigation of submarine

groundwater discharge, Hydrol. Process., 16(11), 2115-2129.

Torrence, C. and G. P. Compo (1998), A practical guide to wavelet analysis, Bull. Am. Meteorol.

Soc., 79(1), 61-78.

32

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

Torres-Freyermuth, A., I. Marino-Tapia, C. Coronado, P. Salles, G. Medellín, A. Pedrozo-

Acuna, R. Silva, J. Candela, and R. Iglesias-Prieto (2012), Wave-induced extreme water

levels in the Puerto Morelos fringing reef lagoon, Natural Hazards and Earth System

Science, 12(12), 3765-3773.

Uchiyama, Y., K. Nadaoka, P. Rölke, K. Adachi, and H. Yagi (2000), Submarine groundwater

discharge into the sea and associated nutrient transport in a sandy beach, Water Resour.

Res., 36(6), 1467-1479.

Valle-Levinson, A., I. Marino-Tapia, C. Enriquez, and A. F. Waterhouse (2011), Tidal variability

of salinity and velocity fields related to intense point-source submarine groundwater

discharges into the coastal ocean, Limnol. Oceanogr., 56(4), 1213-1224.

Vera, I., I. Mariño-Tapia, and C. Enriquez (2012), Effects of drought and subtidal sea-level

variability on salt intrusion in a coastal karst aquifer, Marine and Freshwater Research,

63(6), 485-493.

Werner, A. D. and C. T. Simmons (2009), Impact of sea‐level rise on sea water intrusion in

coastal aquifers, Groundwater, 47(2), 197-204.

Zektser, I. S. and L. G. Everett (2000), Groundwater and the Environment: Applications for the

Global Community, CRC Press.

 

33

714

715

716

717

718

719

720

721

722

723

724

725

726

727

728

729

730

731

732

Figure 1. Chart of the Puerto Morelos fringing coral reef lagoon on the eastern Yucatan

Peninsula of Mexico. Circle represents the location of Pargos spring where the Nortek Vector

and CTD were moored, squares represent the locations of the Nortek Vector at the forereef and

of the CTD within the lagoon, and the ‘X’ represents the meteorological station where winds

were recorded (depth and scale in meters).

34

733

734

735

736

737

738

739

Figure 2. Mean precipitation at nearest station in Cancun, Mexico (~30 km north of Puerto

Morelos). Data provided by Meteorological Assimilation Data Ingest System (MADIS) at

National Oceanic and Atmospheric Administration (NOAA).

35

740

741

742

743

744

Figure 3. Schematic of bathymetry at spring opening with approximate dimensions. Triangle

denotes location of ADV and CT moorings. The ADV was moored at the center of the vent (~7

m depth) with downward orientation. The CT was located about 25 cm above the sampling

volume of the ADV, both within the core or outflow.

36

745

746

747

748

749

Figure 4. Winds and waves at Puerto Morelos lagoon. (a) Hourly wind velocity vectors (blue),

and east (black) and north (red) wind components measured at the meteorological station,

showing the direction the wind was blowing; (b) significant wave heights, Hs, (blue) and peak

wave periods (green) at the forereef; and (c) spectral density of the water level bursts at the

forereef.

37

750

751

752

753

754

755

Figure 5. (a) Time series and (b) spectral contours of the hourly sea level bursts at Pargos

spring; (c) time series and (d) spectral contours of the hourly spring discharge velocities; and (e)

squared-coherence spectra between the sea level variations at Pargos and spring discharge bursts,

with the black line representing the 99 % confidence limit. The two black dashed lines represent

instances of peak high and low tides.

38

756

757

758

759

760

761

762

Figure 6. (a) Sea level variations at Pargos (blue) and the forereef (red), the dashed green line

represents the high tide observed during the first flow reversal event on day 25.5; (b) setup (blue)

and significant wave heights (blue) observed at the jet; (c) TKE (blue) and jet discharge

velocities (red) at the jet opening, with the dashed lines representing 30-hour low-pass filter and

the dotted line representing the zero velocity value; (d) temperatures at the jet (blue) and the

lagoon bottom (black); and (e) salinities at the jet (blue) and the lagoon bottom (black). The

black dashed lines represent instances of peak high and low tides.

39

763

764

765

766

767

768

769

Figure 7. Wavelet and time series of significant wave heights (a) at the forereef; (b) water

level at the jet; (c) jet vertical velocities; and (d) TKE at the jet. In the wavelet plots, shaded

regions on both bottom corners indicate the cone of influence, where edge effects become

important and data is less reliable. The bold black contours in wavelet plots encircle regions of

>95 % confidence.

40

770

771

772

773

774

Figure 8. (a) Sea level variations at Pargos spring and (b) actual (bold, black) and estimated jet

discharge velocities for varying hydraulic conductivities: 5 mm km-1(green), 30 mm km-1(blue)

and 100 mm km-1(red). The dotted black line represents the zero velocity mark.

41

775

776

777

778

779

Figure 9. Schematic of proposed aquifer cave system at the (a) start and (b) middle of low

tides in syzygy. The dashed line represents the halocline or interphase between the two water

layers.

42

780

781

782

783

784