16
Tsunami Squares Approach to Landslide-Generated Waves: Application to Gongjiafang Landslide, Three Gorges Reservoir, China LILI XIAO, 1 STEVEN N. WARD, 2 and JIAJIA WANG 1 Abstract—We have developed a new method, named ‘‘T- sunami Squares’’, for modeling of landslides and landslide- generated waves. The approach has the advantages of the previous ‘‘Tsunami Ball’’ method, for example, separate, special treatment for dry and wet cells is not needed, but obviates the use of millions of individual particles. Simulations now can be expanded to spatial scales not previously possible. The new method accelerates and transports ‘‘squares’’ of material that are fractured into new squares in such a way as to conserve volume and linear momentum. The simulation first generates landslide motion as constrained by direct observation. It then computes induced water waves, given as- sumptions about energy and momentum transfer. We demonstrated and validated the Tsunami Squares method by modeling the 2008 Three Gorges Reservoir Gongjiafang landslide and river tsunami. The landslide’s progressive failure, the wave generated, and its subsequent propagation and run-up are well reproduced. On a laptop computer Tsunami Square simulations flexibly handle a wide variety of waves and flows, and are excellent techniques for risk estimation. Key words: Tsunami squares, numerical modeling, landslide, waves in reservoirs. 1. Introduction Large, high-velocity landslides entering confined water bodies (reservoirs, lakes, fiords, and rivers) may cause potentially extreme tsunami run-up that in- creases the variety, range, and severity of effects attributable to the landslide alone. Alaska’s 1958 Li- tuya Bay tsunami caused by a subaerial landslide has attracted the attention of scholars, and a series of ex- perimental and numerical modeling research projects has been reported (BASU et al. 2010;FRITZ et al. 2009; WEISS et al. 2009). Two-dimensional cross-sections of the landslide and the proximal wave were modeled numerically by MADER and GITTINGS (2002) and QUE- CEDO et al. (2004) with full Navier–Stokes hydrodynamic codes. WARD and DAY (2010) devel- oped a novel ‘‘Tsunami Ball’’ method, which they demonstrated and validated by simulating the land- slide and the propagation of the tsunami to the seaward end of Lituya Bay. WARD (2014)(https:// www.youtube.com/watch?v=6COeNRToYqU) revis- ited this case by using the Tsunami Squares method described herein. Other examples of well-documented rockslide tsunami include the 1934 Tafjord event in Norway, for which HARBITZ et al. (2014a, b) recently applied nu- merical models (denoted GloBouss and DpWaves) based on Boussinesq equations. A similar Boussinesq model (Geowave) has also been applied to several examples of landslide-generated waves, mainly for open-ocean submarine landslides (WATTS et al. 2003; POISSON and PEDREROS 2010;WATTS and TAPPIN 2012). We have developed a new method, named ‘‘T- sunami Squares’’, for modeling of landslides and landslide-generated waves. Tsunami Squares can simulate landslide evolution, wave propagation, the moving mass of impacting water, and the triggering wave process. Tsunami Squares is suited both to long- term hazard assessment in regions where many thou- sands of possible landslides might need to be modeled, for example China’s Three Gorges Reservoir, and to emergencies in which accelerating creep of an unsta- ble slope has been detected and emergency managers require a rapid assessment of the tsunami hazard. 2. Overview of the Gongjiafang Landslide On the afternoon of November 23, 2008, 120 km up-river of China’s Three Gorges Dam and 4 km 1 Engineering Faculty, China University of Geosciences (Wuhan), Wuhan, China. E-mail: [email protected] 2 Institute of Geophysics and Planetary Physics, University of California, Santa Cruz, USA. Pure Appl. Geophys. Ó 2015 Springer Basel DOI 10.1007/s00024-015-1045-6 Pure and Applied Geophysics

Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

  • Upload
    others

  • View
    13

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

Tsunami Squares Approach to Landslide-Generated Waves: Application to Gongjiafang

Landslide, Three Gorges Reservoir, China

LILI XIAO,1 STEVEN N. WARD,2 and JIAJIA WANG1

Abstract—We have developed a new method, named ‘‘T-

sunami Squares’’, for modeling of landslides and landslide-

generated waves. The approach has the advantages of the previous

‘‘Tsunami Ball’’ method, for example, separate, special treatment

for dry and wet cells is not needed, but obviates the use of millions

of individual particles. Simulations now can be expanded to spatial

scales not previously possible. The new method accelerates and

transports ‘‘squares’’ of material that are fractured into new squares

in such a way as to conserve volume and linear momentum. The

simulation first generates landslide motion as constrained by direct

observation. It then computes induced water waves, given as-

sumptions about energy and momentum transfer. We demonstrated

and validated the Tsunami Squares method by modeling the 2008

Three Gorges Reservoir Gongjiafang landslide and river tsunami.

The landslide’s progressive failure, the wave generated, and its

subsequent propagation and run-up are well reproduced. On a

laptop computer Tsunami Square simulations flexibly handle a

wide variety of waves and flows, and are excellent techniques for

risk estimation.

Key words: Tsunami squares, numerical modeling, landslide,

waves in reservoirs.

1. Introduction

Large, high-velocity landslides entering confined

water bodies (reservoirs, lakes, fiords, and rivers) may

cause potentially extreme tsunami run-up that in-

creases the variety, range, and severity of effects

attributable to the landslide alone. Alaska’s 1958 Li-

tuya Bay tsunami caused by a subaerial landslide has

attracted the attention of scholars, and a series of ex-

perimental and numerical modeling research projects

has been reported (BASU et al. 2010; FRITZ et al. 2009;

WEISS et al. 2009). Two-dimensional cross-sections of

the landslide and the proximal wave were modeled

numerically by MADER and GITTINGS (2002) and QUE-

CEDO et al. (2004) with full Navier–Stokes

hydrodynamic codes. WARD and DAY (2010) devel-

oped a novel ‘‘Tsunami Ball’’ method, which they

demonstrated and validated by simulating the land-

slide and the propagation of the tsunami to the

seaward end of Lituya Bay. WARD (2014) (https://

www.youtube.com/watch?v=6COeNRToYqU) revis-

ited this case by using the Tsunami Squares method

described herein.

