Você está na página 1de 16

J. Great Lakes Res.

16(3):380-395
Internat. Assoc. Great Lakes Res., 1990
SUDDEN, EXTREME LAKE ERIE STORM SURGES AND THE INTERACTION
OF WIND STRESS, RESONANCE, AND GEOMETRY
Charles Libicki and Keith W. Bedford
t
Ohio State University
Department of Civil Engineering
470 Hitchcock Hall
2070 Neil Avenue
Columbus, Ohio 43210-1275
ABSTRACT. Lake Erie is at times subject to anomalously large storm surges, whose size and
abruptness cannot be adequately explained by standard resonance theory. Here, we model such wind-
generated surges in an attempt to explain their behavior and in particular to describe the influence of
such factors as non-linear resonance dynamics, planform geometry, bottom topography, andfriction
on surge development, climax, and dissipation. Among the results are: Some previously unnoted
aspects of resonant behavior; the very specialized conditions under which resonance occurs; and the
differences in surge behavior between eastward and westward storm paths, which are caused by
planform and frictional differences (the latter of which are directly linked to depth) between the
eastern and western ends of the lake. In particular, the narrowing of the lake in the eastern basin
serves to "focus" the surge, and produce an exaggerated response at resonance, significantly greater
than that at both sub- and super-resonance. The non-dimensionalized characteristics approach used
here allows explicit analytic solution in many of the idealized linear causes studied, so the effects of
individual terms can be discerned. In the non-linear case, it allows explicit tracking of shock evolu-
tion and dissipation. The results are extended to simulations which make use of actual Lake Erie
planform and topography data.
INDEX WORDS: Storm surges, Lake Erie, physical processes, wind-driven currents, wind setup,
resonance.
INTRODUCTION
As is well known, the passage of weather fronts
over Lake Erie causes significant storm surges. The
forced response component of the surge is usually
initiated during the passage of the wind shear or
pressure gradient over the lake. Very high water
levels occur at the downwind end of the lake, while
significant drawdowns occur at the opposite end.
The free mode response results after passage of the
atmospheric disturbance and consists of a series of
six Kelvin wave oscillation modes. Due to friction,
seiche activity dies away after several seiche peri-
ods. Reviews of the structure, climatology, and
forecasting of such events are found in Platzman
(1963), Irish and Platzman (1962), Pore et al.
(1975), and Schwab (1978), respectively. With
regard to Figure 1, Lake Erie is oriented in a south-
west to northeast direction, is very much longer
ITo whom correspondence should be addressed.
380
than wide, and consists of three successively deeper
basins, progressing toward Buffalo, New York.
The overall shallowness plays a significant role in
the surge response characteristics of the lake, as
does the planform shape at either end of the lake.
Rao (1967), in an important paper describing the
interaction between the wind shear and the lake
response, focused on the role of resonance between
the moving wind shear gradient and the long wave
propagation within the basin. The resonance the-
ory went a long way toward explaining why certain
weather events created unexpectedly large surges
while others with the same wind intensities moving
in the same direction produced no significant
surge. Rao, by using the linear long wave equa-
tions and the method of characteristics, examined
the response of a constant depth lake with per-
fectly reflecting ends for a step jump wind stress
propagating across the lake at a speed of V*. Reso-
LAKE ERIE STORM SURGES
FIG. 1. Lake Erie with depth contours in m.
I
N
/
381
nance was defined as the condition existing when
the wind stress gradient propagation speed equals
the long wave celerity, c* = .Jiii*, where h* is the
lake depth and g is the acceleration due to gravity.
Subresonant and superresonant cases are defined
for v* < c* and V* > c*, respectively. When c* is
not a constant, due to variable depth or nonlinear
interactions, resonance can be defined as the con-
dition that the time of travel of the front and the
free gravity wave are equal. Rao focused on the
relationship between the maximum setup and the
ratio of V* I c*, where setup is defined as the
instantaneous difference between the water surface
elevations at the downwind and the upwind ends of
the lake.
During the last several years, J. S. Dingman and
these authors (Bedford et af. 1980, Bedford and
Dingman, 1982, Dingman and Bedford 1984) have
been comparing various storm surge models for
their suitability in predicting such surges. Model
calculations have been compared to storm surge
data collected from up to fourteen gauges sur-
rounding Lake Erie, and in reviewing each surge
and the accompanying meterological data, several
aspects of the surge development are curious.
First, as the positive surge develops, a rather large,
sometimes sudden, increase in amplitude or "flip"
occurs very rapidly in a relatively confined space
near the basin end. In the notable storm surge of
April, 1979 (whose conditions were reviewed in
Hamblin, 1979), this flip approximately doubled
the surge height. Secondly, these flips only occur at
the Buffalo end of the lake, they do not occur for
positive storm surges at Toledo. Third, it is appar-
ent that, of the actual storms surveyed to date,
resonance as defined above does not appear to be a
significant cause of many of the severe surges.
Examination of the three worst storm surges in
each of the three dominant weather tracks reveals
that only one of the nine, the April 1979 surge, had
significant resonance effects. Neither of the first
two curious aspects are discussed in Rao's work, or
even more recent work, while the relative lack of
importance of resonance is contrary to expecta-
tions.
The objective of this paper is to address the
above problems by extending Rao's approach and
investigating a number of phenomena excluded
from his analysis. These new phenomena include
the effects of the inertia terms, bottom friction,
variable bottom topography, and variable plan-
form topography. We follow Rao's approach of
propagating a simple wind shear across the lake at
constant speeds, but the analysis differs in that
interest is focused not on the setup, but rather on
the time and space development of the maximum
positive surge as well as its amplitude at the down-
wind end. It is at this end of the lake that the
physics is the most complicated and must be under-
stood in order to accurately forecast the surge
flooding.
