tsunami hazard assessment of canada

38
ORIGINAL PAPER Tsunami hazard assessment of Canada Lucinda J. Leonard Garry C. Rogers Ste ´phane Mazzotti Received: 2 June 2013 / Accepted: 18 July 2013 / Published online: 24 August 2013 Ó Her Majesty the Queen in Right of Canada 2013 Abstract We present a preliminary probabilistic tsunami hazard assessment of Canadian coastlines from local and far-field, earthquake, and large submarine landslide sources. Analyses involve published historical, palaeotsunami and palaeoseismic data, modelling, and empirical relations between fault area, earthquake magnitude, and tsunami run-up. The cumulative estimated tsunami hazard for potentially damaging run-up (C1.5 m) of the outer Pacific coastline is *40–80 % in 50 years, respectively one and two orders of magnitude greater than the outer Atlantic (*1–15 %) and the Arctic ( \ 1 %). For larger run-up with significant damage potential (C3 m), Pacific hazard is *10–30 % in 50 years, again much larger than both the Atlantic (*1–5 %) and Arctic ( \ 1 %). For outer Pacific coastlines, the C1.5 m run-up hazard is dominated by far-field subduction zones, but the probability of run-up C3 m is highest for local megathrust sources, particularly the Cas- cadia subduction zone; thrust sources further north are also significant, as illustrated by the 2012 Haida Gwaii event. For Juan de Fuca and Georgia Straits, the Cascadia megathrust dominates the hazard at both levels. Tsunami hazard on the Atlantic coastline is dominated by poorly constrained far-field subduction sources; a lesser hazard is posed by near-field continental slope failures similar to the 1929 Grand Banks event. Tsunami hazard on the Arctic coastline is poorly constrained, but is likely dominated by continental slope failures; Electronic supplementary material The online version of this article (doi:10.1007/s11069-013-0809-5) contains supplementary material, which is available to authorized users. L. J. Leonard (&) G. C. Rogers Geological Survey of Canada, Natural Resources Canada, 9860 West Saanich Road, P.O. Box 6000, Sidney, BC V8L 4B2, Canada e-mail: [email protected] Present Address: L. J. Leonard School of Earth and Ocean Sciences, University of Victoria, PO Box 1700 Station CSC, Victoria, BC V8W 2Y2, Canada S. Mazzotti Geosciences Montpellier, UMR 5243, Universite ´ Montpellier 2, Montpellier, France 123 Nat Hazards (2014) 70:237–274 DOI 10.1007/s11069-013-0809-5

Upload: uvic

Post on 28-Jan-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

ORI GIN AL PA PER

Tsunami hazard assessment of Canada

Lucinda J. Leonard • Garry C. Rogers • Stephane Mazzotti

Received: 2 June 2013 / Accepted: 18 July 2013 / Published online: 24 August 2013� Her Majesty the Queen in Right of Canada 2013

Abstract We present a preliminary probabilistic tsunami hazard assessment of Canadian

coastlines from local and far-field, earthquake, and large submarine landslide sources.

Analyses involve published historical, palaeotsunami and palaeoseismic data, modelling,

and empirical relations between fault area, earthquake magnitude, and tsunami run-up. The

cumulative estimated tsunami hazard for potentially damaging run-up (C1.5 m) of the

outer Pacific coastline is *40–80 % in 50 years, respectively one and two orders of

magnitude greater than the outer Atlantic (*1–15 %) and the Arctic (\1 %). For larger

run-up with significant damage potential (C3 m), Pacific hazard is *10–30 % in 50 years,

again much larger than both the Atlantic (*1–5 %) and Arctic (\1 %). For outer Pacific

coastlines, the C1.5 m run-up hazard is dominated by far-field subduction zones, but the

probability of run-up C3 m is highest for local megathrust sources, particularly the Cas-

cadia subduction zone; thrust sources further north are also significant, as illustrated by the

2012 Haida Gwaii event. For Juan de Fuca and Georgia Straits, the Cascadia megathrust

dominates the hazard at both levels. Tsunami hazard on the Atlantic coastline is dominated

by poorly constrained far-field subduction sources; a lesser hazard is posed by near-field

continental slope failures similar to the 1929 Grand Banks event. Tsunami hazard on the

Arctic coastline is poorly constrained, but is likely dominated by continental slope failures;

Electronic supplementary material The online version of this article (doi:10.1007/s11069-013-0809-5)contains supplementary material, which is available to authorized users.

L. J. Leonard (&) � G. C. RogersGeological Survey of Canada, Natural Resources Canada, 9860 West Saanich Road, P.O. Box 6000,Sidney, BC V8L 4B2, Canadae-mail: [email protected]

Present Address:L. J. LeonardSchool of Earth and Ocean Sciences, University of Victoria, PO Box 1700 Station CSC, Victoria, BCV8W 2Y2, Canada

S. MazzottiGeosciences Montpellier, UMR 5243, Universite Montpellier 2, Montpellier, France

123

Nat Hazards (2014) 70:237–274DOI 10.1007/s11069-013-0809-5

a hypothetical earthquake source beneath the Mackenzie delta requires further study. We

highlight areas susceptible to locally damaging landslide-generated tsunamis, but do not

quantify the hazard.

Keywords Tsunami � Probabilistic hazard analysis � Canada � Earthquake �Landslide

1 Introduction

The occurrence of several major tsunamis globally within the past decade (notably Sumatra

in 2004, Chile in 2010, and NE Japan in 2011), accompanied by catastrophic loss of life,

highlights the need for coastal nations to assess and mitigate their tsunami hazard. This is

particularly true for Canada, which has a longer coastline than any other country and is at

risk from tsunamis generated in three oceans (Fig. 1). Previous studies concern various

historical tsunamis, and the threat that future tsunamis may pose (references in Leonard

et al. 2010). This analysis presents a first attempt to quantify the tsunami hazard on each

coast, from both near- and far-field sources. Earthquakes represent the greatest and most

easily quantified tsunami source; we also consider hazard contributions from mass

movement sources. We do not assess tsunami hazard along inland waterways, although

several historical occurrences involving landslide sources have caused damage and fatal-

ities (e.g. Evans 2001). We exclude the likely common regional hazard of meteorological

tsunamis (e.g. Stephenson and Rabinovich 2009), which result from atmospheric rather

than geologic phenomena and are likely ubiquitous rather than having a distinct source

area. We also omit the potentially devastating but very rare hazard posed by asteroid

impact tsunamis (e.g. Gisler et al. 2011).

This assessment, based on currently available constraints, is intended for use as a starting

point to establish priorities for future work. Identification of poorly constrained sources with

potentially high hazard highlights the need to better quantify both source potential and

tsunami impact. The probabilities determined here will require revision as more data are

collected regarding: the occurrence, frequency, and size of past tsunamis; constraints on

potentially tsunamigenic faults; stability assessments of potential landslide sources; tsunami

modelling of a wide range of potential sources from tsunamigenesis to propagation and

inundation, incorporating fine-resolution bathymetry and coastal topography.

2 Tsunami hazard analysis: methods

Probabilistic tsunami hazard analysis (e.g. Geist and Parsons 2006) involves estimation of

the probability of exceeding specific tsunami wave heights (or run-up) at given locations

from any source, on the basis of either numerical tsunami simulations or empirical analysis

of past tsunami heights (Geist et al. 2009).

2.1 Recurrence and probabilities

If the annual rate, k, of tsunami run-up height Hr0 is known, the probability of exceeding

that run-up at least once in a time period T is given by:

PðHr [ Hr0; TÞ ¼ 1� e�kT ð1Þ

238 Nat Hazards (2014) 70:237–274

123

assuming a Poisson process (Geist and Parsons 2009). For multiple sources with inde-

pendent rates (k1, k2, k3 …) of Hr0, the combined probability of exceedance in time T is

given by:

PðHr [ Hr0; TÞ ¼ 1� e�kcT ð2Þ

Fig. 1 Location maps with potential sources of tsunamis hazardous to the coasts of Canada. Red linesmegathrust plate boundaries. BB Boundary Bay, BI Banks Island, FDF Fraser delta front; G. Antilles GreaterAntilles, HS Haro Strait, MD Mackenzie delta, NF Northern Hispaniola fault, PRT Puerto Rico trench,Scotian sl. Scotian slope, VI Victoria Island. a Local tsunami sources. Yellow stars epicentres of the 1929and 1946 earthquakes that triggered landslide tsunamis, and the 2012 Haida Gwaii tsunamigenic earthquake.b Far-field tsunami sources in the Pacific (left) and Atlantic Ocean (right). Yellow stars epicentres ofearthquakes that spawned major historical far-field tsunamis impacting Canada

Nat Hazards (2014) 70:237–274 239

123

where kc is the cumulative annual rate. For each tsunami source, we estimate annual rates

of run-up of 1.5 and 3 m in each hazard zone (Fig. 2) and probabilities of exceeding those

values in a 50-year period (Eq. 1). Minimum and maximum values do not represent

specific probabilities (e.g. standard deviations) but the combined effect of uncertainties on

the source parameterisation (e.g. tsunami propagation, earthquake magnitude). For mul-

tiple sources, we estimate the cumulative best annual rate (kc_best) as the sum of the best

values of each source:

kc best ¼XN

i¼1

ki best ð3Þ

Uncertainty (upper and lower bounds) of the combined rates is estimated using a standard

geometric sum of the uncertainty on each source, taking into account that all sources are

independent:

kc min ¼ kc best �

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiXN

i¼1

ðki best � ki minÞ2vuut ð4Þ

Fig. 2 Tsunami hazard zones considered for the assessment. Colour is used to distinguish between labelledzones and does not indicate relative hazard. ATL Atlantic, CMC central mainland coast, EHG eastern HaidaGwaii, GI Gulf Islands, GS Georgia Strait, HB Hudson Bay, HS Hudson Strait, INN inner coast, JDF Juan deFuca Strait, NMC northern mainland coast, OUT outer coast, SE-NFLD southeastern Newfoundland, WHGwestern Haida Gwaii, WVI western Vancouver Island. See Online Resource 1 for explanation of zoneboundaries and potential hazard variation within zones

240 Nat Hazards (2014) 70:237–274

123

kc max ¼ kc best þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiXN

i¼1

ðki best � ki maxÞ2vuut ð5Þ

2.2 Tsunami wave height threshold for potential damage

Wave height or amplitude is defined as the zero-to-peak wave amplitude (height above the

state of tide). Mean run-up height (Hr), which determines inundation extent, is approxi-

mately double the wave height or equivalent to the peak-to-trough amplitude (Abe 1995).

Based on historical Pacific tsunamis, damage to boats, docks, and swimmers may occur

due to strong currents for a 0.5-m minimum tsunami amplitude or run-up; more severe

damage and inundation is likely at a 1.5–2-m minimum (Whitmore et al. 2008). We

therefore choose a tsunami run-up (Hr) threshold of 1.5 m (equivalent to 1.5 m peak-to-

trough amplitude or *0.75 m wave height) for potential coastal damage. We also assess a

run-up threshold of 3 m, defined here as having significant damage potential, i.e. major

damage expected that may be geographically extensive.

2.3 Assessment of subduction earthquake sources

Modelling of the many tsunami sources with the potential to impact Canadian coasts is

beyond the scope of this study. Various scenario fault sources have been modelled to assess

their tsunami impact, often with historical constraints; we use the results, along with

palaeoseismic or geodetic recurrence intervals, for preliminary probabilistic analysis. For

Pacific far-field sources, we use a century-long tide gauge dataset for an empirical analysis.

For unconstrained potentially tsunamigenic faults, we use a series of empirical relations to:

(1) derive an estimate of earthquake magnitude and probability based on fault area and

convergence rate, (2) derive the corresponding tsunami run-up in each zone, and (3)

integrate the source event probability with the estimated run-up to provide probabilities of

potentially damaging (C1.5 m) and significant (C3 m) tsunami run-up in each zone (see

Online Resource 2 for complete description and equations).

2.4 Assessment of landslide sources

In contrast to large earthquake tsunamis, significant far-field attenuation occurs during

propagation of landslide tsunamis, so generally only local/regional landslide sources need

to be considered (e.g. Geist and Parsons 2009). However, modelling suggests that very

large failures (e.g. volcanic flank collapses) may result in damaging run-up at transoceanic

distances. Like other natural systems, landslides are generally considered to follow an

inverse power-law frequency–size relationship (e.g. Korup and Clague 2009). With suf-

ficient identification and dating of past slides, the frequency of large events may be

estimated and can be used, along with tsunami modelling, to approximate the probability of

potentially damaging run-up. However, for smaller failures, empirical data are difficult to

collect. Failures are often triggered by earthquakes, and the earthquake statistics of an area

may be used to aid landslide tsunami hazard analysis where slope stability data are

available (e.g. ten Brink et al. 2009); such analyses are beyond the scope of this study.

Coastal landslide tsunamis are typically triggered by failure of steep subaerial slopes or

submarine delta fronts, resulting from ground shaking, rainfall, construction, or other

factors. Isolated cases are noted in Canada (e.g. Leonard et al. 2012a), but without a long-

Nat Hazards (2014) 70:237–274 241

123

term history, the probability of landslide-generated waves cannot be estimated. Instead, we

identify coastlines that may be at risk, based on the landslide susceptibility map of Bo-

browsky and Dominguez (2012), who assessed various factors including topography,

geology, vegetation, and precipitation to divide the Canadian landscape into susceptibility

classes ranging from 1 (lowest) to 6 (highest). We consider coastal areas in classes 5 and 6

to be at risk from landslide-generated waves, as well as neighbouring coastlines that may

also be affected. We also highlight Arctic coastlines that may be susceptible to local waves

from iceberg calving or jokulhlaup events (catastrophic releases of water from a glacier).

3 Pacific coast

In historical time (written) on the British Columbia coast (*150 years of local newspa-

pers), most recorded tsunamis have been far-field events (e.g. Stephenson et al. 2007). The

most damaging were triggered by giant (M [ 9) megathrust earthquakes in Alaska (1964)

and Chile (1960; Fig. 1); the largest since 1964 were the Mw 8.8 2010 Chile and Mw 9.0

2011 Japan events. The only known tsunami fatality in BC resulted from a local landslide

tsunami during a Ms7.3 crustal earthquake on Vancouver Island (1946; Fig. 1). However, a

massive local earthquake and tsunami are known to have occurred in 1700 CE (e.g.

Atwater et al. 2005). On 27 October 2012, a tsunami triggered by an Mw 7.7 thrust

earthquake off Haida Gwaii (formerly the Queen Charlotte Islands) impacted western

Haida Gwaii, with run-up over 8 m above the state of tide documented at several locations.

