You are on page 1of 13

Journal of Oceanography, Vol. 65, pp.

335 to 347, 2009

Accurate Ocean Tide Modeling in Southeast Alaska and


Large Tidal Dissipation around Glacier Bay
D AISUKE INAZU1*, TADAHIRO S ATO2, SATOSHI MIURA2, YUSAKU O HTA2, KAZUYUKI N AKAMURA1,
H IROMI FUJIMOTO2, C HRISTOPHER F. LARSEN3 and TOMOYUKI HIGUCHI1
1
The Institute of Statistical Mathematics/Japan Science and Technology Agency,
Minami-Azabu, Minato-ku, Tokyo 106-8569, Japan
2
Research Center for Prediction of Earthquakes and Volcanic Eruptions, Graduate School of
Science, Tohoku University, Aramaki-Aza-Aoba, Aoba-ku, Sendai 980-8578, Japan
3
Geophysical Institute, University of Alaska Fairbanks, Koyukuk Dr, Fairbanks, AK 99775, U.S.A.

(Received 19 September 2008; in revised form 8 December 2008; accepted 16 January 2009)

An accurate prediction of ocean tides in southeast Alaska is developed using a re- Keywords:
gional, barotropic ocean model with a finite difference scheme. The model skill is ⋅ Bathymetry,
verified by the observational tidal harmonics in southeast Alaska including Glacier ⋅ bottom friction,
Bay. The result is particularly improved in Glacier Bay compared to the previous ⋅ Glacier Bay,
⋅ tidal energy
model described by Foreman et al. (2000). The model bathymetry dominates the model
dissipation.
skill. We re-estimate tidal energy dissipation in the Alaska Panhandle and suggest a
value for tidal energy dissipation of 3.4 GW associated with the M2 constituent which
is 1.5 times the estimation of Foreman et al. (2000). A large portion of the M2 energy
budget entering through Chatham Strait is dissipated in the vicinity of Glacier Bay.
Moreover, it is shown that the developed model has the potential to correct the ocean
tide loading effect in geodetic data more efficiently than the model of Foreman et al.
(2000), especially around Glacier Bay.

1. Introduction tion in each specified area leads to a better estimation


There are many coastal and inland seas with compli- over the Gulf of Alaska. Foreman et al. (2000) estimated
cated coastlines and bathymetry around the Gulf of that the M2 tidal energy dissipation in central Alaska and
Alaska. Ocean tides are strongly confined by the coast- Cook Inlet exceeds 70% of the total tidal energy dissipa-
lines and bathymetry in the seas. The Gulf of Alaska is tion in the Gulf of Alaska, and 7% of the total energy is
known to be a region with large tidal energy dissipation dissipated in the Alaska Panhandle. The portion of the
due to bottom friction (e.g., Miller, 1966; Le Provost and tidal energy dissipated in the Alaska Panhandle is not
Lyard, 1997; Foreman et al., 2000; Egbert and Ray, 2003). necessarily negligible. On the other hand, land elevations
Estimations of the M 2 tidal dissipation in the Gulf of due to the post glacial rebound and present-day ice melt-
Alaska involve relative differences from 30 (Foreman et ing around Glacier Bay in southeast Alaska have been
al., 2000) to 70 (Lyard and Le Provost, 1997) GW, which investigated by geodetic measurements (e.g., Larsen et
leaves much room for improvement. Almost all the tidal al., 2003, 2004; Sato et al., 2008). The GPS and gravim-
energy in the Gulf of Alaska is dissipated in limited ar- eter data include ocean tide components derived in the
eas: central Alaska and Cook Inlet, Alaska Panhandle, Alaska Panhandle, which needs appropriate correction in
northern British Columbia, and around Vancouver Island. order to investigate the dynamics of the solid Earth. The
A more accurate estimation of the tidal energy dissipa- development of an accurate ocean tide model in the Alaska
Panhandle is needed to estimate the tidal energy dissipa-
tion and to correct the geodetic data.
* Corresponding author. E-mail: inazud@aob.geophys.tohoku. Global and regional ocean tide models cover the
ac.jp Alaska Panhandle. One of the global models resolving
Present address: Research Center for Prediction of Earthquakes and the Alaska Panhandle is FES2004 (Lyard et al., 2006).
Volcanic Eruptions, Graduate School of Science, Tohoku University,
However, the FES2004 model yields inaccurate tidal vari-
Aramaki-Aza-Aoba, Aoba-ku, Sendai 980-8578, Japan.
ations in the Alaska Panhandle; for example, the observed
Copyright©The Oceanographic Society of Japan/TERRAPUB/Springer M 2 tidal elevation amplitude and phase at Ketchikan

335
(Skagway) are 188.2 (203.5) cm and 270.1° (283.9°), re-
spectively, while the modeled amplitude and phase are
250.8 (376.3) cm and 276.2° (79.6°), respectively. On the
other hand, Foreman et al. (2000, henceforth F00) devel-
oped an accurate ocean tide model and estimated tidal
energy dissipation due to bottom friction in the northeast
Pacific Ocean including the Alaska Panhandle; however,
the F00 model was not validated in Glacier Bay. Larsen
et al. (2004) conducted ocean tide observations in the
vicinity of Glacier Bay, and the observed tidal harmonics
do not show good agreement with the F00 model; the
observed M2 tidal elevation amplitude and phase at Com-
posite Island are 204.3 cm and 299.7°, respectively, while
the F00 model shows the amplitude and phase to be 176.9
cm and 286.6°, respectively. Therefore, the model result
of F00 may be inappropriate for estimating the tidal en-
ergy dissipation rate around Glacier Bay and removing
the ocean tide components from the geodetic data. In this
study we develop a more accurate ocean tide model, esti-
mate the tidal energy dissipation in the Alaska Panhandle
including Glacier Bay, and show the contribution of the Fig. 1. Computational domain, tide gauges, an ocean bottom
developed ocean tide model to the geodetic aspect. The pressure gauge, GPS stations used in this study, and bottom
topography in meters drawn from the GINA bathymetric
primary purpose is the accurate ocean tide modeling at
data. Open circles, a black circle, and diamonds denote the
the regional scale rather than smaller scales of each harbor
tide gauges, the ocean bottom pressure gauge, and the GPS
and inlet. stations, respectively. An example of the model grids around
Glacier Bay is shown in upper right of the figure.
2. Model Description