The mathematical analysis used in support of
this paper tends to be somewhat involved, whereas
the intent of the paper is to develop the physical
ramifications of the expressions. Rather than
impede the narrative, therefore, the relevant math-
ematics is placed in the accompanying Tables. In
this way verbal summaries of the mathematical
382 UBICKI and BEDFORD
FIG. 2. Model and variable definition sketch.
results can proceed without interruption except for
the most immediately relevant mathematical
statements.
GOVERNING EQUATIONS
Lake Erie is modeled as a one-dimensional system;
that is, vertically and laterally averaged. This can
be justified by its shallowness (Fig. 1) and its rela-
tively narrow width. Indeed, studies of the free
oscillations of Lake Erie (Dingman and Bedford
1984, among many) found that only the sixth mode
shows any transverse character, and is just barely
observable in field data. Also lost to the one-
dimensional analysis is the Coriolis effect. But as
Rao (1967) pointed out, the Coriolis effect simply
transforms a longitudinal standing wave into an
amphidromic (rotating) one, leaving both the
amplitude and period of oscillation virtually
unchanged.
With regard to Figures 1 and 2, and Table 1 and
2, the variables of interest are x*, the distance from
the upwind edge of the lake; t*, the time; B*, the
"channel" top-width; h*, the total average depth, or
the instantaneous area divided by the top-width; t*
the height of the water surface above the still water
level (SWL); g, the acceleration due to gravity; u*,
the area average velocity; and c* = -Vgh* ,
the speed of the free gravity wave. In addition,
there are p, the density of water; T, a measure of
wind stress; and k*, a frictional parameter.
Using standard techniques (Liggett 1975, for
example), a pair of equations describing the con-
servation of mass and conservation of momentum
can be derived in terms of non-dimensional vari-
ables. The equations are presented in Table 1,
To solve the system of equations in their various
forms, the method of characteristics is used.
Descriptions of this procedure are found in Stoker
(1957), Lax (1973), Liggett (1975), and Abbot
(1979) for the general nonlinear cases. Rao's work
was based upon a characteristics solution of the
linear equations. The solution is a method by
which the partial differential equations are rede-
fined as ordinary differential equations along tra-
jectories called characteristics. For this problem,
characteristics take on special significance as the
path of the tip of the main surge coincides with the
main forward characteristic.
The generalized solution is determined by find-
ing the eigenvalues and eigenvectors of the result-
ing pair of governing equations. The eigenvectors
are called Riemann invariants; these are the quan-
tities which are strictly conserved along the charac-
teristics. Table 3, Eqs. 3.3 and 3.4, represent the
invariants and the corresponding characteristics. It
is noted that there are two characteristics: dx/dt =
u c, respectively the forward (+ ) and backward
(-) characteristics, and in general they have curvili-
near paths. For the special linear case considered
by Rao, the characteristics are straight lines with
slopes of 1 in the non-dimensional system.
The general solution to the equations (3.3, 3.4)
GENERAL CHARACTERISTICS SOLUTION
while the nondimensionalization formalism is
detailed in Table 2. The key similarity parameter is
the time, T, it takes a free gravity wave to propa-
gate the length of the basin, L, at a speed of c*,
i.e., c = 1.0. A velocity of zero is set as the bound-
ary condition at either end of the lake in solving
the governing equations, and all their special
forms. Therefore, no energy is lost through the
boundaries.
A number of special cases exist, as noted in
Table 1. For a constant width basin, Term 1 of the
continuity equation, which accounts for the varia-
tion in width, is set to zero. The cases of no wind
stress, constant depth, and no friction remove,
respectively, Terms 2, 3, and 4 from the momen-
tum equation. The special case of small amplitude
disturbances allows a linearization of the equations
(Eqs. 1.3 and 1.4 in Table 1) wherein the velocity u
is considered small with respect to celerity c and
the inertia term (Term 5) is unimportant.
As noted in Table 1, if Terms 1, 3, 4, and 5 are
unimportant, then the original model used by Rao
(1967) is recovered.
SWL
Buffalo
d
x =1
- - - - infinite length
stress band
-- finite stress band
I
a ~ V
- - - - - - - - - - -'"'-=-=-=-=--=-=-=-=-
Toledo
t
wind stress
o
LAKE ERIE STORM SURGES
TABLE 1. General governing equations. All equations are nondimensionalized as per definitions in
Table 2, see also Figures 1 and 2.
1. Full Nonlinear Equations
383
Mass conservation:
Momentum conservation:
2. Small amplitude Form
Mass:
Momentum:
ah aM -M aB
-+-=-
at ax B ax
(1)
aM + a(uM) + c
2
ah = R + c
2
aH _ ku
at ax ax ax h
(5) (2) (3) (4)
ah aM -M aB
-+-=--
at ax B ax
(1)
aM ah aH u
- + c
2
- = R + c
2
- - k -
at ax ax h
(1.1)
(1.2)
(1.3)
(1.4)
3. Simple Linear Form
i.) No friction: Term 4 = 0
ii.) Flat bottom: Term 3 = 0
iii.) Rectangular shape: Term 1 = 0
iv.) No inertia: Term 5 = 0
Mass:
Momentum:
a ~ + aM = 0
at ax
aM = c
2
at = { R inside windfield
at ax 0 outside windfie1d
(1.5)
(1.6)
of the Riemann invariants and their corresponding
trajectories is found by constructing a suitably
dense net or grid through the solution space. At the
net points, where forward and backward charac-
teristics meet, both of their respective equations
hold, and the system is solved simultaneously.