Other examples of well-documented rockslide

tsunami include the 1934 Tafjord event in Norway, for

which HARBITZ et al. (2014a, b) recently applied nu-

merical models (denoted GloBouss and DpWaves)

based on Boussinesq equations. A similar Boussinesq

model (Geowave) has also been applied to several

examples of landslide-generated waves, mainly for

open-ocean submarine landslides (WATTS et al. 2003;

POISSON and PEDREROS 2010; WATTS and TAPPIN 2012).

We have developed a new method, named ‘‘T-

sunami Squares’’, for modeling of landslides and

landslide-generated waves. Tsunami Squares can

simulate landslide evolution, wave propagation, the

moving mass of impacting water, and the triggering

wave process. Tsunami Squares is suited both to long-

term hazard assessment in regions where many thou-

sands of possible landslides might need to be modeled,

for example China’s Three Gorges Reservoir, and to

emergencies in which accelerating creep of an unsta-

ble slope has been detected and emergency managers

require a rapid assessment of the tsunami hazard.

2. Overview of the Gongjiafang Landslide

On the afternoon of November 23, 2008, 120 km

up-river of China’s Three Gorges Dam and 4 km

1 Engineering Faculty, China University of Geosciences

(Wuhan), Wuhan, China. E-mail: [email protected] Institute of Geophysics and Planetary Physics, University

of California, Santa Cruz, USA.

Pure Appl. Geophys.

� 2015 Springer Basel

DOI 10.1007/s00024-015-1045-6 Pure and Applied Geophysics

Page 2: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

downstream of Wushan County (Figs. 1, 2), the

Gongjiafang landslide slipped from the north shore of

the Yangtze and generated a wave that crossed the

river and washed the far bank to a height of 13 m.

The tsunami damaged the Wuxia town docks 3.5 km

upstream, some navigation aids, roads, orange trees,

and a few small boats. Luckily, no ships were passing

through the busy river corridor, and no one was hurt.

The direct economic loss was limited to 800,000

USD (HUANG et al. 2012).

The Gongjiafang landslide happened after a test

impoundment brought the Three Gorges Reservoir to

a new high water level of 172.8 m. The landslide was

located on a scarp slope on the west of Hengshixi

anticline and the east of Wushan syncline (WU et al.

2010). The slide mass consisted of thinly-bedded soft

marl stones intersected with thick limestone and

dolomite limestone. Because of strong weathering

and joint damage, the mass had been fractured into

small rocks and vegetated soils. The high shale con-

tent, large developed fissures, and strong weathering

resulted in numerous discontinuities in the rock. In

November, the lower part of the slide mass was

submerged as the reservoir level rose to its new high

level. The clays in the marlstone and shale softened

easily, aggravating a decrease of rock strength. After

being submerged for a long period, the shallow rock

structure at the slope toe began to fail along tension

joints and the upper dry part soon followed.

3. Tsunami Squares Theory

The two-dimensional Tsunami Squares approach

evolved from the ‘‘Tsunami Ball’’ method which

proved to be effective for simulating flow and flow-

like movements (WARD and DAY 2005, 2008,

2010). Tsunami Squares inherits advantages of the

Figure 1Location of Gongjifang Landslide. The lowest map shows the area 6 km up and down river from the landslide in which the tsunami was

modeled. Populated Wushan County, on the left, is the area affected

L. Xiao et al. Pure Appl. Geophys.

Page 3: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

Tsunami Ball method, for example not needing

special treatment for dry versus wet cells. Moving

materials in the Tsunami Squares method are,

however, made from divisible squares, which ob-

viates the need for millions of individual particles.

Simulations can be extended to spatial scales not

previously possible.

3.1. Mass Flow and Wave Propagation

Typical tsunami calculations for earthquake and

landslide-generated tsunamis (SATAKE and TANIOKA

1995; LIU et al. 2003; TITOV and GONZALEZ 1997)

solve non-linear, long wave continuity, and momen-

tum equations for variations in water column

thickness H(r,t) and depth averaged horizontal water

velocity v(r,t) at points r = (x,y)

oHðr; tÞot

¼ �rh � vðr; tÞHðr; tÞ½ � ð1Þ

and

oHðr; tÞvðr; tÞot

¼ �rh � vðr; tÞvðr; tÞHðr; tÞ½ �� gHðr; tÞrhfðr; tÞ ð2Þ

or

ovðr; tÞot

¼ �vðr; tÞ � rhvðr; tÞ � grhfðr; tÞ

where g is the acceleration of gravity, rh is the

horizontal gradient, f(r,t) is the elevation of the wa-

ter, and t is time. (To relax the long wave assumption

and include wave dispersion, replace Eq. (2) by

Eq. (22) in Appendix 1. For our purposes here, water

wave dispersion is unnecessary.)

For small steps dt, Eqs. (1) and (2) become:

Hðr; t þ dtÞ ¼ Hðr; tÞ � rh � vðr; tÞHðr; tÞ½ �dt ð3Þ

and

Hðr; t þ dtÞvðr; t þ dtÞ ¼ Hðr; tÞvðr; tÞ� rh � vðr; tÞvðr; tÞHðr; tÞ½ �dt

� gHðr; tÞrhfðr; tÞdt

ð4Þ

Tsunami Squares solves equations that are

equivalent to these, but by use of a new approach.

It considers a regular set of N square cells with

dimension Dc (the length of the uniform squares) and

center points ri = (xi,yi). At time t, each cell holds

material (water or landslide mass) of thickness

Hi(t) = H(ri,t), with mean horizontal velocity

vi(t) = v(ri,t) and mean horizontal acceleration

ai(t) = a(ri,t) (Fig. 3, top left). For locations outside

the flow, Hi(t) would be zero. The entire concept of

wave propagation or mass flow involves updating

those conditions to time t ? dt, where dt is some

small time interval.