2.1 Tidal equation


The numerical model used in the present study is tom drag and horizontal eddy viscosity are expected to
configured below. A non-linear, single-layer shallow wa- impede water motions in the inland sea. The coefficients
ter model (Hirose and Yoon, 1996; Kim and Yoon, 1996) of the bottom friction and the horizontal eddy viscosity
is used to simulate the ocean tides in southeast Alaska. are optimized in Subsection 3.2.
Barotropic motions of seawater in spherical coordinates The differential equations are discretized by a sec-
are governed by the following equations of momentum ond-order finite difference scheme to adapt to the stag-
and continuity: gered Arakawa-C grid system which is favorable to wa-
ter mass conservation (Arakawa and Lamb, 1981). The
∂u uu time integration is accomplished using a second-order
+ (u ⋅ ∇)u + f × u = − g∇η − γ b + AH ∇ 2 u, (1) leapfrog scheme associated with the Asselin filter
∂t H
(Asselin, 1972) to suppress high-frequency numerical
noise.
∂η
+ ∇ ⋅ (uH ) = 0. (2 )
∂t 2.2 Bathymetry
Wave propagations and water exchanges are strongly
A vector u is the horizontal depth-integrated velocity, η dissipated by the bottom friction through the shallow
is the sea surface elevation, f is the Coriolis parameter, g straits. The coastlines must be carefully drawn to enable
is the gravitational acceleration (=9.8 m sec –2), and H is seawater to flow appropriately through narrow channels
the total depth of the water column. Dissipation terms and straits. The model grid spacing is set to 1′ × 1′ and
are parameterized by two coefficients: γ b for quadratic the coastlines are based on GTOPO30 topographic data
bottom friction, and AH for horizontal viscosity. The bot- (http://edc.usgs.gov/products/elevation/gtopo30/
tom topography is set in Subsection 2.2. Dissipation pa- gtopo30.html). The computational domain is shown in
rameters are usually determined empirically. Since the Fig. 1. Although some inlets with spatial scales less than
Alaska Panhandle is characterized by complicated coast- a few kilometers are not well resolved in the model, our
lines with shallow banks and fjordic systems, both bot- primary purpose is the accurate modeling of the ocean

336 D. Inazu et al.


Longitude
136 W 134 W 132 W 130 W

59 N

58 N
Latitude

57 N

56 N

55 N

m Fig. 3. Model dependence on tidal elevation conditions and


−300 −200 −100 0 100 200 300
coefficients of the horizontal eddy viscosity with the im-
proved model bathymetry as a function of the coefficient
of the bottom friction. Solid lines indicate simulations with
Fig. 2. Difference between GINA and the multibeam-derived
the tidal elevation condition given by merging the F00 and
data in meters. Blue (red) indicates that the multibeam-de-
F93 models, and dashed line indicates that only the F00
rived data is deeper (shallower) than GINA. Thick gray ar-
model is used for the tidal elevation condition. Black, red,
eas denote no multibeam-derived data.
and blue lines denotes the cases of AH = 1, 0.1, and 10
m 2 sec–1, respectively. The vertical axis indicates the total
rms difference of the four major constituents (M2, S 2, K1,
tides and tidal dissipation at regional scales, such as a and O 1) at the 15 sites.
strait and/or bay. For example, Cross Sound, Glacier Bay,
and Icy Strait are managed to be represented by the spa-
tial resolution. A no-slip condition is assumed along the 1999), which noted that multibeam-derived depth tends
coastlines. River inputs are not included at the coastal to be deeper than depth estimated by satellite-derived
boundary. gravity data in areas shallower than about 1000 m. The
Uncertainties and errors in bathymetric data directly accuracy of the depth estimation from the satellite
affect the computed wave propagation. The bottom bound- altimetry worsens at the coastal region, partly due to
ary condition is basically given by GINA (Lindquist et ambiguities in determination of the gravimetric geoid and
al., 2004; Marks and Smith, 2006), which is originally the dimensions of spherical harmonics (e.g., Sandwell and
interpolated to a 30-second resolution from ETOPO2 Smith, 2001; Rodríguez and Sevilla, 2002).
(Smith and Sandwell, 1997) by a bicubic method (K. Figure 2 indicates that the use of the original GINA
Engle, personal communication, 2007). On the other hand, bathymetry apparently leads faulty ocean modeling in the
the Alaska Panhandle is mostly covered by a large number Alaska Panhandle, where overall depth is basically shal-
of the multibeam-derived depth data under NGDC (NOS lower than a few hundred meters. Therefore we reflect
Hydrographic Survey Data). In addition, Carlson et al. the multibeam-derived data in the model bathymetry. The
(2002) conducted an extensive multibeam bathymetric GINA bathymetry is replaced at model grids where the
observation in the vicinity of Glacier Bay. These multibeam-derived data are obtained, and the accurate
multibeam-derived data are compiled to the gridded data model bathymetry is prepared. The shallowest depth of
by spatial averaging within each model grid. The com- the ocean grids is limited to 5 m for stable computations.
piled multibeam-derived data are compared to GINA (Fig. The bathymetry of the F00 model also includes
2). Large bathymetric errors in GINA are found in the multibeam-derived data, and the model bathymetry of our
Alaska Panhandle and on the shelf edge in the deep ocean. study and F00 probably has many common features. How-
The multibeam bathymetry is deeper overall than GINA ever, certain differences should be considered. For ex-
in the Alaska Panhandle. The deeper multibeam ample, only our model includes the depth observation
bathymetry in the shallow region is possibly supported conducted by Carlson et al. (2002). As described later,
by a report of Japan Hydrographic Association (JHA, the bathymetric condition is most crucial to the model

Ocean Tide Modeling in Southeast Alaska 337


Table 1. Accuracy of the F93 and F00 models for the M2 and K 1 constituents in Dixon Entrance. D (cm) is defined by Eq. (3). The
observed tidal harmonics are derived from table 2 of F93.

M2 K1

Site Longitude (°W) Latitude (°N) F93 F00 F93 F00


Dixon Entrance 132.94 54.56 1.1 6.0 1.0 2.5
Security Cove 132.85 54.75 1.2 7.2 0.7 1.3
Cape Muzon 132.69 54.64 1.8 6.5 0.4 1.7
Egeria Bay 132.97 54.23 6.5 4.9 2.9 1.5
Dixon Rose Spit 131.95 54.26 2.4 6.1 1.1 1.2
Nichols Bay 132.13 54.72 8.1 14.0 4.7 3.5
Cape Chacon 132.03 54.67 2.0 6.6 0.8 1.7
McLean Arm 131.97 54.79 5.0 6.7 1.4 1.6
Kelp Island 131.30 54.88 1.0 6.8 1.4 0.4
Foggy Bay 130.96 54.94 8.3 6.2 2.2 0.8
Brundige Inlet 130.85 54.62 8.6 10.4 2.6 1.8
Qlawdzeet 130.77 54.22 7.2 5.9 2.9 1.4
Prince Rupert 130.33 54.32 4.7 6.0 2.7 0.7