3.1 Far-field tsunamis from Pacific subduction zones

The most complete tsunami record is provided by the tide gauge at Tofino (western

Vancouver Island; Fig. 1), with 61 tsunamis identified from 1906 to 2009 (Wigen 1983;

Stephenson and Rabinovich 2009; WCATWC 2009a, b). Since a 1998 instrument upgrade,

the record contains a few small local/regional earthquake and meteorological tsunamis

(Stephenson and Rabinovich 2009), but it is dominated by far-field events and can be used

to assess their frequency. The Tofino tsunami frequency–amplitude plot (Fig. 3) is a

power-law distribution from which the recurrence of large events can be predicted. For

most of the record, tsunamis down to 6 cm (peak-to-trough) are consistently identified; this

lowers to 3 cm after 1998. From 1975 to 1980 and possibly the early 1980s, the tide gauge

was compromised due to sediment build-up (Stephenson and Rabinovich 2009). We

assume that no tsunamis could be detected from 1975 to 1980, but 1980–1985 was a

quiescent time with no events recorded at tsunami-sensitive Crescent City, CA (Dengler

and Magoon 2006). The shape of the curve at high amplitudes (important for the frequency

of large events) depends on the assumed maximum amplitude. The 1964 Alaska tsunami

(2.4 m peak to trough at Tofino) represents a near maximum, although previous Alaskan

events may have involved longer ruptures (Shennan et al. 2009). We assign a maximum

height of 3.0 ± 0.5 m for far-field tsunamis at Tofino (Fig. 3).

The frequencies of tsunami run-up exceeding 1.5 and 3 m at Tofino are given in

Table 1; uncertainties are propagated from those of the recurrence parameters and maxi-

mum amplitude (Fig. 3). The results only strictly apply to the Tofino tide gauge site; other

sites exhibit different responses for the same event, e.g. Port Alberni usually experiences

larger amplitudes ([5 m in 1964) due to resonance amplification in the Alberni Inlet (Fine

et al. 2009). Response ratios between specific sites tend to be similar for far-field events;

maximum amplitudes at Victoria are typically *0.6 of those at Tofino (Leonard et al.

242 Nat Hazards (2014) 70:237–274

123

2012a), so the Tofino data can be scaled to approximate a recurrence relation for Victoria

(Fig. 3). We assume that the rates estimated for Tofino are applicable to the whole outer

BC coast and Haida Gwaii (Fig. 1), and the lower values at Victoria apply to the rest of

Juan de Fuca Strait and to Queen Charlotte Strait, where recent events recorded at Port

Hardy show similar amplitudes (Stephenson and Rabinovich 2009). We assume that

Georgia and Johnstone Straits and Discovery Passage (zone GS, Fig. 2) are sheltered from

far-field tsunamis, with no damaging waves reported in 1964.

3.2 Cascadia subduction zone

The most recent great earthquake on the Cascadia subduction zone (Fig. 1) preceded the

regional written history, but numerous lines of evidence including First Nations oral

records, Japanese written tsunami records, submerged trees, buried soils and sand layers in

coastal marshes, as well as turbidite deposits offshore, indicate the occurrence of a

Mw * 9 earthquake accompanied by a large tsunami on 26 January 1700 CE (e.g.

Goldfinger et al. 2003; Satake et al. 2003; Leonard et al. 2004; Atwater et al. 2005; Ludwin

et al. 2005). Similar events are inferred throughout the Holocene; recurrence intervals

average *500 years (mean 530 ± 260 years), ranging from *200 to *800 years or more

(Atwater et al. 2004; Goldfinger et al. 2012; Peterson et al. 2013b). Palaeotsunami deposits

associated with subsidence events at the southern entrance to Juan de Fuca Strait imply

three or four megathrust events in the last 1,300 years, all with run-up exceeding 5 m

(Peterson et al. 2013a). Additional megathrust events (Mw C 8) are inferred on the

southern part of the subduction zone; Goldfinger et al. (2012) report turbidite evidence for

22 more Holocene events in the south. Treating southern Cascadia as an independent

source implies an average recurrence interval of *450 years, but the record remains under

Fig. 3 Cumulative frequency distribution of far-field tsunami peak-to-trough wave amplitudes at Tofino(white symbols, solid line, dark grey shading) and Victoria, BC (dashed line, light grey shading). Tofinorecurrence curve based on tide gauge data (error bars assume Poissonian statistics) and tapered to assignedmaximum amplitude; Victoria curve based on a scaling relationship (see text). Empirical data suggest peak-to-trough amplitudes are approximately equivalent to local mean run-up Hr (Abe 1995). Thick grey linesapproximation of frequency of tsunami run-up C1.5 m at Tofino and Victoria

Nat Hazards (2014) 70:237–274 243

123

Table 1 Estimation of potentially damaging run-up on the Canadian coast from tsunamis generated byvarious sources

Sourcea Event frequencyb (/year) Regions affectedc run-up classd

min, best, max min, best, max

FF SZ 1.5 0.0050, 0.0114, 0.0224 WHG, EHG, NMC,CMC, WVI

2

0, 0.0016, 0.0055 JDF, QCS 2

FF SZ 3 0, 0, 0.0030 WHG, EHG, NMC,CMC, WVI

3

CSZ long 0.0013, 0.0019, 0.0037 WHG, EHG, NMC,CMC, QCS, WVI,JDF

3

GSe 2

CSZ south 0.0008, 0.0010, 0.0013 WHG, EHG, NMC,CMC, QCS, WVI-N, JDF

1

WVI-S 2

GSe 0–1

Explorer 0.0010, 0.0030, 0.0152 WHG-N 0, 0, 2

WHG-S, EHG, NMC,CMC, QCS, WVI-S

0, 1, 2

WVI-N 2, 3, 3

Haida Gwaii 0.0005, 0.0013, 0.0038 WHG 2, 3, 3

EHG, NMC, CMC,QCS, WVI-N

0, 1, 2

WVI-S 0, 0, 1

Crustal faults 0.0001, 0.0004, 0.0011 JDF-E 2

0.00005, 0.0002, 0.00055 JDF-W, GS 2

0.0001, 0.0005, 0.0014 EHG, NMC, CMC 1, 2, 2

0, 0, 0.0003 ATL-INN 0, 0, 2

Hawaiian Is. 4 9 10-6, 7 9 10-6,11 9 10-6

WHG, EHG, NMC,CMC, QCS, WVI,JDF

1, 2, 3

Aleutians 1.5 0.0002, 0.0003, 0.0004 WHG, EHG, NMC,CMC, QCS, WVI,JDF

2

Aleutians 3 13 9 10-5, 25 9 10-5,29 9 10-5

3

Gib-Cadiz FF 0.0004, 0.0005, 0.0012 ATL-OUT, SE-NFLD 2

ATL-INN, HS 1

PR trench FF 0.0001, 0.0003, 0.0010 ATL-OUT-S, SE-NFLD, ATL-INN

2

ATL-OUT-N, HS 1

L. Antilles FF 0.0006, 0.0009, 0.0019 ATL-OUT-S 2

SE-NFLD, ATL-INN 1

C-slope 1.5 6 9 10-5, 10 9 10-5,15 9 10-5

ATL-OUT, ARCTIC 2

ATL-INN, HS 1

15 9 10-5, 25 9 10-5,38 9 10-5

SE-NFLD 2

244 Nat Hazards (2014) 70:237–274

123

debate, with some turbidites of questionable origin (e.g. Atwater and Griggs 2012); we

follow Frankel (2011) in assuming a recurrence interval of 1,000 years for the southern

events and assume an uncertainty of ±250 years. We omit a possible independent northern

Cascadia source suggested by Atwater and Griggs (2012).

Tsunami modelling of a full M * 9 Cascadia subduction zone rupture (500-year strain

release) shows that south-western Vancouver Island is most at risk, with 5–8 m wave

amplitudes predicted (16 m maximum), and currents up to 17 m/s in narrow channels and

near headlands (Whitmore 1993; Cherniawsky et al. 2007). In Juan de Fuca Strait,

amplitudes are *50 % of the outer coast, as indicated by the models and by recorded far-

field tsunamis; similar amplitudes are predicted in narrow channels and some bays of the

Gulf Islands, although true maxima in bays and harbours are likely higher than predicted

without a fine-resolution grid. Use of a fine-scale grid in the harbours of Victoria shows

amplitudes up to 4.2 m. In the open southern Strait of Georgia, waves are attenuated to

*10–20 % of outer coast values, with the exception of some local resonances, e.g.

amplitudes in Boundary Bay (Fig. 1) are at least double those in the Vancouver region. At

tide gauges in Burrard Inlet, the 1964 tsunami was about 10 % of the Tofino amplitude

(Spaeth and Berkman 1967). The model does not extend into northern Georgia Strait or

Johnstone Strait; without further constraints, we assume that run-up is similar to that in

southern Georgia Strait.

Modelling of an independent southern Cascadia megathrust source (Cherniawsky et al.

2007) shows that most tsunami energy is directed east and west; run-up for a 500-year

strain release event will likely exceed 1.5 m only on south-western Vancouver Island. Up

to 1.1 m amplitudes occur in the harbours of Victoria; local run-up maxima C1.5 m may

be possible. Without fine-scale modelling in Georgia Strait, similar maxima cannot be

ruled out.

Table 1 continued

Sourcea Event frequencyb (/year) Regions affectedc run-up classd

min, best, max min, best, max

C-slope 3 5 9 10-5, 9 9 10-5,14 9 10-5

ATL-OUT, ARCTIC 3

ATL-INN, HS 2

11 9 10-5, 20 9 10-5,31 9 10-5

SE-NFLD 3

Canary Is. 3 9 10-6, 6 9 10-6,10 9 10-6

ATL-OUT, ATL-INN, SE-NFLD, HS

1, 2, 3

a Source of tsunamigenic earthquakes/landslides. Explorer and Haida Gwaii fault sources defined inTable 2. FF SZ: far-field subduction zones; Aleutians, and C-slope 1.5 and 3: Aleutian and continental slopelandslide sources for tsunami run-up exceeding 1.5 and 3 m. CSZ Cascadia subduction zone, Gib Gibraltar,PR Puerto Rico, L. Antilles Lesser Antillesb Frequency of source events, estimated from average historical/palaeoseismic/predicted recurrencec Zones shown in Fig. 2. WHG: both WHG-N and WHG-S; WVI: WVI-N and WVI-S; JDF: JDF-W andJDF-E&GId run-up class: 0: Hr \ 1.5 m; 1: 2Hr C 1.5 m; 2: Hr C 1.5 m, 2Hr C 3 m; 3: Hr C 3 m. Local mean andmax run-up Hr and 2Hr estimated using Abe (1995) empirical relations (see Online Resource 2) and/orhistorical data, palaeoevent data, modelling, or various assumptions (see text for each source)e Tsunami impact poorly defined. For CSZ long, we assume that mean run-up C3 m is likely only for verylarge (max) events: equivalent recurrence 1,150 years (Witter et al. 2011) (0 min and best probability). ForCSZ south, we assume local maxima C1.5 m are improbable but not impossible (probability 25 % ofearthquake probability; minimum 0)

Nat Hazards (2014) 70:237–274 245

123

Estimated tsunami hazard from Cascadia megathrust ruptures (Table 1) is based on a

combination of modelling (Cherniawsky et al. 2007), empirical relations (for areas north of

the model; equations in Online Resource 2), and palaeoseismic recurrence. Long ruptures

could produce run-up exceeding 1.5 m for all zones and generally C3 m for all regions

except Georgia Strait, where run-up exceeding 3 m is assumed to occur only in the largest

events (recurrence *1,150 years; Witter et al. 2011). Southern ruptures could lead to run-

up generally exceeding 1.5 m (local maxima C3 m) only along south-western Vancouver

Island, with all other areas susceptible only to local maxima C1.5 m.

3.3 Explorer–North America plate boundary

North of Cascadia, the margin transitions from dominantly convergent to strike slip

(Fig. 4). With no large historical earthquakes, debate is ongoing regarding convergence

between the Explorer and North America plates (e.g. Rohr and Furlong 1995; Braunmiller

Fig. 4 Tectonics of the Explorer and Haida Gwaii region. Dark and light grey shaded areas coseismiclocked and transition zones assumed for Explorer subduction (Mazzotti et al. 2003b; Table 2). Red box areaof high seismicity on crustal faults—potential tsunami sources. BP Brooks Peninsula, NFZ Nootka faultzone, PS Puget Sound

246 Nat Hazards (2014) 70:237–274

123

and Nabelek 2002), with contrasting implications for tsunamigenic potential. Despite the

significant internal deformation (e.g. Rabinovich et al. 2008), the existence of an inde-

pendent Explorer plate is supported by GPS data (Mazzotti et al. 2003b), offshore sediment

deformation (Davis and Hyndman 1989), and earthquake slip vectors (Braunmiller and

Nabelek 2002). A subducted slab is present beneath Vancouver Island as far north as

Brooks Peninsula (Cassidy et al. 1998); active subduction is implied by the occurrence of

Episodic Tremor and Slip (Kao et al. 2009). Relative motion of *20 mm/year across the

Nootka fault (Willoughby and Hyndman 2005) suggests *20–25 mm/year convergence at

the southern end of the Explorer–North America boundary, consistent with GPS data

(Mazzotti et al. 2003b). Evidence for active subduction of the Winona block (northern

Explorer) is inconclusive; GPS data cannot discriminate between oblique subduction and

permanent crustal deformation (Mazzotti et al. 2003b). The absence of a subducted slab

north of Brooks Peninsula (Cassidy et al. 1998) implies either a lack of subduction or slow,

recent subduction (Braunmiller and Nabelek 2002).

Here, we treat the Explorer subduction as an independent source, but it is possible that

subduction could occur as an extension of a megathrust earthquake on the rest of the

Cascadia system to the south. Assuming subduction of the Explorer plate (slip tapering to

zero through the Winona block) and a seismogenic zone consistent with GPS and thermal

data (Mazzotti et al. 2003b; Fig. 4; Table 2), we calculate that Mw 7.7 ± 0.3 megathrust

events may occur with an average recurrence interval of *333 years (112–1,005 years).

We also assess a smaller rupture scenario (Winona block excluded); for Mw * 7.4 events,

we calculate an average recurrence interval of *200 years (Table 2). Mw 7.7 events could

produce up to 7 m run-up on north-western Vancouver Island, but on the south-western

coast, northern mainland coast, and southern Haida Gwaii, only local maxima C1.5 m are

likely (Table 1). Larger events could produce run-up consistently exceeding 1.5 m over

these areas, and smaller ones would have a far more localised impact.

3.4 Haida Gwaii margin

Pacific–North America relative plate motion at the Haida Gwaii margin (Fig. 4) is pri-

marily strike slip, accommodated on the near-vertical Queen Charlotte fault, but is

increasingly oblique southward. The 1949 M 8.1 mainly strike-slip event produced only a

small tsunami with reported amplitudes mostly *0.5 m; a reported *3 m wave was likely

landslide-related. A 2001 Mw 6.1 thrust event near the southern Queen Charlotte fault

caused a larger than expected tsunami (up to 23 cm on Vancouver Island; Rabinovich et al.

2008), demonstrating the potential tsunami hazard.