THE IDEALIZED LINEAR CASE-
A BRIEF SUMMARY
Using the linear characteristic equations (4.1, 4.2,
Table 4) Rao (1967) obtained the surge height at
the downwind end, the setup from one end of the
lake to the other (6n, and the residual energy
remaining in the fundamental mode after passage
of the wind stress event. Two types of wind stress
events were considered (see Fig. 2), a stress of
semi-infinite spatial extent and hence duration,
propagating over the lake at a constant rate of
non-dimensional speed V, and a finite stress band
of nondimensional spatial extent, ex, also travelling
at fixed speed, V. The width ex may be greater or
less than the length of the lake (i.e., ex > 1 or ex
< 1). The transition from no stress to full stress
(and back again) is assumed, as in Rao's work, to
take place instantaneously. (The alternative, a
gradual transition from no stress to full stress,
might seem more realistic. But the slope would
then become another free variable and obscure the
relations we are trying to elucidate.)
It is seen from Table 4 that the characteristics are
straight lines with a slope of 1. If, as in Figure 3,
the slope of the forward characteristic emanating
from the left hand edge of the lake (Toledo), where
the wind stress is first felt, is compared to the
inverse propagation speed of the front (itself a
slope), it is seen that three types of conditions are
defined: subresonant, resonant, and superreso-
nant. In the superresonant case, the wind stress
band is essentially travelling to Buffalo at a faster
speed than the free gravity wave in the lake created
by the stress can propagate to Buffalo. In subre-
384 LIBICKI and BEDFORD
TABLE 2. Nondimensional variables and definitions. The asterisk denotes the real variable; please
refer to Figures 1 and 2 and Table 1 for further information.
x* =: distance measured from upwind k* =: frictional parameters
end of lake
R* = ~
g
t* =: time
7* =: wind stress
B* =: top width of lake
p =: water density
h* =: total cross-section average depth
M* =: u* h*
u* =: cross-section average velocity r* =: water surface height above the
still water level
g =: gravitational acceleration
H* =: depth to bottom from the still water level
c* =: ~ * =: speed of free gravity wave
Nondimensional Definitions:
L =: lake length r =: r*gT2/V
T =: Jl (c*t
1
dx H =: H*gP/V
o
x = x*/L h = h*gT2/V
t =: t*/T
U =: u*T/L
C =: c*T/L
B =: B*/L
R
k
sonance, the situation is reversed, while resonance
is the case of simultaneous arrival of both at
Buffalo.
Several other results from Rao's work are of
interest here:
1. The maximum setup for the semi-infinite stress
band occurs at V ~ 00.
2. The maximum surge for the semi-infinite stress
band occurs at V ~ 1.
3. For 1 < V-I < 2, the non-dimensional surge
height is 2VI(l + V).
4. For V < 1, the surge height is based upon Vand
the number of complete excursions that the free
gravity wave has been able to make since the
onset of the front.
5. The maximum setup for a stress band of finite
extent occurs somewhere in the interval 0 < V-I
< 1, depending on the size of Q:.
SUDDEN SURGE HEIGHTS - THE
LINEAR CASE
As mentioned above, the interest here lies in
explaining how the lake surface changes both in
time and space as the wind-stress band approaches
the downwind end. The first result is found by
extending Rao's work. At resonance, to the right
of the main characteristic (the one originating at x
= 0, t = 0) there is no wind stress and M = r = o.
To the left of the main characteristic, wind stress is
applied. All backward characteristics intersecting
the main characteristic have a right-hand side of
zero and, therefore, M = r. Using this relation in
the forward characteristic equation yields d/dt2M
= R or by integration, M(t) = ret) = Rl2t. At the
boundary M must go to zero, but if we look at the
time increment just before the intersection with the
boundary (1 = Of), it is shown that r(l) = ro -
LAKE ERIE STORM SURGES
385
TABLE 3. General method of characteristics formulation and solution. See Lax 1973, Stoker 1957,
Liggett 1975, Abbot 1979, and Figure 3.
Mass:
1. General Equation Form (nondimensional)
oc oc c ou
- + u- + - - = r
l
ot ox 2 ox
(3.1)
Momentum:
ou oc ou
- + 2c - + u - = r
2
ot ox ox
2. System Eigenvalues
Al = u + c forward characteristic
dx
- dt
(3.2)
d .. dx
- backward charactenstIc = -d
- t
3. Conservation Properties
d
for AI: dt<u+2c) = r
2
- r
1
d
for 1..
2
: -(u-2c) = r
2
+ r
1
dt
dx
along dt = u+c
dx
along - = u-c
dt
(3.3)
(3.4)
4. Solution: numerical for most cases where r
1
or r
2
* 0
TABLE 4. Rao's (1967) solution and resonance classes
for the linearized case (see Fig. 3).
01) + M(l -01) = R. In other words, at resonance,
the surge height doubles in the small instant of
time just before it arrives at the downstream
boundary. This sudden increase in surge height at
resonance we call the resonance "flip," and it is
notable that simple linear formulations are satis-
factory in explaining the sudden height increase.
In the superresonant case, V = 00, the surge
evolution takes more time (i.e., more of the extent
of the lake) to develop. In summary, at a point P
(where t = x = 112) where the main backward and
forward characteristic intersect, the elevation is
zero and M = RI2. As the surge approaches the
downwind end, however, M is exchanged for t
until M = 0 and t = R. Therefore, the "flip" that
occurs instantaneously in the resonance case now
takes half the length of the lake to complete. The
ultimate surge heights for the resonant and super-
resonant cases are identical. The more realistic
superresonant case has V < 00, therefore, the
point P lies somewhere between x = 112 and 1 (see
Fig. 3), but t still grows at the expense of M until
M = 0 and t = R at x = 1.