Pick one cell, for example i = 10 (red square,

Fig. 3, top right). With its known velocity and

acceleration, displace the cell material to a new

center point:

~ri ¼ riðtÞ þ viðtÞdt þ 0:5aiðtÞdt2

¼ riðtÞ þ 0:5 ½viðtÞ þ ~vi�dt

¼ ½~xiðtÞ; ~yiðtÞ�ð5Þ

and give it a new mean velocity:

~vi ¼ viðtÞ þ aiðtÞdt

¼ ½~vxiðtÞ; ~vyiðtÞ�ð6Þ

we wish to partition the volume and linear momen-

tum of the material in the displaced cell among the

N original cells. It is logical that the partitioned

volume of the displaced ith cell into the jth original

cell is:

Figure 2Landforms at Guongjiafang after the slide (Yichang Center of

China Geological Survey). The slide stopped funnel-shaped on the

riverbed, mainly distributed below 200 m but with thin deposits

above 220 m. The numbers in the red boxes show the positions

where slide velocities were tracked during modeling

Tsunami Squares Approach to Landslide

Page 4: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

dVji ¼ ðHiD2cÞ 1�

~xi � xj

��

��

Dc

� �

1�~yi � yj

��

��

Dc

� �

if~xi � xj

��

��

Dc

\1 and~yi � yj

��

��

Dc

\1; otherwise dVji ¼ 0

ð7Þ

The product of the two terms on the right of Eq. (7)

is simply a statement of the fractional area overlap of

the ith displaced cell with the jth fixed cell. Clearly

there is no need to run the partitioning through all

j = N cells because, at most, only four fixed cells

overlap the displaced cell (Fig. 3, bottom left). More-

over because ~ri is known and the cells are square, it is

simple to determine which four cells overlap.

Partitioning of vector linear momentum follows in

the same way:

dMji ¼ ðqwHiD2c ~viÞ 1�

~xi � xj

��

��

Dc

� �

1�~yi � yj

��

��

Dc

� �

if~xi � xj

��

��

Dc

\1 and~yi � yj

��

��

Dc

\1; otherwise dMji ¼ 0

ð8Þ

The N updated thickness Hj(t ? dt) and velocity

vj(t ? dt) values in the jth fixed cell comes from

summing and normalizing the volume (Eq. 7) and

momentum (Eq. 8) contributions from all i displaced

cells:

Hjðt þ dtÞ ¼PN

i¼1 dVji

D2c

ð9Þ

vjðt þ dtÞ ¼PN

i¼1 dMji

qwD2cHjðt þ dtÞ ð10Þ

Figure 3Tsunami Squares wave propagation and/or flow simulation concept

L. Xiao et al. Pure Appl. Geophys.

Page 5: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

Because only four of dVji and dMji are non-zero

for each i, the sums in Eqs. (9) and (10) involve

4N terms (not N2). Actually, there may be fewer than

4N terms because there is no need to displace and

partition cells that are dry.

This process has time-stepped a wave propagation

and/or flow simulation on a fixed set of cells while:

1. Conserving material volume exactly. It is possible

to verify that the sum of the four non-zero

partitioned volumes dVji in Eq. (7) is equal to

(HiDc2), the volume of the displaced cell. Equa-

tion (9) simply replaces the ‘‘continuity equation’’

common to most numerical approaches (SATAKE

AND TANIOKA 1995; LIU et al. 2003; HEINRICH 1992;

WALDER et al. 2003) by tracking material from one

cell to another.

2. Conserving linear momentum exactly. dMji ¼qwdVji ~vi sums to the final momentum of the

material in the ith cell. By tracking momentum

from one cell to another, Eq. (10) replaces the

advected component �rh � vðr; tÞvðr; tÞHðr; tÞ½ �in the momentum equation (Eq. 2). For high-speed

landslides and flows, conservation of momentum

is critical for constructing realistic simulations.

For deep water wave propagation, however, fluid

velocities are small and advected momentum is

negligible. This aspect of Tsunami Squares can be

linearized by replacing Eq. (10) by vjðt þ dtÞ ¼ ~vj

where ~vj is determined by use of Eq. (6)

3. Requiring no special treatment of dry cells or any

mention of topography.

4. Reducing a N2 summation to a 4N summation.

5. Obviating the need for a single numerical

derivative.

To show that Eqs. (9) and (10) are equivalent to

Eqs. (3) and (4), Eq. (9) is evaluated with dt very

small (dt � 1) so that vx(r,t)dt � Dc and

vy(r,t)dt � Dc for all cells, and terms with dt2 are

-0

H(rj;t þ dt) = H(rj;t)� H(rj;t)vxðrj;tÞ��

��

Dc

+vyðrj;tÞ��

��

Dc

� �

dt

þXN¼j

i¼1

H(rj,t)vxðri;tÞj j

Dc

� �

dt +XN¼j

i¼1

H(rj,t)vyðri;tÞ��

��

Dc

� �

dt

ð11Þ

The first RHS term in Eq. (11) is the original

material in cell j. The second RHS term in Eq. (11) is

the material originally in cell j that has moved into

adjacent cells. The sums in Eq. (11) only include

those cells in the x and y directions that overlap the

jth cell after their displacements (Eq. 5). They

account for the material originally in adjacent cells

that has moved into cell j. Another statement of

(Eq. 11) is:

Hðrj; t þ dtÞ ¼ Hðrj; tÞ � r � vðrj; tÞHðrj; tÞ� �

dt

ð12Þ

Hence, Eq. (11) is exactly equivalent to Eq. (3)

for vanishingly small dt. We can evaluate Eq. (10) in

the same way:

H(rj;t þ dt)v(rj;t) = H(rj;t)~v(rj;t)� H(rj;t)v(rj;t)

�vxðrj;tÞ��

��

Dc

+vyðrj;tÞ��

��

Dc

� �

dt

+XN 6¼j

i¼1

H(rj,t)v(rj,t)vxðri;tÞj j

Dc

� �

dt

+XN 6¼j

i¼1

H(rj,t)v(rj,t)vyðri;tÞ��

��

Dc

� �

dt

ð13Þ

The first RHS term in Eq. (13) is the original

momentum of cell j. The second RHS term in

Eq. (13) is momentum originally of cell j that was

transferred to adjacent cells. The sums in Eq. (13)

account for momentum originally in adjacent cells

transferred to cell j. Equation (13) is equivalent to the

first three terms in Eq. (4) for vanishingly small

dt. Tsunami Squares updates flow velocities through

the slope of the surface (third RHS term in Eq. (4)) as

a separate step.