Convergence increases from *8 to 15 mm/year from northern to southern Haida Gwaii

(Mazzotti et al. 2003a). Two end-member models are proposed: (1) subduction of the

Pacific plate beneath Haida Gwaii (e.g. Hyndman and Ellis 1981); (2) internal deformation

of both plates (e.g. Rohr et al. 2000). GPS data infer *5 mm/year shortening within the

North America plate, but the 6–10 mm/year residual convergence may be explained by

either underthrusting or shortening of the Pacific plate (Mazzotti et al. 2003a). The sub-

duction model is supported by bathymetry (trench; accretionary prism), receiver function

analysis showing a dipping slab, and by gravity and heat flow data (e.g. Bustin et al. 2007).

We assume subduction on a fault dipping at 28� (Bustin et al. 2007) to calculate an

average recurrence interval of *760 years for megathrust earthquakes of Mw * 7.8

(7.5–8.2; Table 2). Some parameters are constrained over a relatively small area and likely

are not uniform along strike, e.g. dip angle varies with the age of oceanic crust. Also, the

convergence rate used implies a lack of internal deformation of the Pacific plate, e.g.

Nat Hazards (2014) 70:237–274 247

123

Tab

le2

Est

imat

ion

of

eart

hquak

em

agnit

ude

and

recu

rren

ceon

pote

nti

alts

unam

igen

icfa

ult

s

Fau

ltL

eng

th(k

m)

Wid

tha

(km

)M

wb

min

,b

est,

max

Mob

min

,b

est,

max

(91

020

Nm

)S

lip

c

min

,b

est,

max

(m)

Sli

pra

te(m

m/y

ears

)M

ean

recu

rren

ced

min

,b

est,

max

(yea

r)

Ex

plo

rere

85

,17

53

7.5

,1

8.7

57

.4,

7.7

,8

.01

.3,

3.5

,9

.42

.0,

5.4

,1

4.5

14

.4–

18

11

2,

33

3,

1,0

05

Ex

plo

rerf

85

37

.57

.1,

7.4

,7

.70

.5,

1.4

,3

.81

.7,

4.5

,1

2.0

20

–2

56

6,

19

8,

59

8

Hai

da

Gw

aiig

25

0–4

00

32

7.5

,7

.8,

8.2

1.7

,6

.5,

22

2.1

,6

.1,

17

.48

26

6,

76

2,

2,1

74

Mac

ken

zieh

20

0–3

00

30

–8

07

.4,

7.9

,8

.41

.2,

9.1

,49

2.0

,6

.6,

20

.62

98

4,

3,3

01

,1

0,2

98

aE

ffec

tiv

efu

ll-r

up

ture

wid

th(l

ock

edzo

ne

plu

sh

alf

tran

siti

on

zon

e)b

Mo

men

tm

agn

itu

de

from

rup

ture

area

;se

ism

icm

om

ent

from

Mw

(see

Onli

ne

Res

ourc

e2

for

empir

ical

rela

tions)

cM

ean

fault

slip

calc

ula

ted

usi

ng

list

edM

o,

ruptu

rear

ea,

and

anef

fect

ive

shea

rm

odulu

sof

19

10

10

N/m

2

dE

arth

quak

ere

curr

ence

corr

espondin

gto

list

edM

w

eR

up

ture

of

Ex

plo

rer

(fu

ll)

and

Win

on

ase

ctio

ns

(ass

um

edn

ort

hw

ard

lin

ear

dec

reas

eto

zero

inlo

cked

zon

ew

idth

and

slip

)f

Ru

ptu

reo

fE

xp

lore

rse

ctio

no

nly

gF

ault

par

amet

ers

fro

mS

mit

het

al.

(20

03):

ther

mal

lyco

nst

rain

edlo

cked

zon

e(2

2k

m)

and

tran

siti

on

zon

e(2

0k

m);

40

0k

mle

ngth

:m

ax.

po

ten

tial

sub

du

ctio

nex

ten

t.2

50

km

len

gth

:b

ath

ym

etri

cex

pre

ssio

no

fQ

.C

har

lott

eT

erra

ce(a

ccre

tio

nar

yp

rism

)h

So

urc

eo

mit

ted

from

the

cum

ula

tiv

ets

un

ami

haz

ard

du

eto

po

or

con

stra

ints

248 Nat Hazards (2014) 70:237–274

123

2 mm/year intraplate shortening would increase recurrence intervals by *30 %. On

western Haida Gwaii, tsunami run-up may exceed 3 m from local Mw 7.8–8.2 thrust

earthquakes; Mw 7.5 events may result in local maxima of that size (Table 1). run-up is

unlikely to be damaging to the northern mainland and north-western Vancouver Island

from Mw 7.5 events, but may exceed 1.5 m (locally 3 m) in larger events. On south-

western Vancouver Island, local maxima C1.5 m are likely only for the largest events.

Eastern Haida Gwaii may be sheltered; we assume run-up similar to the northern mainland,

but modelling is required.

The occurrence of an Mw 7.7 thrust earthquake off Moresby Island, Haida Gwaii on 27

October 2012 implies that a significant part of the convergence is accommodated by sub-

duction, as in model (1) above (James et al. 2013; Lay et al. 2013). Tsunami impacts docu-

mented on the west coast of Haida Gwaii (e.g. Leonard et al. 2012b) provide a validation of

our empirical methods of tsunami hazard assessment. Although this part of the coast is

unpopulated and there were no witnesses, evidence of tsunami run-up (often exceeding 5 m)

was observed at the heads of many inlets and bays along more than 200 km of coastline.

3.5 Local crustal earthquakes

In the Cascadia forearc (southern Puget Sound to southernmost Georgia Strait; Fig. 4), GPS,

seismicity and palaeoseismic data indicate that *3 mm/year of long-term margin-parallel

shortening is accommodated by crustal earthquakes (e.g. Sherrod et al. 2008). The seismicity

recurrence relation predicts one M C 7 earthquake every*400 years (Hyndman et al. 2003).

None have occurred in the last 150 years, but there is evidence for a tsunamigenic M C 7

thrust earthquake on the Seattle fault*1.1 ka; modelling of an Mw 7.6 source results in[3 m

tsunami waves at Seattle and\1 m in Juan de Fuca Strait (Koshimura et al. 2002). Further

north, the Holocene palaeoseismic record (likely incomplete) includes several events since

*4 ka; mapped fault lengths could produce earthquakes up to M * 7.5 in eastern Juan de

Fuca Strait (e.g. Johnson et al. 2001). In southern Georgia Strait, seafloor offsets indicate

active faulting (Mosher et al. 2000), but displacements are less clear further north. No tsunami

deposits are confirmed around the Canadian Strait of Georgia (e.g. Clague et al. 2000), but

deposits from tsunamis with run-up\3 m are unlikely to be preserved, especially in high-

rainfall regions (Szczucinski 2012).

We outline an approach to assess the tsunami hazard from crustal submarine faults. Based

on the earthquake recurrence relation of a *80 km by 80 km region (box in Fig. 4), we

calculate that Mw 7.5 events may occur every *2,830 years (Table 3), 7.5 being the maxi-

mum magnitude estimated on the Devil’s Mountain fault, which shows Quaternary offsets

offshore Victoria, BC (Hayward et al. 2006). Assuming a uniform frequency of earthquakes

per unit area, we estimate potential tsunami impacts (Table 3). An Mw 7.5 thrust earthquake

can produce mean run-up C1.5 m (locally C3 m) up to 60 km from the source and local

maxima C1.5 m up to 120 km (empirical relations in Online Resource 2). Thus, for any

coastal point, the expected frequency of mean run-up C1.5 m is about half the frequency of

Mw 7.5 earthquakes in a 60-km radius, assuming that roughly 50 % of these earthquakes occur

underwater. Compared with the area discussed above, seismicity rates in the northern and

central Strait of Georgia and western Juan de Fuca Strait are much lower. The related tsunami

hazard is presumably also lower; we assign probabilities that are half those of eastern Juan de

Fuca Strait (Table 3). We include the whole Strait of Georgia in this lower-hazard zone,

because significant seismicity occurs only at the southernmost end.

Another concentration of crustal seismicity occurs in Hecate Strait (location in Fig. 4);

earthquake mechanisms are a combination of thrust and strike slip (Ristau et al. 2007). The

Nat Hazards (2014) 70:237–274 249

123

Ta

ble

3E

stim

atio

no

fp

ote

nti

alim

pac

tat

Pac

ific

coas

tal

site

sfr

om

tsu

nam

isg

ener

ated

by

eart

hq

uak

eso

nlo

cal

cru

stal

fau

lts

Reg

ionsa

Mwb

An

nu

alfr

equen

cyA

(Hr

C1

.5)c

(km

2)

A(2

Hr

C1

.5)c

(km

2)

A(2

Hr

C3

)c(k

m2)

P(H

rC

1.5

m)d

P(H

rC

3m

)d

min

,b

est,

max

min

,b

est,

max

(%in

50

yea

r)m

in,

bes

t,m

ax(%

in5

0y

ear)

JDF

-E&

GI

7.5

0.0

00

1,

0.0

00

4,

0.0

01

1f

5,6

55

22

,61

95

,65

50

.46,

1.5

7,

4.9

70

.23

,0

.79

,2

.49

JDF

-W,

GS

e7

.5L

ow

g0

.23,

0.7

9,

2.4

90

.11

,0

.39

,1

.24

NM

C,

6.8

0.0

03

7h

–9

05

–0

.43

0

CM

C,

7.1

0.0

01

4h

90

53

,470

90

50

.95

0.1

6

EH

G7

.40

.00

01

,0

.000

5h

3,4

70

13

,88

03

,47

00

.27,

1.3

40

.04

,0

.22

aZ

on

esd

efined

inF

ig.

2b

Mw

for

JDF

-E&

GI

bas

edo

nM

xo

fm

app

edfa

ult

s;fo

rH

ecat

eS

trai

t,es

tim

ated

fro

m1

,500

km

2m

axfa

ult

area

(Bu

stin

20

06;

see

On

lin

eR

esou

rce

2fo

rem

pir

ical

rela

tio

ns)

cA

rea

fro

mw

hic

hso

urc

esco

uld

pro

duce

giv

enlo

cal

mea

n(H

r)o

rm

ax(2

Hr)

run-u

pat

apoin

tof

inte

rest

,es

tim

ated

from

empir

ical

rela

tions

giv

enin

Onli

ne

Res

ourc

e2.

Hal

fth

ear

eao

fa

circ

lece

ntr

edo

np

oin

to

fin

tere

st,

assu

med

toli

eh

alf

on

lan

d,

hal

fu

nd

erw

ater

;ra

diu

so

fm

axd

ista

nce

toso

urc

e.F

or

JDF

-E&

GI,

pro

bab

ilit

yo

f2

Hr

fro

m[

60

km

excl

ud

ed(s

uch

area

sli

eu

nd

erla

nd

or

ou

tsid

ese

ism

icit

yar

ea)

dP

rob

abil

ity

of

tsun

ami

run

-up

,fo

rH

rC

1.5

m,b

ased

on

pro

bab

ilit

yo

fea

rth

quak

eo

ccu

rren

ce,av

erag

edo

ver

rele

van

tar

ea.R

edu

ced

by

50

%w

her

eo

nly

2H

rC

1.5

or

3m

isex

pec

ted

eA

reas

po

orl

yco

nst

rain

ed;

arbit

rary

low

pro

bab

ilit

ies

assi

gn

ed(h

alf

tho

seo

fJD

F-E

&G

I)f

Fre

quen

cy(/

6,3

00

km

2)

inJD

F-E

&G

Ifr

om

seis

mic

ity

recu

rren

cere

lati

on

over

giv

enar

ea(s

eeF

ig.

4)

gF

req

uen

cyin

low

-sei

smic

ity

JDF

-Wan

dG

Sas

sum

edto

be

rela

tiv

ely

low

hF

req

uen

cy(/

19

,30

0k

m2)

inH

ecat

eS

trai

tes

tim

ated

from

rup

ture

area

1,5

00

km

2,

effe

ctiv

esh

ear

modulu

s1

91

010

N/m

2,

and

5m

m/y

ear

con

ver

gen

ce(M

azzo

tti

etal

.2

00

3a)

;m

in.

freq

uen

cyfr

om

seis

mic

ity

recu

rren

cere

lati

on

(Bust

in2

00

6)

250 Nat Hazards (2014) 70:237–274

123

long-term strain rate from GPS is significantly larger than that inferred from seismicity,

due to either aseismic/postseismic deformation or infrequent large earthquakes (Bustin

2006). Assuming the latter (parameters in Table 3), earthquakes up to Mw * 7.1 may be

expected every *728 years on average (Table 3). Taking the approach described above

for Juan de Fuca Strait, we estimate the potential tsunami impact on the coasts of eastern

Haida Gwaii and the adjacent mainland (Table 3).

The approach outlined here estimates the maximum probability of potentially damaging

tsunami run-up resulting from submarine crustal earthquakes. Uncertainties remain,

especially in areas of transpression. Fault motion with a strike-slip component will result in

a smaller tsunami than from a pure thrust earthquake of the same magnitude; consequently,

tsunami run-up will be overestimated.

3.6 Local landslides

Locally destructive waves result from landslides on the Pacific coast, with fjords partic-

ularly at risk due to potential failure of their steep walls or submarine delta fronts; rela-

tively small failures can produce high run-up in these long narrow basins. Historical

tsunami sources on the coast of BC and Alaska (Fig. 1) include rockfalls in Knight Inlet

(*16 century and 1999: Bornhold et al. 2007; van Zeyl 2009) and Lituya Bay (1958:

Miller 1960) and submarine slides at Deep Bay (1946: Mosher et al. 2004), Kitimat Arm

(1975: Skvortsov and Bornhold 2007), and Skagway (1994: Kulikov et al. 1996). Some

were triggered by ground shaking (Lituya Bay and Deep Bay) and others by overloading of

sediments at low tide (Kitimat and Skagway); many caused fatalities. Modelling of

potential Fraser River delta foreslope failures suggests that damaging waves could result in

Georgia Strait (Rabinovich et al. 2003). With no frequency–size data available for

potentially tsunamigenic local landslides, we cannot include these sources in the analysis,

but we consider all Pacific coastlines at risk (Sect. 2.4).

3.7 Continental slope landslides

Infrequent ‘‘super-scale’’ slumps offshore Oregon may have produced huge tsunamis

(Goldfinger et al. 2000). Off Vancouver Island, numerous slope failures up to at least 5 km

wide along the Cascadia deformation front appear to be fault controlled and pre-condi-

tioned to fail by strength contrasts between gas-charged and overlying gas hydrate-

cemented sediments (Lopez et al. 2010). Failures are most likely during strong ground

shaking; those that coincide with megathrust seafloor displacement could locally amplify

tsunamis as suggested for the 2004 Sumatra event (Tappin et al. 2007). Slope stability

analysis is needed to determine whether tsunamigenic failures may be triggered by smaller

earthquakes. If continental slope failures were assumed to occur at a similar rate as we

estimate for the Atlantic margin (Sect. 4.5), the hazard is negligible compared to other

Pacific tsunami sources. Making the assumption that failures will only occur during

Cascadia megathrust earthquakes, we omit them as an independent source; they should be

incorporated into probabilistic tsunami hazard analyses of the subduction zone.