(4.1)
(4.3)
(4.4)
(4.5)
(4.2)
dx
R along - =
dt
dx
R along - = -1
dt
d
=
dt
d
=
dt
in stress band
r2 = stress band
4. Classes of resonance based on wind stress speed,
V (see Fig. 3) for linear case above.
a. sub-resonant V< c ; -> V< I
b. resonant V=c ; -> V= 1
a. super-resonant V> c ; -> V> 1
3. Characteristics Solution:
2. General Characteristics Formula:
r
1
= 0
1. Condition: linearized, flat bottom, straight
channel and frictionless
386
x-
Subresonance
FIG. 3. Resonance regimes.
LIBICKI and BEDFORD
x-
Resonance
x-
Superresonance
BOTTOM TOPOGRAPHY AND FRICTION
Accounting for the effects of bottom topography
and friction calls for a more physically realistic
model of lake activity, as is employed iQ current
Lake Erie surge forecasting (Schwab 1978). Both
terms enter into the equations on the right hand
side (Tables 5 and 6), and so do not fundamentally
affect the characteristics formulation or solution
procedure. The characteristic paths are no longer
straight lines but are curvilinear, defined as in Eqs.
5.1,5.2,6.1, and 6.2 in Tables 5 and 6. Because r
TABLE 5. The effects of bottom topography; the lin-
earized case.
1. Conditions: straight channel, no friction,
linearized, variable bottom
topography
2. Characteristics Solution:
d dH dx
-d -- + R along - = c (5.1)
t 2 dx dt
d
d = dH + R along dx = -c (5.2)
t 2 dx dt
characteristics are curves, i.e., straight or concave
up or down
now appears on the right hand side, the final surge
height depends upon its time history and the equa-
tions cannot be directly integrated. Another obser-
vation is that a channel that grows shallower from
the upwind to the downwind end would exhibit a
concave upward main characteristic. This would
lie entirely below the line V = 1, and might not
achieve its peak surge until well above resonance.
Conversely, in a channel that generally deepens
from the upwind end down, the path of the main
characteristic would tend to be concave downward
and, therefore, lie entirely above the line V = 1.
This would imply that resonance-like behavior
might indeed be achieved at resonant speeds or
TABLE 6. The friction effects; the linearized case.
1. Conditions: linearized, variable bottom
topography, straight channel,
includes friction
2. Characteristics Solution:
d dH ku dx
= "2 dx + R - h along dt = c (6.1)
d dH ku dx
di(M+ = "2 + R - h along <it = -c (6.2)
LAKE ERIE STORM SURGES 387
below. Lake Erie, whose western basin is much
shallower than the central or eastern basins, would
be expected to show the latter behavior in the more
common southwest to northeast storms (SW) and
the former behavior in the less frequent northeast
to southwest storms (NE).
An additional factor in the difference between
resonant and superresonant storms is based upon
an earlier discussion. The energy imparted by the
wind stress goes into both the velocity and the
surge height, but as the surge approaches the
boundary the energy is transferred from the veloc-
ity to the surge height, suddenly in the resonant
case, gradually in the superresonant case. Even in
the vicinity of the boundary, therefore, r would be
smaller in the resonant case than the superresonant
because the exchange of kinetic for potential
energy occurs only at the very last instant in the
resonant case. From Long Point to Buffalo, the
final leg of the SW storm's transit, the lake
becomes dramatically shallower. The negative
slope, according to the above equation, acts as a
drain on the conserved quantities. This drain is
proportional to r, which as stated above, would be
depressed at resonance. One would expect, there-
fore, that the final surge height would be greater at
resonance than at superresonance.
Numerical simulations for the linear case (and
nonlinear formulations discussed further on) were
performed with bottom topography data taken
from average depths across 80 nearly evenly spaced
transects along Lake Erie. Simulations were run
for 0.6 ::5 V ::5 1.8. Although not our primary
interest here, simulations at V = 00 were also per-
formed as a theoretical check against Rao's (1967)
results. In every case, they agreed with the results
at V = 1.8 to within a few percent. The SW simu-
lations (Table 7) show a peak at slightly above res-
onance, and a drop-off at higher propagation
speeds (the surge height is 15070 less at V = 1.8
than at V = 1.05). The maximum surge height
over all simulations is 2070 less than for the flat-
bottom case. The nonlinear variable bottom simu-
lations show a maximum surge height and arrival
time virtually identical to the flat-bottom case
(although r is 15070 smaller at V = 1.8). The linear
NE simulations, by contrast, show surge height
over all simulations nearly 35070 greater than for
the flat-bottom case. The nonlinear simulations
exhibit the maximum surge at V = 1.2 with a fall-
off of 7070 by V = 1.8.
The friction formula and the value of the coeffi-
cient, k* = 100 cm
2
/sec, are taken from Schwab
(1978). The expression enters the right hand side of
the momentum and characteristics equations, both
linear and nonlinear, without any effect upon the
form of the invariants. The coefficient is scaled
against the major dimensions of the problem as k
= k*g2L/P = 1.036.
Continuing to build on the experience of the ide-
alized cases, the items of interest are the histories
of u and r along the main forward characteristic.
In the idealized case, the main forward characteris-
tic for superresonance is the first one, for subre-
sonance it is the one that arrives at t = V-I. The
main characteristic is simply the one associated
with the maximum surge. These simulations focus
on r rather than c, as the latter contains the effects
of bottom topography as well as surge height.
Figure 4 is a trace of u and r agains x along the
main characteristic for the SW case with no fric-
tion. The effects of bottom topography are espe-
cially evident in the velocity trace, which peaks at
shallows in the lake and dives where the water
deepens. And yet, the overall picture corresponds
well to the idealized case, with u and r gradually
increasing until near the end, where u must go to
zero and r doubles in the resonance flip. The effect
of friction (Fig. 5) is to flatten out the trace, extend
the distance over which the flip occurs, and in the
end produce a smaller surge at a later time.