Tsunami Squares satisfies the same non-linear

continuity and momentum equations used in tradi-

tional methods, but does so in a way that is more

intuitive (by just moving cell mass and momentum by

appropriate partitioning).

3.2. Gravitational Acceleration

To complete the time step, it is necessary to

update mean cell accelerations ai(t). As is customary

Tsunami Squares Approach to Landslide

Page 6: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

in ‘‘long wave’’ theory, the mean acceleration of

material in the cell is proportional to the slope of

upper surface f(ri,t):

aiðtÞ ¼ agðri; tÞ ¼ �grhfðri; tÞ¼ �grh TðriÞ þ Hðri; tÞ½ �

ð14Þ

H(ri,t) = Hi(t) is material thickness found above,

T(ri) is the fixed topography over which it moves, g is

the acceleration of gravity, and rh is the horizontal

gradient (Fig. 4).

The rhf(ri,t) in Eq. (14) is the only step in which

a numerical derivative must be evaluated. Even this

sole differentiation can, however, be avoided by

fitting a plane to f(ri,t) and its eight adjacent

neighbors, then fixing the horizontal gradient from

the slope of that surface. This plane-fitting approach

helps stabilize the calculation by estimating the

gradient across a two-dimensional region as a func-

tion of adjacent points alone. Another advantage of

the plane-fitting approach is the ability to ‘‘punch

out’’ specific locations near ri by excluding them

from the fit. Where wet cells are near dry ones, the

dry sites would normally be ‘‘punched out’’ during

calculation of the slope of the surface f(ri,t): For

example, if a steep dry cliff was adjacent to a wet

area, the cliff surface would not be included when

computing the slope of the fluid.

In most cases the computation grid is made

sufficiently large that flows or waves do not reach the

ends of the domain in the time period of interest. If

need be, an ‘‘absorbing buffer zone’’ can be included

near domain walls to minimize reflections.

4. Landslide Simulation

A difficulty in simulating landslides is that they

behave partly like a solid, because of a cohesive

fraction, and partly like a fluid, because they flow into

valleys and channels (COUSSOT and MEUNIER 1996).

Landslides have solid characteristics when they ini-

tially fail and when they near a stop (HUNGR and

MCDOUGALL 2009). In between, after a few seconds of

acceleration and before the final seconds of de-ac-

celeration, slide material loses cohesion and takes on

fluid characteristics. This is why Tsunami Squares

does not differentiate fluids from moving slide ma-

terial. For landslide material considered solid during

the first few seconds, the driving force on the square

depends on the slope of the bottom surface. For

landslide material regarded as fluid-like, the driving

force on the square depends upon the slope of the top

surface (Eq. 14).

4.1. Frictional Acceleration

Two types of frictional acceleration are added to

Eq. (14) to oppose sliding, one due to basal friction

(ab) and one due to dynamic friction (ad):

aiðtÞ ¼ agðri; tÞ þ adðri; tÞ þ abðri; tÞ ð15Þ

Basal friction is the simple static resistance of the

sliding surface because of interactions between

moving materials and the rough bed. It depends on

material type, solid fraction, bed roughness, and

normal stress. The acceleration of a cell of material as

a result of basal friction is:

Figure 4Geometry of slide mass and water flow. Tsunami Squares considers moving masses more fluid-like than solid, so the upper surface horizontal

gradient drives the flow for both the slide and the water

L. Xiao et al. Pure Appl. Geophys.

Page 7: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

abðri; tÞ ¼ �lbgv̂slideðri; tÞ ð16Þ

Here, v̂slide is the unit velocity vector and lb is the

basal friction coefficient. Basal friction is treated as a

solid-like moving resistance.

Dynamic friction originates from resistance en-

countered by the moving bulk material through air or

water. Dynamic resistance is intrinsically velocity-

dependent. ‘‘V-square’’ friction, as it is called,

originates from pressure or viscous-like forces acting

on the top and bottom surfaces of moving slides of

thickness H(ri,t):

adðri; tÞ ¼ �ldvðri; tÞ vðri; tÞj j=Hðri; tÞ ð17Þ

Here, ld is the dynamic friction coefficient that

expresses all velocity-dependent properties of particle

interaction. Because dynamic friction increases as

|v2|, it imparts a terminal velocity to motion, with

thicker slides attaining a higher speed. It is weak

during landslide initiation and stoppage, but is

dominant for high-speed moving masses.

Both basal and dynamic friction act to slow the

slide, but are not allowed to reverse the sliding

direction. ld and lb can be functions of time and

space and made as complicated as necessary. For

example, ld of subaerial slides might be less than ld

for subaqueous slides; lb might change from a low

value to a high value as the velocity of the slide falls

below a critical value (WARD and DAY 2006).

Alternatively, the transition could be determined by

an ‘‘angle of repose hrepose’’ of the solidifying

material. When the slope of the top surface drops

below a specific angle, basal friction grows.

Transition between two types of friction enables

Tsunami Squares to embrace the dual behavior of

landslides. When slides move at high speeds, fluid

behavior and dynamic friction dominate and the slope

of the upper surface drives the motion. When slow or

stopped, solid behavior and basal friction dominate.

4.2. Observational Constraints on the Landslide

The landslide model incorporates several obser-

vational constraints: the topography of the bed, the

initial and final landslide shape, and the initial and

final landslide thickness. The Yangtze River, at an

elevation of 173 m in this region, has a generally

parabolic cross-section 470 m wide and 122 m deep.