3.8 Hawaiian Islands landslides

Palaeotsunami data inferring up to *400 m run-up on the south-eastern Hawaiian Islands

are consistent with dating and modelling of debris avalanches off Mauna Loa (Moore et al.

Nat Hazards (2014) 70:237–274 251

123

1994; McMurtry et al. 2004). Based on tsunami modelling of a *5,000-km3 deposit, Ward

(2001) suggests that a volcanic flank collapse could lead to *20 m run-up on the Pacific

coast of North America. However, dispersion was neglected in Ward’s analysis; significant

attenuation is likely to occur over *4,000 km of propagation (Pararas-Carayannis 2002).

For example, more realistic models of Canary Island flank collapses result in wave

amplitudes of *1–3 m after similar propagation in the Atlantic (e.g. Mader 2001), i.e.

*10 % of the Ward (2001) values. Without further constraints, but consistent with the

above models, we assume that large failures could lead to mean tsunami run-up C1.5 m on

the west coast of Canada, but mean run-up C3 m could only result from north-eastern flank

failures, where the main energy would be directed towards Pacific Canada.

Eleven giant Hawaiian submarine landslides have occurred since 1 Ma (McMurtry

et al. 2004). We take the average recurrence interval (*91 ky) as a minimum for local

run-up maxima C1.5 m on the Pacific coast (Table 1). Assuming that failures are equally

likely on all three flanks of the islands, and that western flank slides are the least

hazardous due to wave scattering through the Hawaiian Islands, provides a best estimate

of *136 ky recurrence for mean run-up C1.5 m. We calculate a maximum recurrence

interval of *272 ky with the possibility that only failures on the north-eastern flanks

may present a tsunami hazard for Canada; we choose this value as a minimum recur-

rence for mean run-up C3 m. Thus, the Pacific coast tsunami hazard from Hawaiian

landslides is negligible. Without detailed modelling, we assume that Georgia Strait

would be protected from hazardous waves, but we apply equal hazard to all other Pacific

coastal zones (Table 1).

3.9 Aleutian Islands landslides

Waythomas et al. (2009) modelled eight hypothetical submarine failures

(160–1,440 km3) off the Aleutian margin (Fig. 1). Results suggest a threshold volume of

*400 km3 for slides that may result in tsunami run-up C1.5 m on the Canadian Pacific

coast; run-up greater than *3 m appears possible for failures exceeding *500 km3 in

the central Aleutian arc and *800 km3 in the west. The region has not been surveyed at

a resolution adequate to constrain either the size or frequency of large failures in the

past; Waythomas et al. (2009) use a probabilistic approach developed by Watts (2004) to

calculate an average recurrence interval of *2 ky for failures C100 km3 that could be

triggered by Mw 5.1–8.9 earthquakes. Without further constraints (P. Watts, personal

communication 2011), we assume approximate best estimates of 3 ky (2.5–5 ky) and 4

ky (3.5–8 ky) for the recurrence of mean tsunami run-up exceeding 1.5 and 3 m,

respectively. Without detailed modelling, we assume that Georgia Strait would be pro-

tected from potentially damaging waves, but we apply equal hazard to all other Pacific

coastal zones (Table 1).

4 Atlantic coast

Potential tsunami sources for the east coast of Canada include a few compressive plate

boundaries around the Atlantic Ocean (e.g. 1755 Lisbon tsunami generated offshore south-

western Portugal and observed in Newfoundland; Caribbean subduction), local landslide

sources (e.g. 1929 Grand Banks event), and far-field landslide sources in the eastern

Atlantic Ocean.

252 Nat Hazards (2014) 70:237–274

123

4.1 Gibraltar-Cadiz plate boundary

The catastrophic 1755 Mw * 8.5–9 Lisbon earthquake triggered a damaging tsunami with

many fatalities in western Europe and Morocco (Solares and Arroyo 2004; Muir-Wood and

Mignan 2009). Far-field waves occurred in Brazil and the Caribbean; at Bonavista,

Newfoundland, the harbour drained and then flooded, as did adjacent meadows (Tocque

1846; Ruffman 2006). The source of the earthquake is still under debate, but it must lie in

the region of the diffuse Africa-Eurasia plate boundary off south-western Iberia (e.g.

Baptista et al. 2003; Fig. 1). Of several proposed faults, only the Gibraltar-Cadiz thrust

(GCT) is large enough for an Mw 8.5 event, and an additional source to the northwest may

be required to match all the observations (Gutscher et al. 2009). We assess a range of

possibilities where large earthquakes occur on one or more dip-slip faults; source attri-

butes, slip, and recurrence time required for Mw 8.5 earthquakes are listed in Table 4.

In deep-sea basins offshore the south-western Iberian peninsula, turbidite deposits

dating from the last *9 ky (correlated up to *175 km) probably result from shaking in

large earthquakes (e.g. Gracia et al. 2010). Since 7 ka, all widespread turbidites but one

correlate with tsunami deposits on the coasts of south-western Spain and Portugal (Luque

et al. 2002; Baptista and Miranda 2009); the youngest likely represents the M 8.0 1969

Horseshoe earthquake that triggered only a small tsunami (Gracia et al. 2010). Assuming

that the remaining six extensive turbidite events in the last *9 ky (or five if one had a

climatic trigger) were sourced by events similar to the 1755 event, and that the record is

complete for M * 8.5 events, provides a mean recurrence interval of *1,730 years (or

*2,130 years). This is broadly compatible with proposed fault sources (Table 4).

Observations of relative wave amplitudes in 1755 are matched by modelling of the Mar-

ques de Pombal-Guadalquivir (MdP-G) source (Roger et al. 2010, 2011) or a source oriented

*NNE-SSW such as the GCT (Barkan et al. 2009; Muir-Wood and Mignan 2009). For the

former, tsunami energy is directed towards Newfoundland and South America (Fig. 5). Using

high-resolution bathymetric and topographic grids, run-up exceeds 1.5 m along much of the

south-eastern Newfoundland coast and matches historical accounts of inundation on the

Bonavista peninsula (up to 2.5 m or more; Roger et al. 2010). We consider both the MdP-G

(Baptista et al. 2003) and GCT faults (e.g. Thiebot and Gutscher 2006) viable sources of

‘‘Lisbon-type’’ events, capable of producing run-up in excess of 1.5 m in Atlantic Canada and

possibly greater than 3 m in places (Table 1). We assume that the inner coastlines would be

relatively sheltered, with only local maxima C1.5 m.

4.2 Puerto Rico trench

At the Greater Antilles margin (Fig. 1), convergence is very oblique, with *5 mm/year

compression on the Northern Hispaniola fault (NHF) and*22 mm/year left-lateral strike slip

on the Septentrional and Enriquillo faults (Calais et al. 2002). A 1946–1953 earthquake series

on the NHF included an Ms 8.1 event that spawned up to 5-m tsunami waves, killing 1,790

people in the Dominican Republic (Lander et al. 2002). Further east, strain partitioning is less

apparent. The North America plate subducts obliquely at the Puerto Rico trench beneath

Puerto Rico and the Virgin Islands (relative motion *17 mm/year; Jansma et al. 2000), but

left-lateral slip is likely on offshore faults (LaForge and McCann 2005). Interface slip is

generally oblique, but some events are more orthogonal (e.g. 1943 Mw 7.8; Doser et al. 2005).

No observable tsunami occurred in 1943 or from a 1787 M * 8 event likely on the Puerto

Rico trench (McCann 1985). In addition, modelling of GPS data from Puerto Rico and the

Virgin Islands suggests little or no trench-perpendicular compressional strain accumulation

Nat Hazards (2014) 70:237–274 253

123

Ta

ble

4E

stim

atio

no

fG

ibra

ltar

-Cad

izM

w*

8.5

eart

hquak

ere

curr

ence

from

pote

nti

alfa

ult

sourc

esan

dpal

aeose

ism

icdat

a

Fau

ltso

urc

eaL

eng

th(k

m)

Wid

th(k

m)

Th

ick

nes

s(k

m)

Are

a(k

m2)

Dip

(�)

l(N

/m2)

Sli

pb

(m)

HC

c(m

)S

lip

rate

(mm

/yea

r)M

ean

recu

rren

ced

(yea

r)

Md

P-G

1,1

00

02

4–

45

39

10

10

20

14

–1

84

–5

e2

,80

0–

4,5

00

Ho

rses

ho

ep

lain

17

5–3

50

50

55

69

10

10

10

5.8

4–

5e

1,1

50

–1

,45

0

Gib

ralt

ar-C

adiz

18

02

10

39

10

10

55

2.5

–5

f1

,00

0–

2,0

00

Pal

aeo

seis

mic

1,9

51

(86

8–

2,7

90

)

aP

ote

nti

also

urc

eso

f‘‘

Lis

bo

n-t

yp

e’’

eart

hq

uak

es.M

dP

-G:

Mar

qu

esd

eP

om

bal

-Gu

adal

qu

ivir

(Bap

tist

aet

al.2

00

3);

Ho

rses

hoe

pla

in(S

tich

etal

.2

00

7);

Gib

ralt

ar-C

adiz

thru

st(T

hie

bo

tan

dG

uts

cher

20

06);

pal

aeose

ism

icre

cord

(turb

idit

ean

dts

unam

idep

osi

ts)

of

Lis

bon-t

ype

even

tssi

nce

*9

ka

(Gra

cia

etal

.2

01

0)

bS

lip

req

uir

edto

pro

duce

anM

w8.5

eart

hquak

e(s

eeO

nli

ne

Res

ourc

e2

for

empir

ical

rela

tions)

cH

ori

zon

tal

con

ver

gen

ceeq

uiv

alen

tto

slip

on

the

dip

pin

gfa

ult

sou

rce

dE

arth

qu

ake

recu

rren

ceco

rres

po

nd

ing

toM

w8.5

and

slip

rate

val

ues

.P

alae

ose

ism

icval

ues

are

use

din

the

anal

ysi

se

Afr

ica-

Eu

rasi

aco

nv

erg

ence

rate

(Fer

nan

des

etal

.2

00

3)

fG

ibra

ltar

-Cad

izsu

bd

uct

ion

rate

assu

med

equ

ival

ent

tow

estw

ard

Gib

ralt

arb

lock

mo

tio

n(u

pp

erp

late

tren

chw

ard

moti

on

),re

sult

ing

fro

mex

tru

sio

ndue

totr

ansp

ress

ion

and/

or

slab

roll

bac

k(F

ern

and

eset

al.

20

07;

Ver

nan

tet

al.

20

10)

254 Nat Hazards (2014) 70:237–274

123

(ten Brink and Lopez-Venegas, 2012). An M * 9 event may thus be unlikely, but cannot be

discounted. Globally, the last two M C 9 earthquakes occurred on subduction zones where

such events were deemed impossible due to very oblique convergence (northern Sumatra

2004) or low apparent coupling with frequent moderate events (Japan trench 2011; e.g.

Gutscher and Westbrook 2009).

Turbidites (undated) in the Puerto Rico trench may record plate-boundary earthquakes

(Pilkey 1988). The largest extends for at least 300 km; five entry points suggest widespread

shaking. Probable tsunami deposits are linked to a tsunami in 1918 (Scheffers et al. 2005);

another is dated at maximum 4.2 ka (Taggart et al. 1993). A south-directed deposit on

Anegada Island, dating from *1650 to 1800 CE, may derive from the 1755 Lisbon

tsunami, but modelling suggests a local origin, possibly a 1690 Antilles event of unknown

source (Atwater et al. 2010; Buckley et al. 2011).

Geist and Parsons (2009) estimate a frequency–magnitude relation for the Greater

Antilles based on tectonic slip rates, full locking, global subduction earthquake statistics,

and a 5–30-km deep seismogenic zone dipping at 20�. On the 1,100 km length of the

Puerto Rico trench and NHF, they calculate average recurrence intervals of *300, 1,000,

and *3,000–7,000 years for earthquakes of M 8, 8.5, and 9, respectively.

Modelling shows that a tsunami originating in the Puerto Rico trench region would

propagate more efficiently to Halifax than from any other far-field source (Fig. 6; Xu

2007). Due to the trench orientation, tsunami energy would be directed largely north

towards the Canadian continental margin. Models of an M * 9 thrust earthquake on the

Puerto Rico trench (including the NHF) simulate damaging waves along the US Atlantic

coast, *0.5–3 m at the shelf edge with higher run-up expected (Knight 2006; AGMTHAG

2008; Geist and Parsons 2009). Amplitudes have not been modelled for Atlantic Canada,

but similar or higher values are likely. Modelling is required to determine the magnitude

threshold for earthquakes that could cause damaging run-up in Canada. Based on historical

global tsunami run-up (Online Resource 2), we assume that only M C 8.5 earthquakes

could produce run-up exceeding 1.5 m on most of the Canadian Atlantic coast; we further

assume that run-up from large events could exceed 3 m in places (Table 1). We assume

Fig. 5 Far-field simulation of the 1755 Lisbon tsunami (after Roger et al. 2010). Maximum modelled waveamplitudes (above state of tide) after 9 h of tsunami propagation using MdP-Guadalquivir compositeearthquake source of Baptista et al. (2003)

Nat Hazards (2014) 70:237–274 255

123

that the north-eastern Atlantic coast would be relatively sheltered (local maxima C1.5 m).

Based on the frequency of M 8.5–9 earthquakes given above, we choose 1,000 and

7,000 years as the minimum and maximum recurrence of run-up C1.5 m, with the median

4,000 years as a best estimate.

4.3 Lesser Antilles subduction zone

Convergence is orthogonal in the Lesser Antilles. Preliminary thermal modelling indicates

a southward increase in the megathrust seismogenic zone width due to widening of the

accretionary wedge (Gutscher et al. 2010). Thrust interface earthquakes (up to M * 7.5)

have occurred north of 14.5�N, but not to the south, suggesting that the fault may be locked

(Gutscher et al. 2010) or that subduction is aseismic. Gutscher et al. (2010) suggest that a

*500-km rupture could produce earthquakes up to M * 9; Heuret et al. (2011) note the

potential for M C 8.5 events.

The only tsunami deposits described (on Guadeloupe; *2.4–2.7 ka) are likely sourced

by a submarine slide (Scheffers et al. 2005). The apparent paucity of tsunami deposits is

likely due to poor preservation in the tropical climate and difficulty in distinguishing

between tsunami and hurricane sources (e.g. Day et al. 2008). There are no published

recurrence estimates for great earthquakes in the Lesser Antilles. Without further con-

straints, we assume a range of recurrence intervals based on a comparison with the Cas-

cadia subduction zone (Sect. 3.2). The convergence rate (*20 mm/year; e.g. DeMets et al.

2010) is approximately half that of Cascadia; thus, we assume a range of recurrence

intervals for M * 9 events that is double (1,060 ± 520 years).