Average 4.5 7.2 1.9 1.6

accuracy, rather than the dissipation parameters and the not expected to be drastically improved compared to the
open boundary condition of the tidal elevation. previous F00 model. We aspire to achieve the accuracy
of the modeled tidal harmonics within several centimeters
2.3 Tidal elevation condition for each constituent in the Alaska Panhandle. The advan-
The open boundary condition of the model domain tage of merging the F00 and F93 models for the open
must be as accurate as possible. The F00 model covers boundary condition is shown quantitatively in Subsec-
the northeast Pacific Ocean, and seems to be appropriate tion 3.2.
in its representation of the tidal elevation thanks to the
assimilation of satellite altimeter data in the open sea. 3. Model Validation
Foreman et al. (1993, henceforth F93) computed the ocean
tides surrounding the Queen Charlotte Islands with spa- 3.1 Observed tidal harmonics
tial resolutions finer than those used in F00. We compare Ocean tides have been observed to obtain tidal har-
both the results and ascertain that the F93 model is more monics at permanent and temporary coastal sites, and an
accurate in Dixon Entrance than the F00 model (Table ocean bottom pressure gauge (OBPG). The coastal sta-
1). The F93 model shows the M2 constituent more accu- tions are administered by NOAA or the University of
rately than the F00 model, while the accuracy of the K1 Alaska Fairbanks (UAF, Larsen et al., 2004). The OBPG
constituent modeled by F00 is comparable to that given is under Tohoku University, Japan. The tidal harmonics
by F93. The result of F00 is thus used for the northeast at the sites shown in Fig. 1 are used for model validation
Pacific region and the result of F93 for Dixon Entrance. since some inlets are insufficiently represented in the
These tidal elevation conditions are linearly connected model due to its spatial resolution. At the UAF stations,
around the meridian 134°E for the eight major constitu- the sea level was recorded during periods of less than one
ents (M2, S2, K1, O1, N2, K2, P1, and Q1), which forces month (Larsen et al., 2004), and the tidal harmonics of
the computational boundary. Root-mean-square differ- the four major constituents (M2, S2, K1, and O1) were
ences (the definition is stated below, Eq. (3)) between calculated by a tidal analysis program, BAYTAP-G
the Foreman’s models and the observed tidal harmonics (Tamura et al., 1991). This program selects the optimal
of M 2 are several centimeters at coasts outside the Alaska grouping of tidal constituents based on the Bayesian in-
Panhandle. In addition, the boundary condition of the tidal ference (Akaike, 1980), and preferentially calculates the
elevation does not include the non-linear, shallow water dominant constituent within the group with reduced con-
tides which are concerned at the head of Dixon Entrance tamination by the other minor constituents (Ishiguro et
by F93. Therefore, the model accuracy in this study is al., 1983). We use long tide gauge time series to test sta-

338 D. Inazu et al.


Table 2. Tide gauges and OBPG used in this study.

Site Longitude (°W) Latitude (°N) Agency Duration (days)


1 Skagway 135.33 59.45 NOAA
2 Juneau 134.41 58.30 NOAA
3 Lynn Canal (OBPG) 135.04 58.58 Tohoku Univ. 365
4 Composite Island 136.57 58.89 UAF 28
5 Willoughby Island 136.12 58.61 UAF 14
6 False Bay 134.94 57.97 NOAA
7 Warm Spring Bay 134.83 57.09 NOAA
8 Petersburg 132.98 56.80 NOAA
9 Entrance Island 133.79 56.81 NOAA
10 Port Alexander 134.65 56.25 NOAA
11 Thoms Point 132.08 56.12 NOAA
12 Ketchikan 131.63 55.33 NOAA
13 Elfin Cove 136.34 58.19 NOAA
14 Sitka 135.34 57.05 NOAA
15 Craig 133.14 55.49 NOAA

bility of the tidal harmonics calculated by BAYTAP-G in bathymetry and the tidal elevation at the computational
various durations of the time series. Differences in the boundary have been described in the previous section.
time series duration from one year to two weeks have Here the contribution of the improved boundary condi-
been found to yield small errors within 3% in amplitude tions to the model skill is evaluated as well as the de-
and 3° in phase to the calculated tidal harmonics of the pendence of the model skill on the dissipation parameters.
four major constituents. The record length of the OBPG Figure 3 shows the model skill in cases of tidal elevation
is one year. Thus the tidal harmonics obtained at the fol- conditions and coefficients of the horizontal eddy viscos-
lowing 15 sites are used to validate the modeled tidal ity using the improved bathymetric condition as a func-
harmonics: the 12 stations of NOAA, the 2 temporary sta- tion of the bottom friction coefficient. The total rms dif-
tions of UAF in Glacier Bay, and the OBPG (Table 2). ference is smallest in the case of the tidal elevation con-
dition obtained by merging the F00 and F93 models with
3.2 Sensitivity of the model skill to boundary conditions the dissipation parameters γb = 0.02 and AH = 1 m2 sec –1.
and dissipation parameters The model dependence is apparent on the bottom friction
We compare the tidal harmonics obtained by the tide coefficient and is fairly slight on the horizontal eddy vis-
gauge observations with those of the simulations in this cosity coefficient. The model skill worsens when the F00
subsection. The model skill is evaluated by calculating result is used for the tidal elevation condition. This model
the total root-mean-square (rms) difference between the dependence on the tidal elevation condition is caused by
observed and the modeled tidal harmonics of the four the model accuracy at Clarence Strait. If the original GINA
major constituents at the 15 sites. The rms difference of bathymetry is applied to the model, the total rms differ-
each tidal constituent at each tide gauge and OBPG is ence is around 0.8–1.0, which indicates that the model
defined as: skill is absolutely sensitive to the model bathymetry. The
dependence on the spatial resolution is also an important

{( A
issue. The F00 model employs a finite element grid sys-
)
2
Dc, s = c, s cos Bc, s − ac, s cos bc, s tem, and its spatial resolution is partially finer than our
model, especially in the heads of channels and bays. Nev-
)}
2 12
(
+ Ac, s sin Bc, s − ac, s sin bc, s , (3) ertheless, the original calculation of F00 yields the model
skill of 0.276 which is less accurate than the simulation
using the F00 condition. The discrepancy between the two
where Ac,s (ac,s) and Bc,s (bc,s) are the modeled (observed) is mainly caused by the difference of the model skill in
amplitude and phase of each tidal constituent (c) at each Glacier Bay (not shown), which possibly arises from the
tide gauge (s). The sum of the relative error (Dc,s/ac,s) of difference of the model bathymetry noted in Subsection
the four major constituents at the 15 sites is used to cal- 2.2. The model skill probably depends more strongly on
culate the total rms difference, the model skill. the model bathymetry than on the spatial resolution. Since
The accurate conditions of the multibeam-derived the model dependence on the bottom friction coefficient

Ocean Tide Modeling in Southeast Alaska 339


340
D. Inazu et al.
Table 3. Observed and modeled tidal harmonics for the M2 and K 1 constituents at the 15 sites shown in Fig. 1. D is defined by Eq.
(3). Accuracy of the F00 model is listed as well as the Inazu model.