Freshly exposed slip surfaces on the north bank

reveal the isosceles triangular outline of the landslide

(Fig. 2). The slide was 40 m wide at the top and

160 m wide at the water surface. It formed an upslope

from 120 to 400 m above sea level (350 m above the

deepest part of the river bed) with a horizontal

distance of 290 m from the shoreline (Fig. 5), and

dipped steeply with upper and lower slope angles of

63� and 44�. From measured centerline values, we

interpolated landslide thicknesses across the slide

face, arriving at a total volume of 380,000 m3 and an

average thickness of 17 m. Slide deposits were

distributed at elevations from 210 to 50 m above

sea level, with the thickest part at approximately

90 m; the toe of the landslide deposit is at the floor of

the river bed.

We are also fortunate to have an eyewitness video

of the landslide. This shows that the lower part of the

slide slipped first, followed by the upper part three

seconds later, and that the whole sliding event lasted

approximately 31 s. According to field investigation

the landslide is divided into two parts, below and

above 240 m. The program imposes sequential fail-

ure that keeps the upper part fixed for the first 3 s and

then releases it.

4.3. Landslide Simulation Results

4.3.1 Model Setup

We ran the simulations on 10 m digital topography

that defines the dimension of the cells, Dc = 10. In

general, to maintain reasonable resolution in

Eqs. (5)–(10), it is best to select a time step, dt, such

that vrep 9 dt \ Dc where vrep is some representative

peak velocity of the flow. Because vrep is unknown

before the fact, experimentation is needed to select

dt. In these calculations we took dt = 0.5 s. For

values of friction we set dynamic friction ld equal to

0.05 and 0.2 for landslide material moving on dry

land and under water, respectively. For basal friction

lb gradually changes from 0.2 to 6.0 when the slope

angle becomes less than the 31� angle of repose

measured from the final landslide shape. These values

of friction were derived after several runs to best fit

the observed field data.

Tsunami Squares Approach to Landslide

Page 8: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

4.3.2 Model Results

Figure 6 shows instantaneous profiles of the middle

cross-section and plane view of the slide at four times

from start to finish (a Quicktime animation of the

landslide is available at http://es.ucsc.edu/*ward/

Yangtze1.mov). Rainbow colors indicate sliding ve-

locity. At t = 4 s, the speed of the middle part

exceeded that of the upper part, a consequence of the

sequential failure assumption. At t = 10 s, the upper

part, now moving faster than the lower parts, helps to

push the whole mass deeper into the river. The model

slide moved downward for 26 s, stopping on the

riverbed with a uniform angle of repose and only a

thin layer depositing above water (Fig. 6, lower right

and left), consistent with the behavior of the real

landslide (Fig. 5).

Impact velocity and landslide shape are critical

variables for calculation of wave heights in the

laboratory (FRITZ et al. 2003). We tracked slide-

impact velocities near the river surface and on the

centerline above and below water (five squares in

Fig. 2 (right) and five circles in Fig. 6 (left)).

Figure 7 shows the velocity curves of the tracked

spots. The orange line that follows the above-water

portion at 240 m elevation (#4) has the highest peak

speed of 21 m/s. For the initial four seconds,

however, speeds there are actually slower than in

other places before increasing sharply to peak at

approximately nine seconds. The initial delay reflects

the applied sequential bottom-to-top failure (Fig. 8).

Figure 5Engineering geological profile of the Guongjiafang landslide. The initial and final surfaces provide the thickness of the central section and

residuals

Figure 6Moving landslide profiles at t = 0, 4, 10 and 30 s. For the plane

views (left), background colors brown and red indicate land above

and underwater. The blue to white shading indicates the thickness

of the landslide, thick to thin. The arrows indicate slide direction

and speed. Arrow colors follow the rainbow legend. The red spots

mark the positions where velocities are being tracked in Fig. 7.

Points 1, 2, and 3 are at the water surface (173 m). Points 4 and 5

are above and under water, at 240 m and 140 m. For the cross

sections (right), the rainbow colors represent landslide speeds from

0 to greater than 15 m/s. The lower part of the slide has a higher

speed during first few seconds. The upper part moves fastest later

because of higher gravitational potential energy. A Quicktime

animation of the landslide is available at http://es.ucsc.edu/*ward/

Yangtze1.mov

c

L. Xiao et al. Pure Appl. Geophys.

Page 9: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

Tsunami Squares Approach to Landslide

Page 10: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

At the water surface (locations #1, 2, and 3),

velocity on the left (#2) is much lower and falls to

zero faster than the other two. This is because of

differences in slide thickness and upslope slide

extent. V-square friction (18) offers resistance in-

versely proportional to slide thickness. As apparent in

the left column of Fig. 6, thin bits near slide edges

move slower than thicker bits near the middle.

Upslope extent also affects slide speed and duration.

The more material upslope, the longer it takes to pass

by a given place. The higher on the slope it originated

the higher velocity it acquires in transit. After the

slide mass has completely passed by at the left,

material continues to pass at the middle and right. As

a result, more deposits pile on the middle and right

sides, which agrees with the observed residual

(Fig. 2).

With regard to peak speeds on the centerline (#1,

4, and 5), the above-water position (#4) ranks highest

and the underwater position (#5) ranks lowest. This is

expected, because greater frictional resistance was

applied underwater. At t = 20 s, velocities at posi-

tions #4 and #5 vanish, but masses still pass by

position #1. The last few moving masses slip into the

water there but they stop just below the surface.

Previous analysis of the witness video (HUANG

et al. 2012) inferred a peak slide speed of 11.65 m/s,

somewhat lower than our value. HUANG et al. (2012)

quantified the velocity of the uppermost part of the

landslide, but we know the landslide deformed as it

moved with different velocity profiles at different

locations (e.g. Fig. 7). Measured velocity at the slide

top does not necessarily reflect the slide velocity at

the water level, where waves are produced.