There has been no far-field modelling of a Lesser Antilles megathrust tsunami; in

theory, most far-field energy would be directed eastward, but the potential for damage to

eastern Canada warrants further investigation. Edge waves from the 1960 Chilean tsunami

caused some damage in Vancouver Island inlets, although most energy was directed

elsewhere in the Pacific. Also, the Mid-Atlantic Ridge may steer tsunami waves towards

Fig. 6 Tsunami gain map for Halifax, NS, produced using the all-source Green’s function (figure modifiedfrom Xu 2007). Colours show the wave amplitude at Halifax relative to one unit of vertical displacement atAtlantic Ocean sources with dimensions of 100 km 9 100 km. E.g. a 1,000-km2 source at the Puerto Ricotrench with 10 m initial sea surface displacement results in *1.5 m amplitudes at Halifax. The modelinvolves propagation only; amplitudes would vary if source and run-up effects were considered

256 Nat Hazards (2014) 70:237–274

123

Canada (Fig. 6), as occurred during the 2004 Indian Ocean tsunami (Thomson et al. 2007).

We assume that M * 9 events could produce run-up C1.5 m (locally C3 m) along the

outer coasts of Nova Scotia, southern Newfoundland, and New Brunswick, that run-up

C1.5 m is possible but less likely for the inner coasts and south-eastern Newfoundland, and

that the north-eastern Atlantic coast is sheltered from such waves (Table 1).

4.4 St. Lawrence estuary

Seismicity along the St. Lawrence estuary, Quebec (Fig. 1), likely occurs as a result of

compressional reactivation of Palaeozoic normal faults (Iapetan rift zone) under the

present-day stress field, which is affected by postglacial rebound (e.g. Mazzotti et al.

2005). Earthquake focal mechanisms mostly reflect oblique thrust faulting; tectonic tsu-

namis could result from displacement of the river/estuary floor.

Tsunami modelling of tectonic and landslide sources in the estuary simulates local wave

amplitudes up to 5 m or more; larger sources lead to 1–2 m waves propagating for at least

30 km (El-Sabh et al. 1988; Poncet et al. 2010). Numerous failures are noted along the

estuarine banks and submarine slopes; many relate to the 1663 M * 7–7.5 Charlevoix

earthquake (Locat et al. 2003; Ebel 2011). In the Betsiamites delta region (lower St.

Lawrence), at least three large failures occurred in the last *10 ky, one of which correlates

with the 1663 event (Cauchon-Voyer et al. 2011), suggesting an average recurrence

equivalent to that of M * 7.2–7.8 events in the Charlevoix seismic zone.

With few constraints, we assume that the tectonic tsunami hazard in the St. Lawrence

estuary (inner Atlantic zone; Fig. 2) is negligible, but we take a conservative approach. For

the maximum level of hazard, we assume that run-up C1.5 m (locally exceeding 3 m) may

occur with a return period of *3,300 years (equivalent to the return period of large failures

and presumably large earthquakes, in the lower St. Lawrence). The estuary is also high-

lighted as susceptible to landslide-triggered waves. The shorelines of the Great Lakes

(particularly Lakes Ontario and Erie) may also be subject to tsunami hazard from tectonic

and landslide sources (e.g. Dineva et al. 2004), but inland waterways are outside the scope

of this marine coastline assessment.

4.5 Continental slope landslides

The only historical landslide tsunami on the east coast of North America, triggered by

earthquake shaking at the Grand Banks in 1929 (Fig. 1), caused 28 deaths in Newfound-

land (Heezen and Ewing 1952; Fine et al. 2005). Failures much larger than the 1929 slide

have been mapped (e.g. Piper and McCall 2003; Mosher et al. 2010). Many factors may be

important in tsunamigenesis and propagation (see Online Resource 3), but landslide vol-

ume is a primary control and the only one that can be approximated from available data.

Tsunami modelling of selected Atlantic margin slides and observations from 1929 (details

in Online Resource 3) provide a basis for estimating the threshold size for damaging

tsunamigenic failures. We estimate that tsunami run-up exceeding 1.5 and 3 m may result

from failures that exceed volumes of 40 and 55 km3, respectively. These thresholds may

vary along the coast due to factors such as varying shelf width or the presence of inlets

where resonance amplification may occur.

The frequency of continental slope landslides has varied through the Quaternary, higher

at times of glaciation and deglaciation; on the south-eastern Canadian margin, recurrence

intervals have been estimated at *5 ky for ‘‘small’’ failures and [200 ky for ‘‘large’’

failures (Piper et al. 2003; Piper 2005). High-resolution studies have identified more

Nat Hazards (2014) 70:237–274 257

123

failures in some areas including Flemish Pass and Orphan Basin off south-eastern New-

foundland (Online Resource 3), where recurrence relations can be approximated using

along-slope extent as a proxy for failure size (Fig. 7). However, for most of the margin, the

record of smaller failures closer to the threshold size is incomplete. For the little-studied

Labrador margin (Fig. 1), preliminary data indicate a similar failure distribution to the

Scotian Slope and Orphan Basin (Piper 2007).

To estimate the frequency of potentially damaging tsunami run-up, it is first necessary

to calculate the failure extent associated with a 40-km3 volume. We assume a failure style

similar to the Grand Banks slide. Shallow retrogressive failures (see Online Resource 3)

typically involve the upper 10–20 m of sediment; downdip length likely scales with along-

slope extent. For US Atlantic margin slides, Chaytor et al. (2009) calculate a relation

between volume V and area A: V = 0.0163A1.1. The Grand Banks slide is estimated at

150–200 km3 with an area of *10,000 km2 (Fine et al. 2005; Mosher and Piper 2007),

suggesting a similar relation, V = 0.02A. Relating the volume to the *250 km length

(L) provides V = 0.8L; thus, a failure with a 40-km3 volume has a predicted along-slope

extent of *50 km and a 55-km3 failure an extent of *70 km.

For the outer Atlantic coast, we assume that mean run-up exceeding 1.5 and 3 m could

result from failures with along-slope extents exceeding 50 and 70 km, respectively (only

local maxima for inner coastlines; Table 1). In Orphan Basin, the expected frequency of

slides C50 km in extent is *0.1/ky (Fig. 7), equating to an average recurrence interval of

*10 ky. Similar failures in Flemish Pass have an estimated mean recurrence time of *21

ky. Slope failures with extent C70 km occur every *11.5 ky in Orphan Basin and *45 ky

in Flemish Pass. In the absence of adequate data, we assume that the event frequencies

Fig. 7 Cumulative frequency distributions of Atlantic submarine landslides. Thick grey lines approximationof frequency of failures that could produce tsunami run-up C1.5 m at the Canadian Atlantic coast. Leftcontinental slope failures in 2 areas off Newfoundland since 85 ka (Orphan Basin) and 400 ka (FlemishPass). Data for Orphan Basin (Piper et al. 2011): mass-transport deposits (from ultra-high-resolution seismicreflection data) correlated with cored turbidites. Data for Flemish Pass: large failures since *400 ka (extentC90 km; from seismic profiles; Piper and Campbell 2005); smaller failures since *165 ka (extent C40 km;from higher-resolution data; Huppertz and Piper 2009). Assumed maximum failure extent: 400–600 km inOrphan Basin; 160–200 km in Flemish Pass. Right large volcanic flank failures in the western CanaryIslands of La Palma, El Hierro, and Tenerife since *1 Ma (data from Masson et al. 2002); maximumassumed failure size: from 650 km3 (largest observed) to 1,000 km3

258 Nat Hazards (2014) 70:237–274

123

estimated for Orphan Basin are representative of the entire Canadian Atlantic margin. For

events close to threshold size, most coastal sites are assumed at risk from only one source

section of the continental slope. However, due to the bend in the margin off south-eastern

Newfoundland and the likely landward tsunami travel paths, this zone is considered at risk

from landslide-generated tsunamis from three source areas: Orphan Basin, Flemish Pass,

and the Salar Basin (Online Resource 3). Modelling is required to test this assumption.

4.6 Puerto Rico trench landslides

Large landslides north of Puerto Rico may present a tsunami hazard to the Atlantic coast of

Canada, with an expected directionality similar to tsunamis triggered by earthquakes on the

Puerto Rico trench (Fig. 6). With long estimated recurrence intervals ([200 ky; ten Brink

et al. 2006) and the assumption that such landslides would occur at the same time as

megathrust earthquakes, we omit them as an independent source.

4.7 Canary Islands landslides

Landslide deposits off the western Canary Islands (Fig. 1) suggest major volcanic flank

collapses, although the presence of turbidite sub-units may infer several smaller stages for

each collapse (Masson et al. 2002; Wynn and Masson 2003). Modelling of a western flank

failure of La Palma produces catastrophic tsunami waves at the North American coast

(10–25 m amplitudes for a 500-km3 slide; Ward and Day 2001). However, dispersion

effects were omitted; more realistic models (Mader 2001; Gisler et al. 2006) produce far-

field waves on the North American east coast of no more than 1 or 3 m, with damaging

run-up possible.

Twelve flank failures (\50 to *650 km3) since *1 Ma are identified, the most recent

aged *15 ka (Masson et al. 2002). A 500-km3 slide is expected every *590 ky (Fig. 7).

We choose a threshold volume of 150 km3 (recurrence interval *160 ky; *97–313 ky),

for which Ward and Day (2001) model 3–8-m tsunami waves; we assume that a more

realistic model would yield amplitudes an order of magnitude smaller that may reach 1.5 m

run-up in Atlantic Canada. Although the models of Ward and Day (2001, 2005) overes-

timate amplitudes for a Canary Islands tsunami, they do reveal the relative impact of such

an event along the Canadian Atlantic coast. Focussing by the Laurentian Channel may lead

to amplitudes at some inner coastlines as large as those on the exposed outer coast. Thus,

we apply equal hazard to all susceptible zones, assuming that events with a 160 ky

recurrence could result in mean run-up C1.5 m, 313 ky events could produce mean run-up

C3 m, and 97 ky events could produce only local maxima C1.5 m (Table 1).

4.8 Storegga landslides

Massive landslides have occurred offshore Norway; the Storegga slide complex (Fig. 1)

has failed every *100 ky. For the latest and largest slide (*3,500 km3; *8.1 ka; Solheim

et al. 2005), 3 to [20 m run-up is inferred from tsunami deposits within *900 km

(Bondevik et al. 2005). However, coastal wave run-up in the western Atlantic would not

likely exceed 1.5 m due to attenuation over the[3,000-km distance and wave scattering by

Iceland and the Faroe Islands. Also, the Storegga area is considered currently stable, with

another glacial cycle required to lead to failure (Bryn et al. 2005). Thus, we exclude this

source from the tsunami hazard assessment.

Nat Hazards (2014) 70:237–274 259

123

5 Arctic coast

The Canadian Arctic has only recent tide gauge monitoring, a short written and poorly

recorded oral history; the only potential tsunami account concerns the disappearance of an

Inuit hunting party at the time of a strong earthquake in the 1860s (Ruffman and Murty

2006). Along the Beaufort Sea coast, deposits are linked to observed storm surges in 1944

and 1970 (Reimnitz and Maurer 1978); it can be impossible to distinguish between tsunami

and storm surge deposits (e.g. Shanmugam 2011), so some similar layers may instead

record tsunamis.

Given the lack of subduction zones, Arctic tsunami hazard is probably low, but locally

damaging waves may be generated by landslides, iceberg calving, or jokulhlaups. Far-field

tsunami damage is unlikely. Local tectonic tsunamis are improbable; no tsunamigenic

structures are defined by earthquake concentrations, which relate to postglacial extensional

faulting on eastern Baffin Island and mainly strike-slip faulting in Baffin Bay (Bent 2002).

Deep seismicity beneath the Beaufort Sea slope (Fig. 8) is interpreted as extensional

faulting due to sediment loading (Hasegawa et al. 1979). However, we do consider a

possible fault source beneath the Mackenzie delta (Beaufort Sea). We do not account for

ice effects; ice breakup and transport could increase the potential for near-field damage, but

significant wave attenuation is likely further afield, based on contrasting phenomena during

storm surge events in periods of low versus high ice cover (e.g. Reimnitz and Maurer

1978).

5.1 Mackenzie thrust

GPS data show that the northern Cordillera is moving to the north-northeast relative to the

craton to the east; this is partly taken up by thrust and strike-slip faulting in the Mackenzie

and Richardson Mountains (Leonard et al. 2007; Fig. 8), with a probable *2 mm/year

northward residual motion. A lack of seismicity in the Mackenzie delta region indicates

that either infrequent large thrust earthquakes or aseismic deformation may occur. A thin-

skinned fold-and-thrust belt formed from Palaeocene–Eocene compression; local Holocene

deformation is apparent (Lane 2002). A possible outcome of the ongoing compression is

thrusting of the delta sediments and underlying continental crust northward over the

oceanic crust on a decollement (Hyndman et al. 2005).

In the hypothesis that large tsunamigenic earthquakes may occur on such a locked thrust

fault (Fig. 8), a reasonable range of fault dimensions would indicate Mw 7.9 (7.4–8.4)

events with an average recurrence interval of *3,300 years (Table 2). In the southern

Canada Basin, sediment cores dating back to *8 ka contain four distal turbidites with a

composition consistent with a provenance of Pleistocene–Holocene Mackenzie delta

sediments; Grantz et al. (1996) suggest an origin of earthquake-triggered shallow sub-

marine slides. An average turbidite recurrence interval of *2 ky is consistent with

hypothetical Mackenzie thrust events, but also with the statistical recurrence of M * 7

earthquakes from historical Beaufort Sea seismicity.

An Mw 8.4 Mackenzie thrust earthquake could lead to tsunami run-up C3 m along the

coasts of Yukon and western Northwest Territories, with localised run-up C1.5 m possible

at Banks and Victoria Islands, and north-western Nunavut (Fig. 1). Smaller events could

produce only local maxima C1.5 m. Given the extremely poor constraints, further research

is required before the Mackenzie thrust can be considered a known hazard; thus, we

exclude it from the tsunami hazard assessment.

260 Nat Hazards (2014) 70:237–274

123

5.2 Local landslides

No landslide tsunamis are documented on the Arctic coastline, but several failures are

noted (e.g. Syvitski et al. 1987). Fjords, common along the High Arctic Island coasts and

often populated, are particularly susceptible. Iceberg calving from coastal glaciers and ice

shelves pose a similar hazard, with damaging waves reported from such events, and from

landslides, in Greenland (Ruffman and Murty 2006; Amundson et al. 2008; Macayeal et al.

Fig. 8 Tectonics of the northern Cordillera and the potential Mackenzie thrust (dashed yellow line).Mapped faults shown by black lines, seismicity by orange circles and earthquake mechanisms. Paired yellowarrows indicate styles of current deformation; single thicker arrow indicates northward residual motion.White star location of large submarine landslide scar (Hill et al. 1982). AK Alaska, ELFZ Eskimo Lakes faultzone, NT Northwest Territories, YT Yukon Territory

Nat Hazards (2014) 70:237–274 261

123

262 Nat Hazards (2014) 70:237–274

123

2011). Jokulhlaups from coastal glacier termini could also produce locally damaging

waves (e.g. Lewis et al. 2007).