M2 K1
Observed Inazu model F00 Observed Inazu model F00
Amplitude Phase Amplitude Phase D D Amplitude Phase Amplitude Phase D D
Site (cm) (deg. UTC) (cm) (deg. UTC) (cm) (cm) (cm) (deg. UTC) (cm) (deg. UTC) (cm) (cm)
1 Skagway 203.5 283.9 205.6 282.0 7.0 8.0 52.7 265.4 54.7 264.5 2.1 1.2
2 Juneau 199.5 282.5 194.5 279.7 10.7 9.8 52.0 265.0 53.2 263.7 1.7 0.8
3 Lynn Canal (OBPG) 194.3 281.6 194.3 281.6 0.0 0.1 52.1 264.5 53.7 264.2 1.6 1.9
4 Composite Island 204.3 299.7 199.3 301.7 8.7 51.1 53.6 273.8 54.7 275.0 1.6 5.1
5 Willoughby Island 216.7 298.5 196.0 301.7 23.5 58.8 59.0 273.2 54.5 275.0 4.9 9.0
6 False Bay 181.2 283.2 184.3 281.1 7.3 7.9 50.5 263.4 52.7 263.9 2.2 2.8
7 Warm Spring Bay 164.3 279.3 165.2 278.6 2.3 8.7 48.7 262.3 50.6 262.9 1.9 3.3
8 Petersburg 193.0 285.5 190.1 279.0 21.8 25.8 50.4 268.1 52.9 263.7 4.7 3.1
9 Entrance Island 175.4 279.7 170.3 278.1 6.8 9.4 48.1 265.8 51.3 263.2 4.0 2.9
10 Port Alexander 123.3 276.1 120.1 274.9 4.1 3.0 47.2 264.1 47.4 261.8 1.9 0.8
11 Thoms Point 197.9 272.3 198.9 269.5 9.6 10.0 52.0 261.7 52.2 261.1 0.5 1.3
12 Ketchikan 188.2 270.1 185.4 267.7 8.2 12.7 51.3 261.1 51.1 260.5 0.6 2.1
13 Elfin Cove 124.9 276.1 124.5 276.6 1.2 3.3 47.8 263.9 47.7 263.9 0.2 1.2
14 Sitka 110.6 273.8 108.7 273.8 1.9 3.9 46.2 259.9 46.3 262.4 2.0 2.1
15 Craig 113.5 268.1 117.3 268.1 3.8 4.1 44.7 262.2 46.7 259.4 3.0 1.9
Average 7.8 14.4 2.2 2.6
Fig. 4. Computed M2 (a) amplitude and (b) phase. Contour intervals are (a) 10 cm and (b) 5°.

Fig. 5. Computed K 1 (a) amplitude and (b) phase. Contour intervals are (a) 2 cm and (b) 2°.

shows total rms differences around 0.2–0.3, the model connections of the straits. The simulation model with the
skill seems to be sensitive to the model bathymetry, the finite difference grid system would not reasonable to in-
bottom friction coefficient, and the spatial resolution in vestigate the model dependence on the resolution while
this order. retaining the connection of the straits. The present study
The grid spacing of 1′ × 1′ employed in the present therefore does not demonstrate the model dependence on
study is expected to be a minimal resolution to represent the spatial resolution which will need to be investigated
the connections of the narrow straits in the Alaska Pan- with a simulation model using the finite element grid sys-
handle, and coarser resolution is insufficient to retain the tem. In the context of the ocean tide modeling with the

Ocean Tide Modeling in Southeast Alaska 341


Fig. 6. (a) Mean M2 energy flux and (b) energy dissipation due to bottom friction.

finite element grid, Lefèvre et al. (2000) described that Laurent and Garrett, 2002; St. Laurent et al., 2002). The
accurate model bathymetry is essential in order to im- evaluation requires the spatial distribution of the buoy-
prove the ocean tide model over the Yellow and East China ancy frequency at ocean bottom. However, water tem-
Seas. They also showed the importance of tuning the bot- perature and salinity data derived from World Ocean At-
tom friction coefficient and a minor effect of the mesh las (Levitus and Boyer, 1994; Levitus et al., 1994) in-
refinement. This case is similar to the ocean tide modeling volves an insufficient spatial resolution and incorrect
in the present study. The discussions above will be help- depth for the Alaska Panhandle, and the effect of the in-
ful in general debate on accurate coastal ocean modeling. ternal tides is not quantitatively evaluated in the present
As mentioned above, the most accurate ocean tide study.
representation is achieved with γ b = 0.02 and A H = 1 Table 3 shows the model accuracy at each station
m2 sec–1 under the improved conditions of the boundary for the M2 and K1 constituents of the Inazu and the F00
tidal elevation and the model bathymetry. This simulated models. The model skill is highly improved in Glacier
result of the ocean tide model is called the Inazu model Bay compared to F00 for both the M2 and K1 constitu-
hereafter. The coefficient γb = 0.02 is set as the optimal ents. A slight improvement is found at Ketchikan. We can
coefficient of the bottom friction of our model, since both not clarify the reasons for the differences of the model
the semidiurnal and diurnal constituents show the small- skill between the Inazu and the F00 models, but the dif-
est rms differences around γb = 0.02 and the bottom fric- ference of the model bathymetry may be one reason even
tion coefficient second largely dominates the model skill though the spatial resolution of our model in the Alaska
after the model bathymetry. The model skill hardly de- Panhandle is partly coarser than that of F00. The improve-
pends on the coefficient of the horizontal eddy viscosity, ment at Ketchikan is probably due to the use of the F93
and the coefficient AH = 1 m2 sec–1 is employed for the result instead of the F00 result in Dixon Entrance. The
model based on the Richardson’s four thirds law. The observed M2 and K1 phase at Petersburg lags relatively
optimal bottom friction coefficient larger than a typical behind the modeled phase as well as F00. Although addi-
value (0.003) is possibly due to wall friction (Colbo, tional local dissipation seems to be needed, it is not in-
2006), form drag (Edwards et al., 2004), internal tide vestigated in detail here. A large rms difference is also
generation (e.g., Gargett, 1976; Pawlowics, 2002) in the found at Willoughby Island, which possibly arises from
complex channel/strait system, and effects of tidewater the short record length of only two weeks (Table 2).
glaciers (e.g., Prinsenberg, 1988) in the northern part of Nevertheless, the model accuracy in the overall Alaska
the Alaska Panhandle. Recently, Tanaka et al. (2007) Panhandle is achieved with the rms difference within sev-
quantitatively evaluated the drag due to internal tides eral centimeters for each major constituent. These fea-
within the Kuril Islands by a simple parameterization (St. tures of the model improvements are similar for the S2