5. Wave Generation, Propagation, and Inundation

5.1. Wave Sources

Given an initial distribution of still water H(ri,t)

and the landslide model computed above, there are

Figure 7Slide velocities at five positions above, below, and at the water level. The blue, purple, red, orange, and green lines indicate velocities at

positions 1–5 in Fig. 2

Figure 8Landslide tsunami-generating mechanisms. ‘‘Lift-up’’ (NMT) and

‘‘drag-along’’ (DA) are the main sources of wave generation

L. Xiao et al. Pure Appl. Geophys.

Page 11: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

several ways to introduce waves. The classical

approach simply lifts or drops the water by an

amount equal to the passing slide thickness. Gravita-

tional energy is imparted to the water in this way, but

there is no direct transfer of momentum from the

slide to the water. We call this approach ‘‘no

momentum transfer (NMT)’’. NMT may be adequate

for some tsunami sources, for example submarine

earthquakes, but for high speed slides into water it

does not suitably describe wave generation.

Here, in addition to NMT, we introduce ‘‘drag-

along (DA)’’. DA assumes that extra forces exist at

the water–landslide interface. These forces act to

slow the slide but also accelerate the water, much like

anti-friction. For a submarine landside moving at

velocity vs(ri,t), drag-along acceleration of an over-

lying water layer of thickness H(ri,t) would be:

adaðri; tÞ ¼ cdavsðri; tÞ vsðri; tÞj j=Hðri; tÞ ð18Þ

The drag-along coefficient, cda, may or may not

equal the dynamic friction coefficient ld. Unlike

friction (Eq. 17) that acts only in the direction

opposite to fluid flow, DA (Eq. 18) can accelerate

the flow in any direction that the slide is moving. DA

transfers momentum from the slide to the water and

enhances wave production beyond that of NMT

alone.

5.2. Observational Constraints on the Tsunami

Field workers record tsunami heights at the shore

in two ways: measurement of the wave trail on trees,

docks, and structures, or measurement of the trace of

the dry land–wet land transition. The former records

wave height as it reaches those objects (this is

denoted ‘‘flow depth’’). The latter reveals the highest

elevation the wave reaches on shore (run-up height).

The two types of measurement can lead to sig-

nificantly different results, especially where tsunamis

inundate complex terrain, for example that around the

Gongjiafang site.

A field investigation of wave run-up was con-

ducted by two groups soon after the landslide (DAI

et al. 2010; HUANG et al. 2012). They did not reveal

the methods used for measurement nor did they

indicate precision or local variability. We note the ten

surveyed values from the two groups in Figs. 11 and

12. Wave run-up decayed both upstream and down-

stream from the landslide, and the further from the

landslide, the slower was the rate of decrease. Waves

ran up highest on the north shore where the landslide

occurred. Three-hundred and 4,000 m upstream on

the north shore the impulse wave ran up 13.1 and

1.1 m, respectively.

5.3. Tsunami Simulation Results

5.3.1 Model Setup

Using the same 10-m spaced digital topography as for

the landslide simulation, the Yangtze reservoir was

filled to 173 m elevation extending 6 km up and

downstream and the simulated Gongjiafang landslide

was allowed to fall into the River. For water flows,

both dynamic and basal friction, ld and lb, are equal

to 0, except near the shore where, ld increases to

0.02, because of the higher friction there. For this

study we set the drag-along coefficient, cda, at 0.2.

5.3.2 Model Results

Within seconds, the landslide ‘‘pushes up’’ and

‘‘drags along’’ the water to form a tsunami that

spreads out over the river (Figs. 9, 10). The wave

reaches the opposite bank in approximately 15 s, runs

up on the land, then flows back as a reflected wave.

The return wave peaks over the landslide bank at

t = 61 s. The orange curve in each part of Fig. 9

shows the maximum flow depth up to that time along

the river cross-section.

Figure 10 shows a plot of the tsunami waveforms

at three positions (#1, #2, and #3 in Fig. 9). In

general, the first wave crest is higher than the later

ones. The 17-s interval between the first crests of the

blue and green curves corresponds to the cross-river

propagation time from point #1 to point #3. For long

waves, the cross-river transit time Dt can be ex-

pressed as:

Dt ¼Z#3

#1

ds=ffiffiffiffiffiffiffiffiffiffiffi

ghðsÞp

ð19Þ

where h(s) is the water depth, g is the acceleration of

gravity, and s is the along-path position. For known

Tsunami Squares Approach to Landslide

Page 12: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

Figure 9Waves propagate and reflect in river cross section at the 4th, 10th, 20th, and 61st seconds. The orange lines depict the highest flow height and

have been exaggerated tenfold for better visualization. The yellow spots 1 (right), 2 (middle), and 3 (left) on the water surface mark positions

where flow height is tracked in Fig. 10. A Quicktime animation is available at http://es.ucsc.edu/*ward/Yangtze2.mov

L. Xiao et al. Pure Appl. Geophys.

Page 13: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

water depth, Eq. (19) gives a cross-river transit time

of 17 s, which perfectly matches the modeling result.

The dominant wave period at sites #1 and #3 is

32 s. Wave crests at #1 correlate with troughs at #3,

and vice versa. The dominant period at mid-river site

#2 is 16 s, much less than at the river edge sites.

These features fit the character of standing wave

modes. The period of the nth river mode is:

Tn ¼ 2Dt=n ð20Þ

with Dt calculated by use of Eq. (19). n/2 corresponds

to the number of wavelengths that fit, bank to bank.

Figure 10Wave trains at three positions on a cross-river line. The locations are shown in Fig. 9

Figure 11Extent of the river tsunami after T = 7, 20, and 50 s, and 2:33 and 4:01 min. Numbers in circles show run-up heights in decimeters. Numbers

beside red spots are field data, also in decimeters. A Quicktime animation is available at http://es.ucsc.edu/*ward/Yangtze3.mov

Tsunami Squares Approach to Landslide

Page 14: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

The fundamental (‘‘slosh’’) mode (n = 1) has a pe-

riod of T = 32 s with peak amplitudes of opposite

sign at the river banks, as is apparent from the blue

and green curves in Fig. 9. The n = 1 mode has a

node at the river center, so this oscillation is not seen

in the red curve. The n = 2 overtone has a peak in

mid-river and a period of T = 16 s. It is the largest

contributor to the red curve.