5.3 Continental slope landslides

Retrogressive failures and slumps are widely evident on the Beaufort continental slope

(Hill et al. 1982; Kayen and Lee 1991). Canada Basin is filled with turbidites that likely

derive from stacked mass-transport deposits on the Beaufort and Arctic Archipelago

margins (Mosher et al. 2012); with high sedimentation rates, the Mackenzie delta may have

tsunamigenic failure potential. Hill et al. (1982) speculate that a large landslide scar at the

southern Beaufort shelf–slope interface (Fig. 8) results from creep, but other failure modes

are possible, e.g. failure on a weak gassified layer caused by gas hydrate dissociation (e.g.

Kayen and Lee 1991). Mosher (2009) suggests that the slide may have been tsunamigenic.

No failures have been dated. Southern Canada Basin turbidites, which likely relate to

Mackenzie delta failures of unknown tsunami potential, have an average Holocene

recurrence interval of *2 ky (Grantz et al. 1996).

Fig. 9 Probabilistic tsunami hazard of Canada; colours correspond to the probability of exceedance ofgiven tsunami run-up in a 50-year period (best estimate cumulative values) a Tsunami hazard map forprobability of potentially damaging run-up (exceeding 1.5 m) in 50 years. b Tsunami hazard map forprobability of significant run-up (exceeding 3 m) in 50 years. c Tsunami hazard comparison betweenrepresentative regions (zones defined in Fig. 2). Minimum, best, and maximum cumulative probability ofexceedance (in 50 years) of tsunami run-up of 1.5 m (left) and 3 m (right) for each region

b

Fig. 9 continued

Nat Hazards (2014) 70:237–274 263

123

Debris flow deposits on the Baffin margin (Aksu and Piper 1987) may result from

gradual glaciogenic processes or potentially tsunamigenic failures. The turbidite record

from an abyssal plain site suggests at least eight events of unknown origin in the last *50

ky (Benetti 2006); further studies are needed to constrain the history of slope failures. No

disturbances are apparent in the epicentral region of the 1933 Mw 7.4 Baffin Bay earth-

quake (Li et al. 2010).

Earthquake shaking may be necessary to trigger large continental slope failures, but the

sediment may be pre-conditioned to fail by various factors (e.g. Mosher et al. 1994). A

similar seismic hazard level is calculated for the Arctic and Atlantic margins (Halchuk and

Adams 2008). In the absence of better constraints, we assume that potentially damaging

tsunami run-up due to continental slope landslides may occur along the margins of Canada

Basin and Baffin Bay at a rate similar to that calculated for the Atlantic margin (Table 1).

Without constraints on how the Arctic coastline may be affected, we apply a uniform

hazard to the entire zone, although some coastlines are undoubtedly sheltered from slope

failure tsunamis, particularly when ice cover is high (most of the year); hazard will be

higher at times of lower ice extent.

6 Cumulative tsunami hazard of the Canadian coastline

The estimated cumulative tsunami hazard is given for each zone in Fig. 9 in terms of probability

of exceedance of 1.5 and 3 m run-up in a 50-year period (Online Resource 4 also shows the

maximum run-up expected within 100 to 2,500-year time periods). Regions with apparent

negligible hazard may still be vulnerable to local landslide-generated waves (Fig. 10). Tsunami

hazard on the outer Pacific coast (P(Hr C 1.5) & 40–80 %; R & 30–100 years equivalent

mean recurrence interval) is, respectively, one and two orders of magnitude greater than the outer

Atlantic (P(Hr C 1.5) & 1–15 %; R & 300–1,700 years) and Arctic (P(Hr C 1.5) \ 1 %;

R & 6,500–17,000 years). For run-up exceeding 3 m, the Pacific hazard (P(Hr C 3) &10–30 %; R & 150–560 years) is significantly higher than both the Atlantic (P(Hr C 3) &1–5 %; R & 650–4,000 years) and Arctic (P(Hr C 3) \1 %; R & 7,000–20,000 years). On

the outer Pacific coast, the 1.5 m? hazard is dominated by far-field subduction zones; the

3 m? hazard is almost entirely contributed by local megathrust faults (Fig. 11). For inner Pacific

coasts (Juan de Fuca and Georgia Straits), the Cascadia subduction zone contributes most hazard

at both levels. Tsunami hazard on the Atlantic coast is dominated by poorly constrained far-field

subduction zone sources at both the 1.5? and 3 m? run-up levels. Tsunami hazard on the Arctic

coastline remains uncertain, but this region is assumed to be sheltered from far-field tsunamis; the

hazard is provided by local landslide sources.

7 Discussion and recommendations

The assessment presented here quantifies the tsunami hazard from potential sources in a

simplified probabilistic way to estimate probabilities of run-up at two levels (C1.5 and

3 m), but we do not assess the probabilities of larger run-up (critical for emergency/land-

use guidelines, etc.). An improved probabilistic hazard assessment will require quantifi-

cation of a wide range of source scenarios, as well as tsunami modelling using fine-

resolution bathymetry and onland topography.

Representing the greatest overall tsunami hazard, improved tsunami models of the

Cascadia subduction zone are needed that involve a greater range of source models and

264 Nat Hazards (2014) 70:237–274

123

detailed bathymetric/topographic data to more accurately assess coastal run-up. Hazard

analysis will also benefit from comparison with other subduction zones. At near-field sites,

rupture variations have a large impact on run-up; for the same earthquake magnitude, a

narrow rupture produces a larger tsunami than a wide one, due to the shallower concen-

tration of fault slip beneath relatively deep water (Geist 2005). Splay fault rupture and

landslides also influence local run-up, as indicated for the 1964 Alaska and 2004 Sumatra

events (DeDontney and Rice 2011; Suleimani et al. 2011). A useful case study by Priest

et al. (2009) assesses potential tsunami run-up at Cannon Beach, Oregon, from a wide

range of Cascadia megathrust rupture scenarios. Variations include high-slip patches and

splay fault rupture; Mw 8.3–9.4 earthquakes with *8–38 m fault slip result in maximum

run-up of 9–30 m (preferred scenario *10 m). Additional variations could incorporate

large coseismic failures at the deformation front, and rupture that extends northward to

include subduction of the Explorer plate.

The October 2012 Haida Gwaii tsunami on the Queen Charlotte margin provides a

useful validation of our empirical methods for assessing tsunami hazard in areas with a

lack of historical events. Documentation of significant tsunami run-up (locally more than

8 m) confirms the contribution of a historically quiescent subduction thrust to the tsunami

hazard of western Haida Gwaii. Ongoing analysis of this event will contribute to an

improved understanding of tsunami hazard in the region. Subduction along the Explorer

margin, also historically quiescent, is identified as a source of significant tsunami hazard

Fig. 10 Canadian coastlines (blue) susceptible to local waves triggered by subaerial or submarinelandslides or glacial calving. Hazard is based on the landslide susceptibility map of Bobrowsky andDominguez (2012) and on the presence of glacial fjords

Nat Hazards (2014) 70:237–274 265

123

for north-western Vancouver Island (Online Resource 4). Better constraints are needed on

the extent and nature of subduction, as well as modelling to better assess the tsunamigenic

potential of such megathrust earthquakes; parallels can be drawn from observations of the

Haida Gwaii event. Future work should also seek evidence for past tsunamigenic earth-

quakes (particularly tsunami deposits, but also turbidites, landslide deposits, and vertical

coseismic displacement indicators). The possibility of tectonic tsunamis in the St. Law-

rence estuary also merits further study.

Fig. 11 Relative contribution of all considered source types to the cumulative probabilistic tsunami hazard(1.5 and 3 m run-up: best estimate) of each hazard zone (zones in Fig. 2), shown in order of decreasing1.5 m hazard. Note: different scale for upper versus lower panels. Data details (and a more complete versionof this figure, with individual sources) in Online Resource 4. Continental slope landslide sources are notincluded for Pacific zones; compared to most other sources, they are expected to contribute a negligiblehazard, and triggering is most likely from large near-field megathrust tsunamigenic earthquakes (i.e. not anindependent source)

266 Nat Hazards (2014) 70:237–274

123

Far-field subduction zones are potentially significant sources of tsunami hazard on the

Atlantic coast (Fig. 11). Events similar to the 1755 Lisbon tsunami have a relatively well-

constrained frequency. Nova Scotia may be sheltered from this source (Fig. 5), but a

greater hazard may be posed by the north-eastern Caribbean, where further research and

modelling is required to constrain potential megathrust events and their tsunami impact.

For the Arctic coasts of Yukon and Northwest Territories, the hypothetical Mackenzie

thrust may present the highest tsunami hazard, but at present, we consider the uncertainties

too great to include this source in the national hazard map. Future studies need to constrain

the parameters and tsunamigenic potential of this fault and to search for tsunami deposits.

Our analysis of Atlantic and Arctic continental slope failures similar to the 1929 Grand

Banks event involves a number of simplistic assumptions that should be tested with

modelling and additional data. These include: (1) the recurrence relation of failures in

Orphan Basin is applicable to other parts of the margin; (2) the threshold volume of

damaging tsunamigenic failures is 40 km3, and this value applies for the whole Atlantic

(and Arctic) margin; (3) most sites are at risk from only one source area directly offshore.

The continental margin is far from uniform; future studies should model a wide range of

landslide parameters and source areas to ascertain coastal sites that may be susceptible to

damaging waves from relatively small failures.

The inner Pacific coast, especially Georgia Strait, is sheltered from most tsunami

sources. For this and many Atlantic and Arctic areas, landslide tsunamis are likely the

greatest hazard, but are difficult to quantify (Fig. 10). Probabilistic analyses will require (1)

identification of potential sources; (2) evidence for past tsunamigenic events to establish

recurrence relations and/or slope stability analyses that incorporate seismic hazard; (3)

tsunami modelling of a wide range of failures. In the Arctic, locally damaging waves are

also likely from glacial sources, a hazard that may increase with climate warming. Tsunami

modelling should take sea ice into account; it may increase the hazard locally but decrease

it further afield; seasonal hazard variations are likely.

Acknowledgments We acknowledge helpful discussions and input from John Adams, Jan Bednarski,Calvin Campbell, Josef Cherniawsky, Scott Dallimore, Isaac Fine, Phil Hill, Sabine Hippchen, Roy Hy-ndman, Carmel Lowe, Brian MacLean, David Mazzucchi, David Mosher, David Piper, Alexander Rabi-novich, Alan Ruffman, Franck Saint-Ange, Fred Stephenson, and Kelin Wang. The manuscript wasimproved by useful comments from three anonymous reviewers. Several figures were prepared with the aidof GMT software (Wessel and Smith 1995). This is Earth Sciences Sector contribution 20130130.

References

Abe K (1995) Estimate of tsunami run-up heights from earthquake magnitudes. In: Tsuchiya Y, Shuto N(eds) Tsunami: progress in prediction, disaster prevention and warning. Advances in technologicalhazards research, vol 4. Kluwer, The Netherlands, pp 21–35

Aksu AE, Piper DJW (1987) Late Quaternary sedimentation in Baffin Bay. Can J Earth Sci24(9):1833–1846

Amundson JM, Truffer M, Luthi MP, Fahnestock M, West M, Motyka RJ (2008) Glacier, fjord, and seismicresponse to recent large calving events, Jakobshavn Isbræ, Greenland. Geophys Res Lett35(22):L22501. doi:10.1029/2008GL035281

Atlantic and Gulf of Mexico Tsunami Hazard Assessment Group (AGMTHAG) (2008) Evaluation oftsunami sources with the potential to impact the US Atlantic and Gulf coasts—a report to the NuclearRegulatory Commission. US Geological Survey Administrative Report, 300 pp

Atwater BF, Griggs GB (2012) Deep-sea turbidites as guides to Holocene earthquake history at the Cascadiasubduction zone—alternative views for a seismic-hazard workshop. US Geological Survey Open-FileReport 2012-1043, 58 pp

Nat Hazards (2014) 70:237–274 267

123

Atwater BF, Tuttle MP, Schweig ES, Rubin CM, Yamaguchi DK, Hemphill-Haley E (2004) Earthquakerecurrence inferred from paleoseismology. In: Gillespie AR, Porter SC, Atwater BF (eds) The qua-ternary period in the United States. Elsevier Science, New York, pp 331–350

Atwater BF, Musumi-Rokkaku S, Satake K, Tsuji Y, Ueda K, Yamaguchi DK (2005) The orphan tsunami of1700—Japanese clues to a parent earthquake in North America. US Geological Survey ProfessionalPaper 1707, 144 pp

Atwater BF, ten Brink US, Buckley M, Halley RS, Jaffe BE, Lopez-Venegas AM, Reinhardt EG, Tuttle MP,Watt W, Wei Y (2010) Geomorphic and stratigraphic evidence for an unusual tsunami or storm a fewcenturies ago at Anegada, British Virgin Islands. Nat Hazards. doi:10.1007/s11069-010-9622-6

Baptista MA, Miranda JM (2009) Revision of the Portuguese catalog of tsunamis. Nat Hazards Earth SystSci 9(1):25–42

Baptista MA, Miranda JM, Chierici F, Zitellini N (2003) New study of the 1755 earthquake source based onmulti-channel seismic survey data and tsunami modeling. Nat Hazards Earth Syst Sci 3(5):333–340

Barkan R, ten Brink U, Lin J (2009) Far field tsunami simulations of the 1755 Lisbon earthquake: impli-cations for tsunami hazard to the US east coast and the Caribbean. Mar Geol 264(1–2):109–122

Benetti S (2006) Late Quaternary sedimentary processes along the western North Atlantic margin. Ph.D.thesis, University of Southampton, U.K. 188 p

Bent AL (2002) The 1933 Ms = 7.3 Baffin Bay earthquake: strike-slip faulting along the northeasternCanadian passive margin. Geophys J Int 150:724–736

Bobrowsky PT, Dominguez MJ (2012) Landslide susceptibility map of Canada. Geological Survey ofCanada, Open File 7228, scale 1:6 million. doi:10.4095/291902

Bondevik S, Løvholt F, Harbitz C, Mangerud J, Dawson A, Svendsen JI (2005) The Storegga Slidetsunami—comparing field observations with numerical simulations. Mar Pet Geol 22(1–2):195–208

Bornhold BD, Harper JR, McLaren D, Thomson RE (2007) Destruction of the First Nations village ofKwalate by a rock avalanche-generated tsunami. Atmos Ocean 45(2):123–128

Braunmiller J, Nabelek J (2002) Seismotectonics of the explorer region. J Geophys Res 107(B10):2208.doi:10.1029/2001JB000220

Bryn P, Berg K, Forsberg CF, Solheim A, Kvalstad TJ (2005) Explaining the storegga slide. Mar Pet Geol22:11–19

Buckley ML, Wei Y, Jaffe BE, Watt SG (2011) Inverse modeling of velocities and inferred cause ofoverwash that emplaced inland fields of boulders at Anegada, British Virgin Islands. Nat Hazards.doi:10.1007/s11069-011-9725-8