342 D. Inazu et al.


and O1 constituents. cate that the remainder of the M2 energy flux entering
The M 2 and K 1 tidal harmonics computed from the through Clarence Strait and Chatham Strait after frictional
Inazu model are shown in Figs. 4 and 5. We see that the dissipation exceeds or offsets the energy inflows through
phase lag of the M2 (K1) constituent notably changes from Sumner Strait and Cross Sound, respectively. The Inazu
285° to 300° (268° to 275°) at the entrance to Glacier model shows an energy flux of 2.8 GW entering through
Bay. The change of the phase lag of both the constituents Chatham Strait, which is quite larger than the F00 esti-
is mostly equivalent in the time domain. The change of mation of only 0.6 GW. The difference in the estimation
the phase lag seems to arise from the shallow depth less probably arises because the line integral crossing Chatham
than 50 m of the entrance to Glacier Bay derived from Strait by F00 includes both Chatham Strait and Sumner
the observation of Carlson et al. (2002). Cokelet et al. Strait (M. G. G. Foreman, personal communication,
(2004) conducted a surface current observation in the vi- 2008).
cinity of Glacier Bay. The observed current is mostly ac- We now specify the region of the energy sink in the
counted for by tidal currents and shows a large spatial Alaska Panhandle. The energy dissipation due to bottom
change near the entrance to Glacier Bay, which possibly friction at each model grid is calculated by the following
indicates the robust change of the phase lag of tidal el- formula (e.g., Le Provost and Lyard, 1997):
evation in Glacier Bay. The change of the phase lag of
the M2 (K1) constituent at the entrance to Glacier Bay is 1 T
substantially greater than that of F00 which computed a Db = ∫0 ργ b u
3
dt, (5)
T
phase lag of 290° (270°) in Glacier Bay. The disagree-
ment of the phase lag suggests that the tidal energy dissi- where ρ is seawater density (=1028 kg m–3), T is the tidal
pation due to bottom friction should be reevaluated be- period, γ b is the bottom friction coefficient of the Inazu
cause the phase lag in the shallow region is strongly af- model (=0.02), and u is the horizontal depth-integrated
fected by the bottom friction. velocity. Figure 6(b) shows the M 2 dissipation due to
bottom friction, which is similar to plate 5(a) of F00. Large
4. Energy Flux and Dissipation dissipation is identified at the mouth of Chatham Strait,
The representation of the regional ocean tides in the the overall Sumner Strait, and the entrance to Glacier Bay.
Alaska Panhandle is improved particularly in the vicin- A notable difference is found as larger dissipation around
ity of Glacier Bay as demonstrated in the previous sec- the entrance to Glacier Bay.
tion. F00 suggested that tidal energy dissipation due to The results mentioned above are summarized as fol-
bottom friction is large at the entrance to Glacier Bay lows. The M2 energy flux of 1.5 GW entering through
and Cross Sound (see their plate 5). It is worth Clarence Strait is mainly dissipated at the head of Clarence
reinvestigating the tidal energy dissipation in the Alaska Strait and Sumner Strait. The remainder of the M2 energy
Panhandle owing to the improvement given by the Inazu flux from Clarence Strait is likely to exceed the energy
model. The tidal energy flux is calculated by the follow- flux entering from Sumner Strait because of the mean
ing formula (e.g., Kowalik and Proshutinsky, 1993; Lyard, runoff of 0.5 GW. The M2 energy flux of 2.8 GW enters
1997; Lyard and Le Provost, 1997): through Chatham Strait. The large portion of the energy
flux (2.1 GW) enters through Icy Strait, dissipates at the
F = ρH  u 2 + gη u,
1 entrance to Glacier Bay and Cross Sound, and the remain-
2 
( 4) der of 0.3 GW leaves Cross Sound. Consequently, the M2
tidal energy is dissipated at a rate of 3.4 GW in the Alaska
where ρ is seawater density (=1028 kg m–3), H is the Panhandle. A notable advantage of the Inazu model over
depth, g is the gravitational acceleration (=9.8 m sec–2), F00 is the accurate tidal prediction in the vicinity of Gla-
η is the sea surface elevation, and u is the horizontal cier Bay, which is expected to improve the evaluation of
depth-integrated velocity. The average of Eq. (4) over the the tidal energy dissipation. The evaluated M2 tidal dissi-
tidal period gives the mean energy flux of the constitu- pation is 1.5 times the F00 estimation (2.24 GW). The K1
ent. The mean M2 energy flux over one period is shown tidal dissipation is estimated to be 0.08 GW by the simi-
in Fig. 6(a), which is similar to the result of F00. The lar manner. The value is mostly equivalent to the F00 es-
mean M2 energy flux entering the Alaska Panhandle is timation (0.07 GW) though the accuracy of the K1 con-
calculated by the line integrals similar to F00, which re- stituent is also improved in the vicinity of Glacier Bay as
veals 1.5 and 2.8 GW entering through Clarence Strait well as the M2 constituent.
and Chatham Strait, respectively, and 0.5 and 0.3 GW
leaving through Sumner Strait and Cross Sound, respec- 5. Prediction of Tidal Signal in Geodetic Measure-
tively. Another line integral evaluates 2.1 GW inflow en- ment
tering Icy Strait from Chatham Strait. The results indi- In this section, an efficiency of the Inazu model is

Ocean Tide Modeling in Southeast Alaska 343


verified by predicting the tidal signals in two GPS dis-
placement time series. The GPS sites used here are lo-
cated at a station QUIC under UAF, and a station AB50
under Plate Boundary Observatory (Fig. 1). The observed
tidal signal of GPS consists of two parts. One is a direct
(cm)

0.09
0.11
D
effect called the body tide which is the surface deforma-
tion of the solid Earth by the Newtonian attraction due to
displacement for the M2 and O1 constituents. The phase shows the local phase lag behind the equilibrium tide at each site. D
Table 4. Observed and predicted tidal signals in the vertical displacement at the GPS sites. The amplitude indicates the vertical

the moon and sun. The other is an indirect effect mainly


Phase

293.3
293.4
(deg.)
−6.2

−7.1
−7.3

0.0
due to the oceanic tide loading (OTL). The OTL effect in
O1

southeast Alaska is known to exceed one half of the body


AB50 (134.55°W, 58.42°N)

Amplitude

tide effect (Sato et al., 2008). The accuracy of the ocean


5.74

5.73
5.73

5.38

0.78
0.77
(cm)

tide model is an important issue for investigating the dy-


namics of the solid Earth. We then evaluate the skill of
the Inazu model to predict the tidal signals in the GPS
data by comparing our ocean tide model with F00.
(cm)

0.08
0.06

Vertical and horizontal components of the surface


D

displacement are decomposed from the original GPS data


is the rms difference defined by Eq. (3) between the observed and the predicted tidal signals.

by a “precise point positioning” (PPP) method which was


Phase

187.4
186.8
(deg.)
−8.2

−10.9
−10.2

0.0

initially introduced by Zumberge et al. (1997). The cal-


M2

culation is performed by a software “GpsTools ver. 0.6.3”


(Takasu and Kasai, 2005; Takasu, 2006) which is a GPS/
Amplitude

GNSS (Global Navigation Satellite System) analysis


1.65

1.66
1.63

4.03

2.41
2.44
(cm)

package. The detailed PPP analysis strategies are same


as Sato et al. (2008), which enables one to extract pre-
cise displacement time series from raw GPS data. The
semidiurnal and diurnal constituents in the displacement
(cm)

0.17
0.17
D

time series are calculated by BAYTAP-G.