Landslide-generated tsunami transit rapidly and

negatively affects Yangtze shores for many kilome-

ters in our model, as in the real event. Figure 11

shows a map of wave propagation at five typical

moments. Within 7 s, slide masses hit the river and

water waves radiate outward in an arc (Fig. 11a).

They cross the river and wash the far bank in

approximately 20 s (Fig. 11b). Within 50 s the waves

have propagated a distance of 2 km (Fig. 11c). In

2.5 min the first wave reaches Wushan County

(Fig. 11d). For several minutes afterwards, the signal

echoes between shores before dying out (Fig. 11e).

Collection of simulated run-up heights along

shore line enables preparation of a wave-decay curve

(Fig. 12). This includes broad variations with dis-

tance especially within the first 4 km up and down

river. These variations probably result from interfer-

ence of many positive and negative waves reflected

from the curved river shores (Fig. 11a, for example).

Some sites are naturally prone to higher or lower run-

up because of local geography. Likewise, waves run-

up higher in such narrowing water channels as valley

and branch stream outlets than at straight shore

locations.

The highest wave run-up in the simulation,

17.2 m, was located near the landslide. Maximum

run-up height on the opposite bank was 14.2 m. The

wave height had dropped to 0.4 m by the time it

reached the docks at Wushan County. Field observa-

tions measured 13 m at the landslide and 12 m at the

opposite bank. For the south shore, the observed data

(red dots) are scattered around the simulation result

(red line). The observed north shore values (blue

spots) are slightly higher than the values calculated.

Considering that the specific run-up quantities mea-

sured in the field are not completely clear nor were

any uncertainties assigned, we cannot make a formal

statement of goodness of fit, but the run-up heights on

both the north and south shores are reasonable fits

with the observations, with correlation coefficient

R2 = 0.88.

6. Conclusions

This paper introduces Tsunami Squares, a new

approach for modeling of landslide-generated waves.

Tsunami Squares has the advantages of the previous

Tsunami Ball method, for example, special separate

treatment for dry and wet cells is not needed, but it

Figure 12Run-up height decay as a function of distance along the north (blue) and south (red) shores. The landslide occurred on the north shore. The

scattered dots are the observed values on the north (blue) and south (red) shores

L. Xiao et al. Pure Appl. Geophys.

Page 15: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

obviates the use of millions of individual particles.

Simulations can be expanded to spatial scales not

possible previously. The new method accelerates and

transports squares of material that are fractured into

new squares in such a way as to conserve volume and

linear momentum.

The simulation first computes landslide motions

on dry land. A novel aspect of Tsunami Squares is

that it considers landslides as part solid and part fluid.

Landslides are more fluid-like within a few seconds

of initiation after the material loses cohesion. Solid

landslides are driven by the slope of the bottom

surface. Fluid-like landslides are driven by the slope

of the top surface.

The second step in the simulation introduces

water. The falling landslide generates waves by two

means. One is the simple uplift of the fluid, known as

‘‘no momentum transfer’’. The second mechanism is

drag-along, a force representing the interaction be-

tween the slide and water. Velocity-dependent drag-

along contributes substantially to tsunami wave

generation from high-speed landslides.

We have demonstrated and validated Tsunami

Squares by modeling the 2008 Three Gorges Reser-

voir Gongjiafang landslide and river tsunami. The

landslide’s progressive failure, generated wave, and

subsequent propagation and run-up are well repro-

duced (WARD and XIAO 2013).

On a laptop computer, Tsunami Square simula-

tions flexibly handle a wide variety of waves and

flows, and are excellent techniques for risk estimation,

hazard assessment, and emergency management

applications.

Acknowledgments

L. Xiao was supported by the National Natural

Science Foundation of China (no. 41202247). We

thank Professor K. Yin, from the China University of

Geosciences (Wuhan), for supporting the geology

background research and field investigation. We

thank Dr Simon Day, from University College

London, for careful revision. We also thank Chongq-

ing Three Gorges Reservoir Geological Hazards

Prevention and Control Office for supplying the

10-meter digital elevation map.

Appendix 1: Inclusion of wave dispersion

Long wave theory assumes that the depth-aver-

aged horizontal acceleration of a water column is

proportional to the gradient of the fluid surface as

stated in Eq. (21):

ovðr; tÞot

¼ �vðr; tÞ � rhvðr; tÞ � grhfðr; tÞ ð21Þ

According to linear dispersive wave theory

(WARD AND DAY 2010), the depth-averaged horizontal

acceleration of a water column is proportional to the

gradient of the fluid surface smoothed over a di-

mension comparable with the water depth. Linear

dispersion can be accommodated in tsunami squares

simply by replacing Eq. (21) by Eq. (22)

ovðr; tÞot

¼ �vðr; tÞ � rhvðr; tÞ � grhfsmoothðr; tÞ

ð22Þ

where

fsmoothðr; tÞ ¼ fðr; tÞ � SðrÞ

¼Z

fðr0; tÞSðr� r0Þdr0ð23Þ

and

SðrÞ ¼ Re

Z

k

eik�r tanhðkhÞ4p2 kh

dk ð24Þ

This indicates that short wave (kh � 1) contri-

butions to the surface gradient impart less depth-

averaged acceleration to the water column than do

longer wave (kh � 1) contributions. As a result,

short waves fall behind long waves, as dictated by

linear dispersive theory. For the applications in this

paper, water wave dispersion is not important.

REFERENCES

BASU D, DAS K, GREEN S, JANETZKE R, STAMATAKOS J. (2010),

Numerical Simulation of Surface Waves Generated by a

Subaerial Landslide at Lituya Bay Alaska. Journal of Offshore

Mechanics and Arctic Engineering, 132, p 41101.

COUSSOT P, MEUNIER M. (1996), Recognition, classification and

mechanical description of debris flows. Earth-Sci Rev, 40(1996),

209–227.