Bustin AMM (2006) The crustal structure, deformation from GPS, and seismicity related to oblique con-vergence along the Queen Charlotte margin, British Columbia. Ph.D. thesis, University of Victoria,Canada. 226 pp

Bustin AMM, Hyndman RD, Kao H, Cassidy JF (2007) Evidence for underthrusting beneath the QueenCharlotte margin, from teleseismic receiver function analysis. Geophys J Int 171:1198–1211

Calais E, Mazabraud Y, Mercier de Lepinay B, Mann P, Mattioli G, Jansma P (2002) Strain partitioning andfault slip rates in the northeastern Caribbean from GPS measurements. Geophys Res Lett 106:1–9

Cassidy JF, Ellis RM, Karavas C, Rogers GC (1998) The northern limit of the subducted Juan de Fuca platesystem. J Geophys Res 103(B11):26949–26961

Cauchon-Voyer G, Locat J, Leroueil S, St-Onge G, Demers D (2011) Large-scale subaerial and submarineHolocene and recent mass movements in the Betsiamites area, Quebec, Canada. Eng Geol 121:28–45

Chaytor JD, ten Brink US, Solow AR, Andrews BD (2009) Size distribution of submarine landslides alongthe US Atlantic margin. Mar Geol 264(1–2):16–27

Cherniawsky JY, Titov VV, Wang K, Li J-Y (2007) Numerical simulations of tsunami waves and currentsfor southern Vancouver Island from a Cascadia megathrust earthquake. Pure Appl Geophys164(2–3):465–492

Clague JJ, Bobrowsky PT, Hutchinson I (2000) A review of geological records of large tsunamis atVancouver Island, British Columbia, and implications for hazard. Quat Sci Rev 19(9):849–863

Davis EE, Hyndman RD (1989) Accretion and recent deformation of sediments along the northern Cascadiasubduction zone. Geol Soc Am Bull 101(11):1465–1480

Day S, Kilburn C, McGuire B (2008) Tectonic threats in the Caribbean. Issues in Risk Science 8, BenfieldUCL Hazard Research Centre, London, 32 p

DeDontney N, Rice JR (2011) Tsunami wave analysis and possibility of splay fault rupture during the 2004Indian Ocean earthquake. Pure Appl Geophys. doi:10.1007/s00024-011-0438-4

DeMets C, Gordon RG, Argus DF (2010) Geologically current plate motions. Geophys J Int 181:1–80Dengler LA, Magoon OT (2006) Reassessing Crescent City, California’s tsunami risk. Proceedings of the

8th US national conference on earthquake engineering, San Francisco, California, Paper 1451, 10 pp

268 Nat Hazards (2014) 70:237–274

123

Dineva S, Eaton D, Mereu R (2004) Seismicity of the southern Great Lakes: revised earthquake hypocentersand possible tectonic controls. Bull Seismol Soc Am 94(5):1902–1918

Doser DI, Rodriguez CM, Flores C (2005) Historical earthquakes of the Puerto Rico-Virgin Islands region(1915–1963). In: Mann P (ed), Active tectonics and seismic hazards of Puerto Rico, the Virgin Islands,and offshore areas. Geol Soc Am Spec Pap 385:103–114

Ebel JE (2011) A new analysis of the magnitude of the February 1663 earthquake at Charlevoix, Quebec.Bull Seismol Soc Am 101(3):1024–1038

El-Sabh MI, Murty TS, Dumais J-F (1988) Tsunami hazards in the St. Lawrence Estuary, Canada. In: El-Sabh MI, Murty TS (eds) Natural and man-made hazards. D Reidel Publishing, The Netherlands,pp 201–213

Evans SG (2001) Landslides. In: Brooks GR (ed) A synthesis of geological hazards in Canada. Geol SurvCan Bulletin 548, pp 43–79

Fernandes RMS, Ambrosius BAC, Noomen R, Bastos L, Wortel MJR, Spakman W, Govers R (2003) Therelative motion between Africa and Eurasia as derived from ITRF2000 and GPS data. Geophys ResLett 30(16):1828. doi:10.1029/2003GL017089

Fernandes RMS, Miranda JM, Meijninger BML, Bos MS, Noomen R, Bastos L, Ambrosius BAC, RivaREM (2007) Surface velocity field of the Ibero-Maghrebian segment of the Eurasia-Nubia plateboundary. Geophys J Int 169:315–324

Fine IV, Rabinovich AB, Bornhold BD, Thomson RE, Kulikov EA (2005) The Grand Banks landslide-generated tsunami of November 18, 1929: preliminary analysis and numerical modeling. Mar Geol215(1–2):45–57

Fine IV, Cherniawsky JY, Rabinovich AB, Stephenson F (2009) Numerical modeling and observations oftsunami waves in Alberni Inlet and Barkley Sound, British Columbia. Pure Appl Geophys165:2019–2044

Frankel AD (2011) Summary of November 2010 meeting to evaluate turbidite data for constraining therecurrence parameters of great Cascadia earthquakes for the update of the national seismic hazardmaps. US Geological Survey Open-File Report 2011–1310. 13 pp

Geist EL (2005) Local tsunami hazards in the Pacific Northwest from Cascadia subduction zone earth-quakes. US Geological Survey Professional Paper 1661-B, 17 pp

Geist EL, Parsons T (2006) Probabilistic analysis of tsunami hazards. Nat Hazards 37:277–314Geist EL, Parsons T (2009) Assessment of source probabilities for potential tsunamis affecting the US

Atlantic coast. Mar Geol 264(1–2):98–108Geist EL, Parsons T, ten Brink US, Lee HJ (2009) Chapter 4: Tsunami probability. In: Bernard EN,

Robinson AR (eds) The sea, vol 15: tsunamis. Harvard University Press, Cambridge, pp 93–136Gisler G, Weaver R, Gittings ML (2006) SAGE calculations of the tsunami threat from La Palma. Sci

Tsunami Hazards 24(4):288–301Gisler G, Weaver R, Gittings M (2011) Calculations of asteroid impacts into deep and shallow water. Pure

Appl Geophys 168:1187–1198Goldfinger C, Kulm LD, McNeill LC, Watts P (2000) Super-scale failure of the southern Oregon Cascadia

margin. Pure Appl Geophys 157(6–8):1189–1226Goldfinger C, Nelson CH, Johnson JE, the Shipboard Scientific Party (2003) Holocene earthquake records

from the Cascadia subduction zone and northern San Andreas Fault based on precise dating of offshoreturbidites. Annu Rev Earth Planet Sci 31:555–577

Goldfinger C, Nelson CH, Morey AE et al (2012) Turbidite event history: methods and implications forHolocene paleoseismicity of the Cascadia subduction zone. US Geological Survey Professional Paper1661-F, 170 pp

Gracia E, Vizcaino A, Escutia C, Asioli A, Rodes A, Garcia-Orellano J, Lebreiro S, Goldfinger C (2010)Holocene earthquake record offshore Portugal (SW Iberia): testing turbidite paleoseismology in aslow-convergence margin. Quat Sci Rev 29:1156–1172

Grantz A, Phillips RL, Mullen MW, Starratt SW, Jones GA, Sathy Naidu A, Finney BP (1996) Character,paleoenvironment, rate of accumulation, and evidence for seismic triggering of Holocene turbidites,Canada Abyssal Plain, Arctic Ocean. Mar Geol 133:51–73

Gutscher M-A, Westbrook GK (2009) Great earthquakes in slow-subduction, low-taper margins. In: Lal-lemand S, Funicello F (eds) Subduction zone geodynamics. Springer, Berlin, pp 119–133

Gutscher M, Dominguez S, Westbrook GK, Leroy P (2009) Deep structure, recent deformation and analogmodeling of the Gulf of Cadiz accretionary wedge: implications for the 1755 Lisbon earthquake.Tectonophysics 475:85–97

Gutscher M-A, Westbrook GK, Marcaillou B, Graindorge D, Gailler A, Pichot T, Maury RC (2010) Alongstrike variations in the width of the seismogenic zone of the Lesser Antilles subduction predicted bythermal modeling. Geophys Res Abstr 12(EGU2010):10600

Nat Hazards (2014) 70:237–274 269

123

Halchuk S, Adams J (2008) Fourth generation seismic hazard maps of Canada: maps and grid values to beused with the 2005 National Building Code of Canada. 1 CD-ROM

Hasegawa HS, Chou CW, Basham PW (1979) Seismotectonics of the Beaufort Sea. Can J Earth Sci16(4):816–830

Hayward N, Nedimovic ClearyM, Calvert AJ (2006) Structural variation along the Devil’s Mountain faultzone, northwestern Washington. Can J Earth Sci 43:433–446

Heezen BC, Ewing M (1952) Turbidity currents and submarine slumps, and the 1929 Grand Banksearthquake. Am J Sci 250(12):849–873

Heuret A, Lallemand S, Funiciello F, Piromallo C, Faccenna C (2011) Physical characteristics of subductioninterface type seismogenic zones revisited. Geochem Geophys Geosyst 12(1). doi:10.1029/2010GC003230

Hill PR, Moran KM, Blasco SM (1982) Creep deformation of slope sediments in the Canadian Beaufort Sea.Geo-Mar Lett 2(3–4):163–170

Huppertz TJ, Piper DJW (2009) The influence of shelf-crossing glaciation on continental slope sedimen-tation, Flemish Pass, eastern Canadian continental margin. Mar Geol 265:67–85

Hyndman RD, Ellis RM (1981) Queen Charlotte fault zone: microearthquakes from a temporary array ofland stations and ocean bottom seismographs. Can J Earth Sci 18(4):776–788

Hyndman RD, Mazzotti S, Weichert D, Rogers GC (2003) Frequency of large crustal earthquakes in PugetSound—southern Georgia Strait predicted from geodetic and geological deformation rates. J GeophysRes 108(B1):2033. doi:10.1029/2001JB001710

Hyndman RD, Cassidy JF, Adams J, Rogers GC, Mazzotti S (2005) Earthquakes and seismic hazard in theYukon-Beaufort-Mackenzie. Can Soc Explor Geophys Rec, pp 32–66

James T, Rogers G, Cassidy J, Dragert H, Hyndman R, Leonard L, Nykolaishen L, Riedel M, Schmidt M,Wang K (2013) Field studies target 2012 Haida Gwaii earthquake. Eos Trans AGU 94(22):197–198

Jansma PE, Mattioli GS, Lopez A, DeMets C, Dixon TH, Mann P, Calais E (2000) Neotectonics of PuertoRico and the Virgin Islands, northeastern Caribbean, from GPS geodesy. Tectonics 19(6):1021–1037

Johnson SY, Dadisman SV, Mosher DC, Blakely RJ, Childs JR (2001) Active tectonics of the DevilsMountain fault and related structures, northern Puget Lowland and eastern Strait of Juan de Fucaregion, Pacific Northwest. US Geological Survey Professional Paper 1643, 65 pp

Kao H, Shan S-J, Dragert H, Rogers G (2009) Northern Cascadia episodic tremor and slip: a decade oftremor observations from 1997 to 2007. J Geophys Res 114:B00A12. doi:10.1029/2008JB006046

Kayen RE, Lee HJ (1991) Pleistocene slope instability of gas hydrate-laden sediment on the Beaufort seamargin. Mar Georesour Geotechnol 10(1–2):125–141

Knight B (2006) Model prediction of Gulf and southern Atlantic coast tsunami impacts from a distributionof sources. Sci Tsunami Hazards 24(5):304–312

Korup O, Clague JJ (2009) Natural hazards, extreme events, and mountain topography. Quat Sci Rev28(11–12):977–990

Koshimura S, Mofjeld HO, Gonzalez FI, Moore AL (2002) Modeling the 1100 bp paleotsunami in PugetSound, Washington. Geophys Res Lett 29(20):1948. doi:10.1029/2002GL015170

Kulikov EA, Rabinovich AB, Thomson RE, Bornhold BD (1996) The landslide tsunami of November 3,1994, Skagway Harbor, Alaska. J Geophys Res 101(C3):6609–6615

LaForge RC, McCann WR (2005) A seismic source model for Puerto Rico, for use in probabilistic groundmotion hazard analyses. In: Mann P (ed) Active tectonics and seismic hazards of Puerto Rico, theVirgin Islands, and offshore areas. Geol Soc Am Spec Pap 385:223–248

Lander JF, Whiteside LS, Lockridge PA (2002) A brief history of tsunamis in the Caribbean Sea. SciTsunami Hazards 20(2):57–94

Lane L (2002) Tectonic evolution of the Canadian Beaufort Sea—Mackenzie Delta region’’ a brief review.Can Soc Explor Geophys Rec, pp 49–56

Lay T, Ye L, Kanamori H, Yamazaki Y, Cheung KF, Kwong K, Koper KD (2013) The October 28, 2012Mw 7.8 Haida Gwaii underthrusting earthquake and tsunami: slip partitioning alonf the QueenCharlotte fault transpressional boundary. Earth Planet Sci Lett. doi:10.1016/j.epsl.2013.05.005

Leonard LJ, Hyndman RD, Mazzotti S (2004) Coseismic subsidence in the 1700 great Cascadia earthquake:coastal estimates versus elastic dislocation models. Geol Soc Am Bull 116:655–670

Leonard LJ, Hyndman RD, Mazzotti S, Nykolaishen L, Schmidt M, Hippchen S (2007) Current deformationin the northern Canadian Cordillera inferred from GPS measurements. J Geophys Res 112(B11401).doi:10.1029/2007JB005061

Leonard LJ, Rogers GC, Hyndman RD (2010) Annotated bibliography of references relevant to tsunamihazard in Canada. Geological Survey of Canada, Open File 6552, 269 pp

Leonard LJ, Rogers GC, Mazzotti S (2012a) A preliminary tsunami hazard assessment of the Canadiancoastline. Geological Survey of Canada, Open File 7201, 126 pp. doi:10.4095/292067

270 Nat Hazards (2014) 70:237–274

123

Leonard LJ, Bednarski J, Fine I, Cherniawsky J, Wright C (2012b) The Haida Gwaii tsunami of October 27,2012. Risky Ground: Centre for Natural Hazard Research newsletter, Simon Fraser Univ Winteredition: 10–11

Lewis T, Francus P, Bradley RS (2007) Limnology, sedimentology, and hydrology of a jokulhlaup into ameromictic High Arctic lake. Can J Earth Sci 44:791–806

Li G, Piper DJW, Campbell DC (2010) The Quaternary Lancaster Sound trough-mouth fan, NW Baffin Bay.In: Atlantic Geoscience Society Abstracts, Atl Geol 46:59

Locat J, Martin F, Levesque C, Locat P, Leroueil S, Konrad JM, Urgeles R, Canals M, Duchesne MJ (2003)Submarine mass movements in the upper Saguenay Fjord, (Quebec, Canada), triggered by the 1663earthquake. In: Locat J, Mienert J (eds) Submarine mass movements and their consequences. Advancesin Natural and Technological Hazards Research, vol 19. Kluwer, The Netherlands, pp 509–519

Lopez C, Spence G, Hyndman R, Kelley D (2010) Frontal ridge slope failure at the northern Cascadiamargin: margin-normal fault and gas hydrate control. Geology 38(11):967–970