Based on Farrell (1972), the OTL effect is estimated
293.5
293.3
Phase
(deg.)
−5.9

−7.6
−7.6

0.0

by the convolution integral of mass loading over global


oceans with Green’s function. The preliminary Earth
O1

model of Dziewonski and Anderson (1981) is applied to


QUIC (136.59°W, 58.91°N)

Amplitude

the Green’s function. Here the Earth model assumes the


5.72

5.71
5.71

5.33

0.82
0.82
(cm)

elastic and hydrostatic Earth (Dehant et al., 1999), and


the instantaneous response of the solid Earth to tide-gen-
erating force (e.g., Kantha and Clayson, 2000), namely
zero phase lag of the body tide behind the equilibrium
(cm)

0.09
0.18

tide. The computation of the body tide and the OTL ef-
D

fect to predict the GPS displacement are performed by a


program GOTIC (Sato and Hanada, 1984). The OTL ef-
187.7
184.6
Phase
(deg.)

−17.5
−16.4

−10.7

0.0

fect at the GPS sites is absolutely derived from the Inazu


M2

model or the F00 model in the Alaska Panhandle. The


global FES2004 model is used to estimate the contribu-
Amplitude

1.23
1.32

1.19

3.92

2.77
2.75

tion to the OTL effect from outside the Alaska Panhan-


(cm)

dle.
The observed and predicted tidal signals at the GPS
sites are compared in Table 4. The vertical component of
Inazu model

Inazu model

the displacement is used for the comparison since the


Observation

Body tide

vertical displacement shows the OTL effect larger than


Total

F00

F00

the horizontal displacement at the GPS sites. The skill of


OTL

the prediction is evaluated for the M2 and O1 constitu-


ents. For the diurnal constituent, the O1 constituent is
compared rather than the K1 constituent since it is well
known that the K 1 constituent observed from GPS is
strongly biased associated with the satellite orbital error

344 D. Inazu et al.


(e.g., Hatanaka et al., 2001; Allinson et al., 2004; Choi et than the other region in southeast Alaska as well as seen
al., 2004). in the ocean tide representation. The ocean tide correc-
The result of Table 4 is as follows. In the case using tion in the GPS and gravimeter data is expected to be
the Inazu model instead of the F00 model, the observed performed throughout southeast Alaska to elucidate the
M2 signal at QUIC is accurately predicted while the M2 Earth’s load response and tectonic motions.
prediction may slightly worsen at AB50. This is why the The bottom friction coefficient optimized in the
Inazu model is more accurate in the vicinity of Glacier present study is 0.02 which is larger than the typical value
Bay than the F00 model, and may be a little inaccurate of 0.003. The Alaska Panhandle is formed not only by
near Juneau (Table 3). For the O1 constituent, the accu- the main channel/strait system but also by many small
racy is mostly equivalent at the two GPS sites in the cases inlets which are not resolved by the developed ocean tide
of both the ocean tide models since the Inazu model is model. Specific dissipation mechanisms as mentioned in
just as accurate as the F00 model for the diurnal constitu- Subsection 3.2 which our barotropic model does not ex-
ents. As well as the representation of the ocean tide shown plicitly capture are needed to be clarified to explain fur-
in Table 3, the advantage of the Inazu model over the F00 ther accurate tidal energy dissipation. In southeast Alaska,
model has been demonstrated in terms of the geodetic the glacier yields the tidewater glaciers but does not form
aspect, especially for the M2 constituent and in the vicin- ice shelves (e.g., Meier and Post, 1987). It is worth in-
ity of Glacier Bay. Thus the Inazu model is probably a vestigating effects of sea ice on ocean tides and tidal cur-
prospect of the ocean tide correction for the geodetic data, rents in the Alaska Panhandle as well as Hudson Bay
and is expected to be used for geodetic, seismic, and tec- (Prinsenberg, 1988) and Okhotsk Sea (Ono et al., 2008)
tonic researches in the Alaska Panhandle. though the sea-ice effects may not be robust. The current
observation in Glacier Bay was conducted in summer by
6. Summary and Discussion Cokelet et al. (2004). However, all-year current and/or
An accurate, regional ocean tide model in southeast conductivity-temperature-depth observations are prefer-
Alaska has been developed. The model accuracy for the able in order to investigate the effects of the tidewater
semidiurnal and diurnal constituents is within several glaciers or sea ice on the tidal currents. Observations that
centimeters in the overall Alaska Panhandle including can distinguish brief features of internal structures in each
Glacier Bay with an optimal coefficient of the bottom fric- bay and strait are likely required for quantitative evalua-
tion under accurate conditions of the bathymetry and the tion of the drag due to the internal wave which is em-
boundary tidal elevations. The dependence of the model ployed by Tanaka et al. (2007). In addition, long-term
skill has been evaluated on the boundary conditions and and robust observations of the sea level are also expected.
the dissipation parameters. The model accuracy shows a The duration of the sea level observations in Glacier Bay
crucial sensitivity to the model bathymetry, and also ap- under UAF is less than one month. Table 3 shows that the
parent dependence on the bottom friction coefficient. The observed tidal elevation amplitude of the M2 and K 1 con-
developed ocean tide model is highly improved in the stituents at Composite Island is smaller than that at
vicinity of Glacier Bay compared to F00. The improve- Willoughby Island. On the contrary, the observed ampli-
ment may be thanks to the model bathymetry though the tude of the S2 and O 1 constituents at Composite Island is
detailed reason can not be clarified. larger than that at Willoughby Island (not shown). The
The developed ocean tide model gives an estimation reason for the odd relationship among the semidiurnal
of the M2 tidal energy dissipation due to bottom friction and diurnal tides in Glacier Bay is not clarified but the
that is 1.5 times larger than that of F00 while the K 1 tidal short duration of the observation data may give rise to
energy dissipation is as much as F00. A large portion of misunderstanding. In order to gain more accurate under-
the tidal energy flux entering through Chatham Strait is standing and representation of the ocean tide, the quality
dissipated in the vicinity of Glacier Bay. If the total M2 and quantity of the observation data should be highly
tidal energy dissipation in the Gulf of Alaska is 33.5 GW ensured as much as possible.
as indicated by F00, the M2 tidal energy dissipation in There are thus many essential problems for profound
the Alaska Panhandle estimated by the developed ocean understanding of the ocean tides. Also, stratification and
tide model accounts for 10% of the total energy dissipa- form drag in the inlets can increasingly cause further tidal
tion while the F00 model accounts for 7% of the total energy dissipation (e.g., Stigebrandt, 1999; Edwards et
energy dissipation. Such a difference seems considerable al., 2004; Cushman-Roisin et al., 2005). Local long phase
and may be expected in other regions in the Gulf of delay at Petersburg seems an interesting problem (Table
Alaska. The developed ocean tide model is also shown to 3), which is not appropriately represented by either F00
contribute to more accurate prediction of the tidal sig- or our model. Such highly accurate ocean tide modeling
nals of the GPS time series than the F00 model. The im- postulates a sophisticated grid system (e.g., Chen et al.,
provement is more notable in the vicinity of Glacier Bay 2003) and accurate boundary conditions of bathymetry