Tsunami Squares Approach to Landslide

Page 16: Tsunami Squares Approach to Landslide-Generated Waves ...ward/papers/Tsunami Squares-final.pdf · 3. Tsunami Squares Theory The two-dimensional Tsunami Squares approach evolved from

DAI Y, WANG Y, YIN K, CHEN L, LIU B. (2010), Surge Survey and

Calculation Analysis of a Landslide in Wushan County in the

Three Gorges Reservoir. Journal of Wuhan University of Tech-

nology, 32(19), 14–71 (In Chinese).

FRITZ HM, HAGER WH, MINOR HE. (2003), Landslide generated im-

pulse waves. 1. Instantaneous flow fields. Exp Fluids, 35, 505–519.

FRITZ HM, MOHAMMED F, YOO J. (2009), Lituya Bay Landslide

Impact Generated Mega-Tsunami 50th Anniversary. Pure Appl

Geophys (166), 153–175.

HARBITZ CB, GLIMSDAL S, LOVHOLT F, KVELDSVIK V, PEDERSEN GK,

A JENSEN. (2014), Rockslide tsunamis in complex fjords: from an

unstable rock slope at Akerneset to tsunami risk in western

Norway. Coast Eng, 88(2014), 101–122.

HARBITZ CB, VHOLT FL, BUNGUM H. (2014), Submarine landslide

tsunamis: how extreme and how likely? Nat Hazards, 72(3),

1341–1374. doi:10.1007/s11069-013-0681-3.

HEINRICH P. (1992), Nonlinear water waves generated by sub-

marine and aerial landslides. Journal of Waterway, Port, Coastal

and Ocean Engineering, 118, 249.

HUANG B, YIN Y, LIU GN, WANG SC. (2012), Analysis of waves

generated by Gongjiafang landslide in Wu Gorge, Three Gorges

Reservoir, on November 23, 2008. Landslides, 9(3), 395–405.

HUANG B, YIN Y, WANG S, CHEN X, LIU G, JIANG Z, LIU J. (2013), A

physical similarity model of an impulsive wave generated by

Gongjiafang landslide in Three Gorges Reservoir, China.

Landslides, 1–13.

HUNGR O. (1995), A model for the runout analysis of rapid flow

slides, debris flows, and avalanches. Can. Geotech. J., 32(1995),

610–623.

HUNGR O, MCDOUGALL S. (2009), Two numerical models for land-

slide dynamic analysis. Computer and Geosciences, 35(2009),

978–992.

LIU PL, LYNETT P, SYNOLAKIS CE. (2003), Analytical solutions for

forced long waves on a sloping beach. J Fluid Mech, 478,

101–109.

MADER CL, GITTINGS ML. (2002), Modeling the 1958 Lituya Bay

mega-tsunami. Science of Tsunami Hazards, 20(5), 241–250.

POISSON B, PEDREROS R. (2010), Numerical modeling of historical

landslide-generated tsunamis in the French Lesser Antilles.

Natural Hazards and Earth System Sciences (10), 1281–1292.

QUECEDO M, PASTOR M, HERREROS MI. (2004), Numerical modelling

of impulse wave generated by fast landslides. Int J Numer Meth

Eng, 59, 1633–1656.

SATAKE K, TANIOKA Y. (1995), Tsunami generation of the 1993

Hokkaido Nansei-Oki earthquake. Pure Appl Geophys, 144(3–4),

803–821.

TITOV VV, GONZALEZ FI. (1997), Implementation and testing of the

method of splitting tsunami (MOST) model: US Department of

Commerce, National Oceanic and Atmospheric Administration,

Environmental Research Laboratories, Pacific Marine Environ-

mental Laboratory.

WALDER JS, WATTS P, SORENSEN OE, JANSSEN K. (2003), Tsunami

generated by subaerial mass flows. Journal of Geophysical Re-

search, 108(B5), 2236–2254.

WARD SN. (2014), Lituya Bay Tsunami. https://www.youtube.com/

watch?v=6COeNRToYqU.

WARD SN, DAY S. (2010), The 1958 Lituya bay landslide and

tsunami–a tsunami ball approach. Journal of Earthquake and

Tsunami, 4(4), 285–319.

WARD SN, DAY S. (2005), Tsunami thoughts. CSEG RECORDER.

WARD SN, DAY S. (2006), Particulate kinematic simulations of

debris avalanches: interpretation of deposits and landslide

seismic signals of Mount Saint Helens, 1980 May 18. Geophys.

J. Int, 167, 991–1004.

WARD SN, DAY S. (2008), Tsunami Balls: A Granular appproach to

Tsunami runup and inundation. Communications in computa-

tional Physics, 3(1), 222–249.

WARD SN, XIAO L. (2013), Yangtze Tsunami. YOUTUBE MOVIE

https://www.youtube.com/watch?v=JBa8z9oPgLI.

WATTS P, TAPPIN DR. (2012), Geowave Validation with Case

Studies: Accurate Geology Reproduces Observations. In Y

YAMADA, K KAWAMURA, K IKEHARA et al. (Eds.), Submarine Mass

Movements and Their Consequences (31, pp. 517): Springer

Netherlands.

WATTS P, GRILLI ST, KIRBY JT, FRYER GJ, TAPPIN DR. (2003),

Landslide tsunami case studies using a Boussinesq model and a

fully nonlinear tsunami generation model. Natural Hazards and

Earth System Sciences, 3, 391–402.

WEISS R, FRITZ HM, WUNNEMANN K. (2009), Hybrid modeling of the

mega-tsunami runup in Lituya Bay after half a century. Geophys

Res Lett, 36(9), L9609.

WU B, LI H, YAO M. (2010), Deformation and Failure Mechanism

of Slope in Area from Gongjiafang to Dulong of Wuxia County,

Chongqing. Chinese Journal of Underground Space and Engi-

neering, 6(2010), 1656–1659 (in Chinese).

YIN K, LIU Y, WANG Y, JIANG Z. (2012), Physical Model Ex-

periments of landslide-induceds surge in Three Gorges reservoir.

Earth Science–Journal of China University of Geosicences,

37(5), 1067–1074 (in Chinese).

(Received August 8, 2014, revised January 15, 2015, accepted January 16, 2015)

L. Xiao et al. Pure Appl. Geophys.