The same general pattern is noted (Fig. 6) in the
nonlinear NE cases with u and r increasing gradu-
ally to the boundary and u showing local peaks at
shallows in the basin. The flip is greatly exagger-
ated, however, doubling in the last percent of
travel to an even greater height than predicted by
the flat-bottom system. With the addition of fric-
tion, the maximum appears, not at the impinge-
ment of the first characteristic, but later at T =
1.16 (Fig. 7). Frictional resistance, as modeled
here, consumes 20070 of the surge height in the SW
simulations with both the linear and nonlinear
models. In contrast, the NE simulations have fric-
tional losses of 45070 with the linear model and
31070 with the nonlinear model (Table 7). The dif-
ference, predominantly due to the shallowness of
the western basin, points out how sensitive NE
simulations are to the way frictional losses are
modeled. The notion may be extended to other
losses as well, such as vertical acceleration and
absorbing boundaries, neither of which is modeled
here, but both of which are aggravated in western
basin. Indeed, at the eastern end of the lake, the
prediction of a resonance flip depends strongly on
the presence of a reflecting boundary, whereas at
388 LIBICKI and BEDFORD
TABLE 7. Surge height, time of arrival, and associatedpropagation speedfor various conditions. a) Linear Model;
b) Nonlinear Model, R* = 10 dynes per cm
1
(R = 0.1372); c) Nonlinear Model, R* = 20 dynes per cm
1
(R = 0.2744).
Cases
Constant Width, Flat Bottom:
i. V =
ii. resonance
iii. resonance, shock solution
Constant Width, Variable Bottom
i. Southwest
no friction
friction
ii. Northeast
no friction
friction
Variable Width, Variable Bottom
i. Southwest
no friction
friction
ii. Northeast
no friction
friction
T V /R
a) b) c) a) b) c) a) b) c) a) b) c)
1.0 /0.131/0.252 1.0 /0.953/0.914 1.0 /0.954/0.918
1.0 /0.131/0.251 1.0 /0.95110.917 1.0 /0.954/0.915
/0.129/0.247 /0.977/0.996 /0.942/0.901
0.977/0.1311* 0.997/0.952/* 1.05 /1.1 /* 0.977/0.954/*
.764/0.105/* 0.979/0.963/* 1.05 /1.05 /* 0.764/0.763/*
1.344/0.164/0.308 0.999/0.957/0.932 1.8 /1.2 /1.2 1.344/1.195/1.122
0.876/0.113/0.216 1.358/1.16110.903 1.8 /1.8 /1.8 0.876/0.823/0.788
1.427/0.185/* 0.979/0.925/* 1.05 /1.1 /* 1.427/1.3511*
1.206/0.157/0.309 0.898/0.917/0.844 1.15 /1.1 /1.2 1.206/1.144/1.126
1.555/0.186/0.345 1.025/0.958/0.950 1.2 /1.2 /1.15 1.555/1.355/1.256
0.925/0.116/0.230 0.999/1.006/0.978 1.8 /1.2 /1.2 0.925/0.848/0.839
*simulation fails due to negative depths
the western end of the lake, the shore is more akin
to a flood plain.
VARIABLE WIDTH
Variable width can have a profound influence on
the development of surges, especially near reso-
nance. The term accounting for variable width
appears in the continuity equation, but not in the
momentum equation when presented in non-
conservation form. The derivative of width is
divided by the width and so non-dimension-
alization is automatic. The term enters the linear
characteristic equations in Table 8.
The linear equations can be solved explicitly in
the case of resonance. As before, the back charac-
teristic intersects the main forward characteristic
o
Toledo Buffalo
FlG. 4. Surge height (--) and velocity (....) versus
distance for the dominant characteristic - SWstorm, no
friction.
FlG. 5. Surge height (--) and velocity () versus
distance for the dominant characteristic-SW storm,
friction.
LAKE ERIE STORM SURGES 389
o
immediately from the quiescent zone when M = r
= 0, and so, as before, M = r. Along the forward
characteristic dx/dt = 1, so the derivative in x can
be replaced by one in t. Consider a basin that is a
constant width until a distance d from the down-
wind end, and then narrows linearly to a value of a
at the end. This case is solved as with the constant
width case up to x = 1 - d. At that point, dB/dx
takes on a constant negative value and the equa-
tion for the height becomes
2 [ ( 3 ) ( 1 ) -1/2 ]
~ = 3R 2(1 - d) + db 1 - b d(b - 1) (1)
Toledo Buffalo
FIG. 6. Surge height (--) and velocity (...) versus
distance for the dominant characteristic- NE storm, no
friction.
which, after expanding in a binomial series, gives
~ = R + R 2 - d + R(3 - 4d/3) + . .. (2)
4 b 16b
2
1. Conditions: variable channel width, linearized,
flat bottom, no friction
TABLE 8. The effect of variable width; the linear
response.
In these equations, b = 1/(1 - a). To both first and
second order, therefore, the surge height is
increased by constriction of the basin at the down-
ward end and also by delaying the onset of this
constriction (although, as the length over which
the constriction takes place grows too small, reflec-
tion begins to dominate over focusing (Lamb
1983.
To calculate surge heights for other propagation
speeds, both families of characteristics must be
considered, and so a numerical solution is per-
formed. The results are presented in Figure 8. The
basin is taken as having a constant width from x =
oto x = 0.75 and then narrowing linearly to the
end with a = 0.2, which approximates the narrow-
ing of Lake Erie from Long Point to Buffalo. At
resonance, the surge height is more than doubled
over the rectangular case, while at infinite propa-
gation speed it is augmented by only 100/0.
Full linear and nonlinear simulations were per-
formed with the basin shape taken from 80 tran-
sects of Lake Erie as for the bottom topography.