Ocean Tide Modeling in Southeast Alaska 345


and tidal elevation, as well as elucidating the dissipation Cont. Shelf Res., 25, 227–242.
mechanisms. It is finally remarked that the ocean Dehant, V., P. Defraigne and J. M. Wahr (1999): Tides for a
modeling should be validated not only by oceanographic convective Earth. J. Geophys. Res., 104, 1035–1058.
but also by geodetic observations. Dziewonski, A. M. and D. L. Anderson (1981): Preliminary
reference Earth model. Phys. Earth Planet. Inter., 25, 297–
356.
Acknowledgements
Edwards, K. A., P. MacCready, J. N. Moum, G. Pawlakd, J. M.
Comments by anonymous reviewers on an earlier Klymake and A. Perline (2004): Form drag and mixing due
version of this manuscript were helpful and constructive. to tidal flow past a sharp point. J. Phys. Oceanogr., 34,
The authors wish to thank Dr. K. Ichikawa of Kyushu 1297–1312.
University for his valuable comments. Thanks also to Dr. Egbert, G. D. and R. D. Ray (2003): Semi-diurnal and diurnal
N. Hirose of Kyushu University for his help with our tidal dissipation from TOPEX/Poseidon altimetry. Geophys.
numerical operations. Most calculations were performed Res. Lett., 30, 1907, doi:10.1029/2003GL017676.
on the NEC SX-6 system at The Institute of Statistical Farrell, W. E. (1972): Deformation of the Earth by surface loads.
Mathematics. Funding was provided by a CREST (Core Rev. Geophys. Space Phys., 10, 761–797.
Research for Evolutional Science and Technology) pro- Foreman, M. G. G., R. F. Henry, R. A. Walters and V. A.
Ballantyne (1993): A finite element model for tides and reso-
gram of Japan Science and Technology Agency, and a
nance along the north coast of British Columbia. J. Geophys.
Grant-in-Aid for Scientific Research (17253003) of
Res., 98, 2509–2531.
MEXT (Ministry of Education, Culture, Sports, Science Foreman, M. G. G., W. R. Crawford, J. Y. Cherniawsky, R. F.
and Technology) of the Japanese government. A US par- Henry and M. R. Tarbotton (2000): A high-resolution as-
ticipant was supported by the National Science Founda- similating tidal model for the northeast Pacific Ocean. J.
tion grant EAR-9870144 and University of Alaska. Geophys. Res., 105, 28629–28651.
Gargett, A. E. (1976): Generation of internal waves in the Strait
References of Georgia, British Columbia. Deep-Sea Res., 23, 17–32.
Akaike, H. (1980): Likelihood and Bayes procedure. p. 1–13. Hatanaka, Y., A. Sengoku, T. Sato, J. M. Johnson, C. Rocken
In Bayesian Statistics, ed. by J. M. Bernardo, M. H. De and C. Meertens (2001): Detection of tidal loading signals
Groot, D. V. Lindley and A. F. M. Smith, University Press, from GPS permanent array of GSI Japan. J. Geod. Soc. Ja-
Valencia. pan, 47, 187–192.
Allinson, C. R., P. J. Clarke, S. J. Edwards, M. A. King, T. F. Hirose, N. and J.-H. Yoon (1996): Barotropic response to the
Baker and P. R. Cruddace (2004): Stability of direct GPS wind in the Japan Sea. Proc. 4th CREAMS Workshop, 39–
estimates of ocean tide loading. Geophys. Res. Lett., 31, 43.
L15603, doi:10.1029/2004GL020588. Ishiguro, M., H. Akaike, M. Ooe and S. Nakai (1983): A
Arakawa, A. and V. R. Lamb (1981): A potential enstrophy and Bayesian approach to the analysis of Earth tides. p. 283–
energy conserving scheme for the shallow water equations. 292. In Proc. 9th Int. Symp. Earth Tides, ed. by J. T. Kuo,
Mon. Wea. Rev., 109, 18–36. E. Schweizerbart’sche Verlangsbuchhandlung, Stuttgart.
Asselin, R. (1972): Frequency filter for time integrations. Mon. Japan Hydrographic Association (1999): Research on bottom
Wea. Rev., 100, 487–490. topography using satellite altimeter data. Japan
Carlson, P. R., P. Hooge, G. Cochrane, A. Stevenson, P. Dartnell Hydrographic Association Survey Research Report, 96, 89
and K. Lee (2002): Multibeam bathymetry and selected pp. (in Japanese).
perspective views of main part of Glacier Bay, Alaska. U.S. Kantha, L. H. and C. A. Clayson (2000): Numerical Models of
Geological Survey Open-File Report 02-391. Available at Oceans and Oceanic Processes. Academic Press, San Di-
http://geopubs.wr.usgs.gov/open-file/of02-391 ego, 940 pp.
Chen, C., H. Liu and R. C. Beardsley (2003): An unstructured Kim, C.-H. and J.-H. Yoon (1996): Modeling of the wind-driven
grid, finite-volume, three-dimensional, primitive equations circulation in the Japan Sea using a reduced gravity model.
ocean model: Application to coastal ocean and estuaries. J. J. Oceanogr., 52, 359–373.
Atmos. Oceanic Technol., 20, 159–186. Kowalik, Z. and A. Y. Proshutinsky (1993): Diurnal tides in the
Choi, K., A. Bilich, K. M. Larson and P. Axelrad (2004): Modi- Arctic Ocean. J. Geophys. Res., 98, 16449–16468.
fied sidereal filtering: Implications for high-rate GPS posi- Larsen, C. F., K. A. Echelmeyer, J. T. Freymueller and R. J.
tioning. Geophys. Res. Lett., 31, L22608, doi:10.1029/ Motyka (2003): Tide gauge records of uplift along the north-
2004GL021621. ern Pacific-North American plate boundary, 1937 to 2001.
Cokelet, E. D., A. D. Jenkins and L. L. Etherington (2004): A J. Geophys. Res., 108, 2216, doi:10.1029/2001JB001685.
transect of Glacier Bay ocean currents measured by acous- Larsen, C. F., R. J. Motyka, J. T. Freymueller, K. A. Echelmeyer
tic Doppler current profiler (ADCP). Proc. 4th Glacier Bay and E. R. Ivins (2004): Rapid uplift of southern Alaska
Science Symp., 80–83. caused by recent ice loss. Geophys. J. Int., 158, 1118–1133.
Colbo, K. (2006): Lateral Reynolds stress and eddy viscosity Le Provost, C. and F. Lyard (1997): Energetics of the M 2
in a coastal strait. J. Phys. Oceanogr., 36, 770–783. barotropic ocean tides: An estimate of bottom friction dis-
Cushman-Roisin, B., A. J. Willmott and N. R. T. Biggs (2005): sipation from a hydrodynamic model. Prog. Oceanogr., 40,
Influence of stratification on decaying surface seiche modes. 37–52.