The results are shown in Table 7 and Figure 8. As
in the idealized case, the surge is increased with the
addition of variable width, although by a factor of
only 1.5 in the frictionless case and 1.6 in the case
with friction (1.4 and 1.5, respectively, in the non-
linear model), rather than double. This is to be
expected, in part due to the width's varying along
the entire length, in part due to the lake's shallow-
ing as it narrows toward Buffalo, and in part due
to the inexact matching between a fixed propaga-
tion speed and a varying wave speed. Again, the
maximum effect is felt at resonance. The surge
Buffalo
(7.2)
(7.1)
d M JB dx
dt(M+~ ) = R - Ba;{ along dt =
d (M r) R M JB I dx -I
dt -" = +Ba;{ a ong dt =
Toledo
o
2. Characteristics Solution:
FIG. 7. Surge height (--) and velocity (....) versus
distance for the dominant characteristic- NE storm,
friction.
390 LIBICKI and BEDFORD
2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4
2 . 5 +-__-.L ....l..-__-.L ....I..-__----'- ....l..-__----L _\_ 2 . 5
".--....
o
c
a 2.0
en
c
Q)
E
"TI 1. 5
c
a
c
'-../
4
2.0
1.5
1: 1.0
CJ)
Q)
I
Q) 0.5
0'1
I-
:::J
if)
2
5
_____________ _
__________________ _
1.0
0.5
2 1.8 1.6 1.4 1.2 1 0.8 0.6
o. 0 0 .0
0.4
Front Propagation Speed (nondimensional)
FIG. 8. Maximum surge height versus front propagation speed for (1) idealized rectangular basin, linear system;
(2) idealized rectangular basin nonlinear system; (3) idealized narrowing basin, linear system; (4) full nonlinear
system, SWstorm; and (5) full nonlinear system, NE storm.
height at V = 1.8 is 31 % less than at V = 1.1
(32070 in the nonlinear model).
The NE case is affected much less strongly by
variable width (by 10% in both the friction and
frictionless cases) as is to be expected given less
narrowing in the western basin.
THE NONLINEAR SYSTEM
In the nonlinear system (Table 3), surge height and
velocity are not simply proportional to the wind
stress R; the absolute value of R is important. Here
ris fixes at 10 dynes/cm
2
, which yields R = R*P/
D = 0.1372. Additional simulations were per-
formed with a wind stress of 20 dynes/cm
2
,
approximating the April 1979 storm.
The concept of resonance in the nonlinear sys-
tem is not as direct as in the linear system. As the
surge builds, U and c increase, which in turn
increases the speed of the characteristic. At V = 1,
therefore, the main forward characteristic arrives
ahead of the front and so behaves like subreson-
ance. Exact resonant coupling would then have the
wind front progress at exactly the same speed as
the main characteristic. Back characteristics meet
this main characteristic immediately upon crossing
the front. The Riemann invariant of the back char-
acteristic is then identical to its value in the quies-
cent zone, as in the linear case. The solution indi-
cates that just before the surge arrives
= c
2
- 1 = (1 + Rt ) 213 - 1 (3)
and once it hits
The resonance flip more than doubles the surge
LAKE ERIE STORM SURGES 391
and substituting it into a binomial expansion for ~
gives
height. Taking a Taylor's series expansion of T
(time of arrival) as a function of R
~ = R ( 1 - ~ R ) + O(R3) (6)
This states that to first order of R, the nonlinear
resonance solution corresponds to the linear solu-
tion, and to second order of R, the travel time is
reduced and surge height increases, although not in
proportion to R.
The difficulty with the above analysis is that as
the speed of the main forward characteristic
increases, it overtakes the straight characteristics
of the same family in the quiescent zone. At that
point, according to the above analysis, the solution
is double-valued and appeal must be made to "gen-
eralized" or "weak" solutions. These solutions take
into account formation of a shock, a point across
which the solution passes abruptly from one state
to another, where derivatives are unbounded, and
yet the integral form of the governing equations is
still satisfied in a weak sense. Physical examples of
shocks include sonic booms during a airplane's
transition to supersonic flight and, of more rele-
vance here, hydraulic jumps and bores (Stoker
1957). Forward shocks occur when at any point the
characteristic speed dx/dt is greater on the left
than on the right, which is clearly the case at the
front in the exact resonance case. Therefore, at
resonance in Lake Erie, a shock is created immedi-
ately and is continually aggravated as the wind
proceeds along the lake. Attention has not been
drawn to this fact in the previous literature on this
problem.
A shock, physically manifested by a hydraulic
jump when dealing with the shallow water wave
equations, is formally predicted here in the case of
perfect resonance, although actually observing one
in storm conditions would be nearly impossible.
Still, one-dimensional modeling that takes it into
account has its value if only to determine a practi-
cal effect on the surge formation and height. Lax
(1973) discussed how the addition of viscous
damping to the equations does not affect the prop-
agation of the shock nor the conserved quantities
on either side, but merely thickens the transition
zone. In this way, a physical jump may not be in
COMPARISON TO STEADY STATE
In posing the question, does resonance produce an
exaggerated dynamic response, we may counter-
exaggerated with respect to what? An appropriate
baseline is the steady-state response, where the
wind has built up slowly and has been blowing long
evidence, but a jump in the conserved quantities
over a larger distance would be.
The method of characteristics, as laid out above,
cannot account for the effects of a shock in the
solution space. We must return to the conservation
form of the governing equations in u and c, invoke
the Rankine-Hugoniot relations for stating how
variables vary across the shock, and construct a
numerical solution. A solution is constructed mak-
ing use of the standard characteristics method in
the area of smooth solution, the shock relations
where necessary, and the Lax admissibility or
entropy condition (Lax 1973).
At the chosen value of R, the resulting surge
height is 1.3070 less than for the solution that
ignores shock formation, and the time of arrival is
2.6% greater (Table 8). As might be expected, the
fraction of the surge height consumed in the shock
increases with increasing R.
Simulations at other than resonant conditions
were performed without taking shock formation
into account. The maximum surge is produced at V
= 1.05, which is slightly above resonance. The
surge heights at subresonance are significantly
lower and at superresonance are not. For superre-
sonance, the time of the maximum surge coincides
with the arrival of the first characteristic at the
downwind boundary (which occurs approximately
5% earlier than in the linear case). For subreson-
ance, the maximum surge occurs when the wind
front arrives at the downwind boundary. Surge
heights are reduced with respect to the linear sys-
tem by approximately 5%.
Incorporating the extra effects of bottom topog-
raphy, friction, and variable width into the nonlin-
ear formulation proceeds in much the same way as
with the linear system. In particular, width varia-
tion appears in the continuity equation only, which
is common to both the linear and the nonlinear
system, and so the term that accounts for variable
width appears on the right hand side, with opposite
sign in each family, as with the linear system. The
results of adding the extra factors are presented in
Table 7, and were discussed previously, each factor
in its respective sub-section.
(5) T = 1 - ~ R + O(R2)
392 LIBICKI and BEDFORD
'0
+-
Q)

Q)

.... ".
\. 6 April 1979
/
\. .. :/
.........
'.
.... .
.....
enough for friction to damp out all transients. (An
alternative baseline is V = <Xl, or V is some value
off resonance. But these were discussed previously,
and are displayed graphically in Figure 8.) Under
steady state, the velocity and all derivatives with
respect to time go to zero and the equations
become hdrldx = R and = O. This system
predicts a ratio of water surface displacement to
wind stress (t/ R) of 0.42 at Buffalo for the SW
wind and 0.61 at Toledo for NE. Referring Figure
8, the dynamic displacement at resonance for the
SW storm is 172010 greater than the steady state
displacement at R = IOdynes/cm
2
By contrast,
the dynamic displacement is only 40010 greater than
the steady state for the NE storm. Even this figure
is likely an overestimate due to the questionable
assumption of a reflective boundary at the western
end of the lake. For comparison, the flat, rectan-
gular, frictionless, linear system predicts a 1000/0
exaggeration due to dynamic effect.
CASE HISTORIES
-to
The clearest and perhaps only example of resonant
coupling occurred in the storm of 6 April 1979
(Hamblin 1979). By contrast, the greatest positive
surge at Toledo on record was produced by a storm
that did not exhibit the features of resonance
described above (Strommen 1973).
The storm of 6 April 1979 was the most severe
storm of the winter for the Great Lakes region. It
brought winds of up to 125 km/h, freezing rain,
waves of up to 5 m, and property damages in the
millions of dollars. It produced the largest Toledo-
to-Buffalo setup ever recorded, going back to 1900:
4.1 m at 0700 EST 6 April, and one of the largest
surge amplitudes ever recorded at Buffalo: 2.4 m.
With reference to the data in Figures 9 and 10
(Hamblin 1979), the storm was marked by the
sharpness of the wind stress front, across which
temperatures also dropped from 5C to 2C. At
Buffalo, for example, the axial component of the
wind stress jumped from 2 to 18 dynes/cm
2
in the
space of 1 hour. The front progressed from west to
east, arriving at Cleveland at 2200 5 April and Buf-
falo 5 hours later, or approximately at resonance
speeds.
The lake, in turn, exhibited many of the features
of a resonant response. The exaggerated surge is
mentioned above. The time of the maximum surge
coincided with the arrival of the peak wind stress.
With reference to Figure 10, the surge grew as it
moved across the lake and then more than doubled
.. . .
......... ..... "\.
..... -"
" ",
......... ",
.... .......
............. -...-
18 00 12 18
14 November 1972
FIG. 9. Water level at Buffalo (--) and Toledo (..)
for the April 1979 and November 1972 storms.
in its last hour of travel, producing the resonance
flip.
An additional feature is readily explained by
characteristic analysis:
Curiously, despite the fact that winds tended to
peak earlier in the western end of the lake, the
maximum water level excursion occurred about six
hours earlier at the eastern end of Lake Erie than at
the western end (Hamblin 1979).
As indicated above, for a sharp wind front, the
maximum surge is associated with the first charac-
teristic to spend its entire transit beneath the wind
LAKE ERIE STORM SURGES
393
'20
//i/
,-
~ ~ ~ -, .. '
'20
o
-5
-20
-20
I
Toledo
200 6 April 1979
Buffalo Toledo
300 6 April 1979
Buffalo
'5 ------- --------------
o
'20
-20
Toledo
4'00 6 April 1979
Buffalo
FIG. 10. Estimated water surface profiles (--) and wind stress profiles (....) for the April 1979 storm.
stress. The statement also holds for the maximum
negative surge and the first back characteristic.
Such a characteristic takes off from the downwind
border just before the resonance flip and arrives at
the upwind border approximately 7 hours (L/c)
later. Also note (Figs. 9 and 10) that the sudden
increase and gradual decay at Buffalo is mirrored
in a gradual decrease and sudden recovery (albeit
not quite as sudden) as Toledo.
NE storms, as noted above, are much less sensi-
tive to resonant coupling. The worst storm on
record, in terms of water level rise at Toledo,
occurred in November 1972 (Fig. 9b). The storm
center, which first consolidated over the Texas
Panhandle, moved within 40 hours to a position
over southeast Ohio generating maximum overlake
winds of 115 km/h out of the NE. The waters at
Toledo began to climb at 800 EST 13 November
and reached a peak of 1.2 m above the mean 26
hours later (Strommen 1973). Considering the rela-
tively slow buildup of both wind and water it is
safe to say resonance played no role in this surge.