346 D. Inazu et al.


Lefèvre, F., C. Le Provost and F. H. Lyard (2000): How can we 573–587.
improve a global ocean tide model at a regional scale? A Sandwell, D. T. and W. H. F. Smith (2001): Chapter 12,
test on the Yellow Sea and the East China Sea. J. Geophys. Bathymetric estimation. p. 441–457. In Satellite Altimetry
Res., 105, 8707–8726. and Earth Sciences, ed. by L.-L. Fu and A. Cazenave, Aca-
Levitus, S. and T. P. Boyer (1994): World Ocean Atlas 1994, demic Press, San Diego.
vol. 4, Temperature. NOAA Atlas NESDIS, vol. 4, NOAA, Sato, T. and H. Hanada (1984): A program for the computation
Silver Spring, Md. of oceanic tidal loading effects ‘GOTIC’. Publ. Int. Lat.
Levitus, S., R. Burgett and T. P. Boyer (1994): World Ocean Obs. Mizusawa, 18, 29–47.
Atlas 1994, vol. 3, Salinity. NOAA Atlas NESDIS, vol. 3, Sato, T., S. Miura, Y. Ohta, H. Fujimoto, W. Sun, C. F. Larsen,
NOAA, Silver Spring, Md. M. Heavner, M. Kaufman and J. T. Freymueller (2008):
Lindquist, K. G., K. Engle, D. Stahlke and E. Price (2004): Earth tides observed by gravity and GPS in southeastern
Global topography and bathymetry grid improves research Alaska. J. Geodyn., 46, 78–89.
efforts. EOS, Trans. Am. Geophys. Union, 85, 186. Smith, W. H. F. and D. T. Sandwell (1997): Global sea floor
Lyard, F. (1997): The tides in the Arctic Ocean from a finite topography from satellite altimetry and ship depth sound-
element model. J. Geophys. Res., 102, 15611–15638. ings. Science, 277, 1956–1962.
Lyard, F. and C. Le Provost (1997): Energy budget of the tidal St. Laurent, L. C. and C. Garrett (2002): The role of internal
hydrodynamic model FES94.1. Geophys. Res. Lett., 24, tides in mixing the deep ocean. J. Phys. Oceanogr., 32,
687–690. 2882–2899.
Lyard, F., F. Lefèvre, T. Letellier and O. Francis (2006): Mod- St. Laurent, L. C., H. L. Simmons and S. R. Jayne (2002): Es-
elling the global ocean tides: Modern insights from timating tidally driven mixing in the deep ocean. Geophys.
FES2004. Ocean Dyn., 56, 394–415. Res. Lett., 29, 2106, doi:10.1029/2002GL015633.
Marks, K. M. and W. H. F. Smith (2006): An evaluation of pub- Stigebrandt, A. (1999): Resistance to barotropic tidal flow in
licly available global bathymetry grids. Mar. Geophys. Res., straits by baroclinic wave drag. J. Phys. Oceanogr., 29, 191–
27, 19–34. 197.
Meier, M. F. and A. Post (1987): Fast tidewater glaciers. J. Takasu, T. (2006): High-rate Precise Point Positioning: detec-
Geophys. Res., 92, 9051–9058. tion of crustal deformation by using 1-Hz GPS data. GPS/
Miller, G. R. (1966): The flux of tidal energy out of the deep GNSS Symp. 2006, Tokyo, 52–59.
oceans. J. Geophys. Res., 71, 2485–2489. Takasu, T. and S. Kasai (2005): Development of precise orbit/
Ono, J., K. I. Ohshima, G. Mizuta, Y. Fukamachi and M. clock determination software for GPS/GNSS. Proc. 49th
Wakatsuchi (2008): Diurnal coastal-trapped waves on the Space Sciences and Technology Conference, 1223–1227.
eastern shelf of Sakhalin in the Sea of Okhotsk and their Tamura, Y., T. Sato, M. Ooe and M. Ishiguro (1991): A proce-
modification by sea ice. Cont. Shelf Res., 28, 697–709. dure for tidal analysis with a Bayesian information crite-
Pawlowicz, R. (2002): Observations and linear analysis of sill- rion. Geophys. J. Int., 104, 507–516.
generated internal tides and estuarine flow in Haro Strait. Tanaka, Y., T. Hibiya and Y. Niwa (2007): Estimates of tidal
J. Geophys. Res., 107, 3056, doi:10.1029/2000JC000504. energy dissipation and diapycnal diffusivity in the Kuril
Prinsenberg, S. J. (1988): Damping and phase advance of the Straits using TOPEX/POSEIDON altimeter data. J.
tide in western Hudson Bay by the annual ice cover. J. Phys. Geophys. Res., 112, C10021, doi:10.1029/2007JC004172.
Oceanogr., 18, 1744–1751. Zumberge, J. F., M. B. Heflin, D. C. Jefferson, M. M. Watkins
Rodríguez, G. and M. J. Sevilla (2002): Correlation between and F. H. Webb (1997): Precise point positioning for the
sea surface topography and bathymetry in shallow shelf efficient and robust analysis of GPS data from large net-
waters in the Western Mediterranean. Geophys. J. Int., 150, works. J. Geophys. Res., 102, 5005–5017.

Ocean Tide Modeling in Southeast Alaska 347

You might also like