394 LIBICKI and BEDFORD
CONCLUSIONS
By extending, augmenting, and re-examining pre-
vious storm surge analyses, it is apparent that the
interaction of geometry and friction plays an
extremely important role in the development of the
surge at the ends of Lake Erie. Wind stress reso-
nance appears to playa much reduced role in such
development, but when it does occur, leads to sud-
den and dramatic surge development in the eastern
end of the lake. The following generalizations also
can be made.
The maximum surge is typically associated with
the first characteristic to spend its entire transit
beneath the wind stress. This implies that at subre-
sonant speeds, the maximum surge coincides with
the arrival of the stress front at the downwind
boundary, and at superresonant speeds, the time
corresponds to the transit time for the free gravity
wave.
In the idealized case (flat, rectangular basin, no
friction), surge height increases with stress band
propagation speed until reaching a plateau at reso-
nance. This is true for both the linear and nonlin-
ear models. Resonance is characterized by a sud-
den exchange of kinetic energy for surge height at
the downwind boundary, and results in a flip in the
water surface profile.
The chief effect upon the main surge of includ-
ing the nonlinear terms (at typical and stress val-
ues) is to decrease the surge height and time of
arrival by approximately the same fraction. At res-
onance, the surge height more than doubles at the
downwind boundary rather than exactly doubling,
as with the linear model. Formation of a hydraulic
shock is predicted at exact resonance, although the
surge height dissipated in the shock zone is small.
The inclusion of variable bottom topography
generally exaggerates the surge height in basins
that shallow from upwind to downwind and dimin-
ishes surge height in the reverse case. In the former
case, the maximum surge height may not be
achieved until well above resonance, whereas in the
latter, surge heights at superresonance are dimin-
ished with respect to resonance.
Friction diminishes surge heights and thickens
the shock transition region. The large role played
by frictional losses in storms progressing from
northeast to southwest indicates the importance of
properly accounting for these and other losses,
such as nonreflective boundaries, in this type of
storm.
Variable width can serve to focus and thereby
exaggerate surge heights by 40 to 500/0. This exag-
geration is especially pronounced at resonance at
the Buffalo end of the lake. Indeed, this focusing is
the major factor for the dominance of resonance
over both subresonance and superresonance in
these types of storms.
The criteria for resonance are fairly strict: a
sharp wind front that remains intact for most of
the length of the lake, wind and propagation veloc-
ities that are aligned with the lake axis, and a prop-
agation speed that closely matches the free gravity
wave speed. Most destructive storms do not meet
these criteria and so resonance is not a major con-
tributing factor. When such a storm does occur, as
it did in April 1979, resonance can greatly exagger-
ate both the intensity and the suddenness of the
surge.
ACKNOWLEDGMENTS
This research was financed in part by National
Oceanographic and Atmospheric Administration,
Ohio Sea Grant contract NA79AA-D-00120 and
NA80AA-D-0100, and Cray Research, Inc. Their
support is appreciated.
REFERENCES
Abbot, M. B. 1979. Computational Hydraulics: Ele-
ments of the Theory of Free Surface Flows. Boston,
Massachusetts: Pitman Publishing, Ltd.
Bedford, K. W., and Dingman, J. S. 1982. Model inter-
comparison of Lake Erie storm surge resonance pre-
dictions. In Applying Research to Hydraulic Prac-
tice, pp. 661-673. New York: Hydraulics Division,
American Society of Civil Engineers.
____ , Prater, M., and Dingman, J. 5.1980. Great
Lakes storm surge flood forecasting methods. In
Proceedings of the National Symposium on Urban
Stormwater Management in Coastal Areas, pp.
289-298. New York: Hydraulics Division, American
Society of Civil Engineers.
Dingman, J. 5., and Bedford, K. W. 1984. The Lake
Erie response to the January 26, 1978 cyclone. J.
Geophys Res. 89 (C4):6427-6445.
Hamblin, P. F. 1979. Note: Great Lakes storm surge of
April 6, 1979. J. Great Lakes Res. 5:312-315.
Irish, S. M., and Platzman, G. W. 1962. An investiga-
tion of the meterological conditions associated with
extreme wind tides on Lake Erie. Mon. Weather Rev.
90:39-47.
Lamb, H. L. 1932. Hydrodynamics. New York, New
York: Dover Publications.
LAKE ERIE STORM SURGES
395
Lax, P. D. 1973. Hyberbolic Systems of Conservation
Laws and the Mathematical Theory of Shock Waves.
Philadelphia: Society for Industrial and Applied
Mathematics.
Liggett, J. A. 1975. Basic equations of unsteady flow.
In Unsteady Flow in Open Channels, K. Mahmood
and V. Yevjevich, eds. Ford Collins, Colorado:
Water Research Publications.
Platzman, G. W. 1963. The dynamical prediction of
wind tides on Lake Erie. Meteorol. Monogr. 4(26):
1-141.
Pore, N. A., Perotti, H. P., and Richardson, W. S.
1975. Climatology of Lake Erie storm surges at Buf-
falo and Toledo. NOAA Tech. Memo. NWS TDL
54, Washington, DC: Nat'l. Ocean. and Atmos.
Admin.
Rao, D. B. 1967. Response of a lake to a time-
dependent stress, J. Geophys. Res., 27:1697-1708.
Schwab, D. J. 1978. Simulation and forecasting of
Lake Erie storm surges. Mon. Weather Rev. 106:
1476-1487.
Stoker, J. J. 1957. Water Waves: The Mathematical
Theory with Applications. New York, New York:
lnterscience.
Strommen, N. D. 1973. Fall storm and high lake levels
spell disaster around the Great Lakes. Mariner's
Weather Log 17:66-69.

Você também pode gostar