Ludwin RS, Dennis R, Carver D, McMillan AD, Losey R, Clague J, Jonientz-Trisler C, Bowechop J, WrayJ, James K (2005) Dating the 1700 Cascadia earthquake: great coastal earthquakes in Native stories.Seismol Res Lett 76(2):140–148

Luque L, Lario J, Civis J, Silva PG, Zazo C, Goy JL, Dabrio CJ (2002) Sedimentary record of a tsunamiduring Roman times, Bay of Cadiz, Spain. J Quat Sci 17(5–6):623–631

Macayeal DR, Abbot DS, Sergienko OV (2011) Iceberg-capsize tsunamigenesis. Ann Glaciol 52(58):51–56Mader C (2001) Modeling the La Palma landslide tsunami. Sci Tsunami Hazards 19(3):150–170Masson DG, Watts AB, Gee MJR, Urgeles R, Mitchell NC, Le Bas TP, Canals M (2002) Slope failures on

the flanks of the western Canary Islands. Earth-Sci Rev 57(1–2):1–35Mazzotti S, Hyndman RD, Fluck P, Smith AJ, Schmidt M (2003a) Distribution of the Pacific/North America

motion in the Queen Charlotte Islands-S. Alaska plate boundary zone. Geophys Res Lett 30(14):1762.doi:10.1029/2003GL017586

Mazzotti S, Dragert H, Henton J, Schmidt M, Hyndman R, James T, Lu Y, Craymer M (2003b) Currenttectonics of northern Cascadia from a decade of GPS measurements. J Geophys Res 108(B12):2554.doi:10.1029/2003JB002653

Mazzotti S, James TS, Henton J, Adams J (2005) GPS crustal strain, postglacial rebound, and seismic hazardin eastern North America: the Saint Lawrence valley example. J Geophys Res 110(B11301). doi:10.1029/2004JB003590

McCann WR (1985) On the earthquake hazards of Puerto Rico and the Virgin Islands. Bull Seismol Soc Am75(1):251–262

McMurtry GM, Watts P, Fryer GJ, Smith JR, Imamura F (2004) Giant landslides, mega-tsunamis, and paleo-sea level in the Hawaiian Islands. Mar Geol 203:219–233

Miller DJ (1960) Giant waves in Lituya Bay, Alaska. US Geological Survey Professional Paper 354-C,pp 51–86

Moore JG, Bryan WB, Ludwig KR (1994) Chaotic deposition by a giant wave, Molokai, Hawaii. Geol SocAm Bull 106(7):962–967

Mosher DC (2009) Submarine landslides and consequent tsunamis in Canada. Geosci Can 36(4):179–190Mosher DC, Piper DJW (2007) Analysis of multibeam seafloor imagery of the Laurentian Fan and the 1929

Grand Banks landslide area. In: Lykousis V, Sakellariou D, Locat J (eds) Submarine mass movementsand their consequences. Advances in Natural and Technological Hazards Research, vol 27. Springer,The Netherlands, pp 77–88

Mosher DC, Moran KM, Hiscott RN (1994) Late Quaternary sediment, sediment mass flow processes andslope instability on the Scotian Slope. Sedimentology 41(5):1039–1061

Mosher DC, Cassidy JF, Lowe C, Mi Y, Hyndman RD, Rogers GC, Fisher M (2000) Neotectonics in theStrait of Georgia: first tentative correlation of seismicity with shallow geological structure in south-western British Columbia. In: Current Res, Geological Survey of Canada, paper 2000-A22, 9 pp

Mosher DC, Monahan PA, Barrie JV, Courtney RC (2004) Coastal submarine failures in the Strait ofGeorgia, British Columbia: landslides of the 1946 Vancouver Island earthquake. J Coast Res20(1):277–291

Mosher DC, Xu Z, Shimeld J (2010) The Pliocene Shelburne mass-movement and consequent tsunami,western Scotian Slope. In: Mosher DC, Shipp RC, Moscardelli L, Chaytor JD, Baxter CDP, Lee HJ,Urgeles R (eds) Submarine mass movements and their consequences. Advances in Natural andTechnological Hazards Research, vol 28. Springer, The Netherlands, pp 765–775

Mosher DC, Shimeld J, Hutchinson D, Lebedeva-Ivanova N, Chapman CB (2012) Submarine landslides inArctic sedimentation: Canada Basin. In: Yamada Y, Kawamura K, Ikehara K, Ogawa Y, Urgeles R,Mosher D, Chaytor J, Strasser M (eds) Submarine mass movements and their consequences. Advancesin Natural and Technological Hazards Research, vol 31. Springer, The Netherlands, pp 147–157

Nat Hazards (2014) 70:237–274 271

123

Muir-Wood R, Mignan A (2009) A phenomenological reconstruction of the Mw 9 November 1st 1755earthquake source. In: Mendes-Victor LA, Oliviera CS, Azebedo J, Ribeiro A (eds) The 1755 Lisbonearthquake: revisited. Geotechnical, geological, and earthquake engineering 7. Springer, Berlin,pp 121–146

Pararas-Carayannis G (2002) Evaluation of the threat of mega tsunami generation from postulated massiveslope failures of island stratovolcanoes on La Palma, Canary Islands, and on the island of Hawaii. SciTsunami Hazards 20(5):251–277

Peterson CD, Cruikshank KM, Darienzo ME, Wessen GC, Butler VL, Sterling SL (2013a) Coseismicsubsidence and paleotsunami run-up records from latest Holocene deposits in the Waatch Valley, NeahBay, Northwest Washington, USA: links to great earthquakes in the northern Cascadia margin. J CoastRes 29(1):157–172

Peterson CD, Clague JJ, Carver GA, Cruikshank KM (2013b) Recurrence intervals of major paleotsunamisas calibrated by historic tsunami deposits in three localities: port Alberni, Cannon Beach, and CrescentCity, along the Cascadia margin, Canada and USA. Nat Hazards. doi:10.1007/s11069-013-062-1

Pilkey OH (1988) Basin plains; giant sedimentation events. Geol Soc Am Spec Pap 229:93–99Piper DJW (2005) Late Cenozoic evolution of the continental margin of eastern Canada. Nor J Geol

85:305–318Piper DJW (2007) Labrador geological framework report to C-Core. unpublished report, 9 ppPiper DJW, Campbell DC (2005) Quaternary geology of Flemish Pass and is application to geohazard

evaluation for hydrocarbon development. Pet Resour Reserv Spec Pap 43:29–43Piper DJW, McCall C (2003) A synthesis of the distribution of submarine mass movements on the eastern

Canadian margin. In: Locat J, Mienert J (eds) Submarine mass movements and their consequences.Advances in Natural and Technological Hazards Research, vol 19. Kluwer, The Netherlands,pp 291–298

Piper DJW, Mosher DC, Gauley B-J, Jenner K, Campbell DC (2003) The chronology and recurrence ofsubmarine mass movements on the continental slope off southeastern Canada. In: Locat J, Mienert J(eds) Submarine mass movements and their consequences. Advances in Natural and TechnologicalHazards Research, vol 19. Kluwer, The Netherlands, pp 299–306

Piper DJW, Tripsanas E, Mosher DC, MacKillop K (2011) Paleoseismicity of the continental margin ofeastern Canada: rare regional failures and associated turbidites in Orphan Basin. Unpublished manu-script, 38 pp

Poncet R, Campbell C, Dias F, Locat J, Mosher D (2010) A study of the tsunami effects of two landslides inthe St. Lawrence estuary. In: Mosher DC, Shipp RC, Moscardelli L, Chaytor JD, Baxter CDP, Lee HJ,Urgeles R (eds) Submarine mass movements and their consequences. Advances in Natural andTechnological Hazards Research, vol 28. Springer, The Netherlands, pp 755–764

Priest GR, Goldfinger C, Wang K, Witter RC, Zhang Y, Baptista AM (2009) Confidence limits for tsunami-inundation limits in northern Oregon inferred from a 10,000-year history of great earthquakes at theCascadia subduction zone. Nat Hazards. doi:10.1007/s11069-009-9453-5

Rabinovich AB, Thomson RE, Bornhold BD, Fine IV, Kulikov EA (2003) Numerical modelling of tsunamisgenerated by hypothetical landslides in the Strait of Georgia, British Columbia. Pure Appl Geophys160(7):1273–1313

Rabinovich AB, Thomson RE, Titov VV, Stephenson FE, Rogers GC (2008) Locally generated tsunamisrecorded on the coast of British Columbia. Atmos Ocean 46(3):343–360

Reimnitz E, Maurer DK (1978) Storm surges in the Alaskan Beaufort Sea. US Geological Survey Open-FileReport 78–593, 26 pp

Ristau J, Rogers GC, Cassidy JF (2007) Stress in western Canada from regional moment tensor analysis.Can J Earth Sci 44:127–148

Roger J, Baptista MA, Mosher D, Hebert H, Sahal A (2010) Tsunami impact on Newfoundland, Canada, dueto far-field generated tsunamis: implications on hazard assessment. In: Proceedings of the 9th USnational and 10th Canadian conference on earthquake engineering. Toronto, July 25–29, Paper 1837,6 pp

Roger J, Baptista MA, Sahal A, Accary F, Allgeyer S, Hebert H (2011) The transoceanic 1755 Lisbontsunami in Martinique. Pure Appl Geophys 168(6–7):1015–1031

Rohr KMM, Furlong KP (1995) Ephemeral plate tectonics at the Queen Charlotte triple junction. Geology23(11):1035–1038

Rohr KMM, Scheidhauer M, Trehu AM (2000) Transpression between two warm mafic plates: the QueenCharlotte fault revisited. J Geophys Res 105(B4):8147–8172

Ruffman A (2006) Documentation of the farfield parameters of the November 1, 1755 ‘‘Lisbon’’ tsunamialong the shores of the western Atlantic Ocean. In: Program and Abstracts, International TsunamiSociety Third Tsunami Symposium. Honolulu, HI, May 23–25

272 Nat Hazards (2014) 70:237–274

123

Ruffman A, Murty T (2006) Tsunami hazards in the Arctic regions of North America, Greenland and theNorwegian Sea. In: Program and Abstracts, International Tsunami Society Third Tsunami Symposium.Honolulu, HI, May 23–25

Satake K, Wang K, Atwater BF (2003) Fault slip and seismic moment of the 1700 Cascadia earthquakeinferred from Japanese tsunami descriptions. J Geophys Res 108(B11):2535. doi:10.1029/2003JB002521

Scheffers A, Scheffers S, Kelletat D (2005) Paleo-tsunami relics on the southern and central Antillean Islandarc. J Coast Res 21(2):263–273

Shanmugam G (2011) Process-sedimentological challenges in distinguishing paleo-tsunami deposits. NatHazards. doi:10.1007/s11069-011-9766-z

Shennan I, Bruhn R, Plafker G (2009) Multi-segment earthquakes and tsunami potential of the Aleutianmegathrust. Quat Sci Rev 28(1–2):7–13

Sherrod BL, Mazzotti S, Haugerud R (2008) Comparison of geodetic and paleoseismic rates of deformationin the Puget Sound-Georgia Basin, Pacific Northwest. Eos Trans Am Geophys Union 89(53):AbstractT21B-1953

Skvortsov A, Bornhold B (2007) Numerical simulation of the landslide-generated tsunami in Kitimat Arm,British Columbia, Canada, 27 April 1975. J Geophys Res 112(F02028). doi:10.1029/2006JF000499

Smith AJ, Hyndman RD, Cassidy JF, Wang K (2003) Structure, seismicity, and thermal regime of the QueenCharlotte transform margin. J Geophys Res 108(B11):2539. doi:10.1029/2002JB002247

Solares JMM, Arroyo AL (2004) The great historical 1755 earthquake: effects and damage in Spain.J Seismol 8(2):275–294

Solheim A, Berg K, Forsberg CF, Bryn P (2005) The Storegga Slide complex: repetitive large scale slidingwith similar cause and development. Mar Pet Geol 22:97–107

Spaeth MG, Berkman SC (1967) The tsunami of March 28, 1964, as recorded at tide stations. Coast andGeodetic Survey Technical Bulletin 33, US Department of Commerce, 86 pp

Stephenson FE, Rabinovich AB (2009) Tsunamis on the Pacific coast of Canada recorded in 1994–2007.Pure Appl Geophys 166(1–2):177–210

Stephenson F, Rabinovich AB, Solovieva ON, Kulikov EA, Yakovenko OI (2007) Catalogue of tsunamis,British Columbia, Canada: 1700–2007. P.P. Shirshov Institute of Oceanology, Moscow

Stich D, Mancilla F, Pondrelli S, Morales J (2007) Source analysis of the February 12th 2007, Mw 6.0Horseshoe earthquake: implications for the 1755 Lisbon earthquake. Geophys Res Lett 34:L12308.doi:10.1029/2007GL030012

Suleimani E, Nicolsky DJ, Haeussler PJ, Hansen R (2011) Combined effects of tectonic and landslide-generated tsunami runup at Seward, Alaska during the Mw 9.2 1964 earthquake. Pure Appl Geophys168:1053–1074

Syvitski JPM, Burrell DC, Skei JM (1987) Chapter 5: Subaqueous slope failure. In: Fjords: processes andproducts. Springer, New York, pp 175–209

Szczucinski W (2012) The post-depositional changes of the onshore 2004 tsunami deposits on the AndamanSea coast of Thailand. Nat Hazards 60(1):115–133

Taggart BE, Lundberg J, Carew JL, Mylroie JE (1993) Holocene reef-rock boulders on Isla de Mona, PuertoRico transported by a hurricane or seismic sea wave. Geol Soc Am Abstr with Programs 25(6):A–A61

Tappin DR, McNeil LC, Henstock T, Mosher D (2007) Mass wasting processes–offshore Sumatra. In:Lykousis V, Sakellariou D, Locat J (eds) Submarine mass movements and their consequences.Advances in Natural and Technological Hazards Research, vol 27. Springer, The Netherlands,pp 327–336

ten Brink US, Lopez-Venegas AM (2012) Plate interaction in the NE Caribbean subduction zone fromcontinuous GPS observations. Geophys Res Lett 39:L10304. doi:10.1029/2012GL051485

ten Brink US, Geist EL, Lynett P, Andrews, B (2006) Submarine slides north of Puerto Rick and theirtsunami hazard. In: Mercado-Irizarry, A, Liu P (eds) Caribbean Tsunami Hazard, Proceedings of theNSF Caribbean tsunami workshop. World Scientific Publishing Co. Pte. Ltd., pp 67–90

ten Brink US, Lee HJ, Geist EL, Twichell D (2009) Assessment of tsunami hazard to the US east coast usingrelationships between submarine landslides and earthquakes. Mar Geol 264(1–2):65–73

Thiebot E, Gutscher M-A (2006) The Gibraltar Arc seismogenic zone (part 1): constraints on a shallow eastdipping fault plane source for the 1755 Lisbon earthquake provided by seismic data, gravity andthermal modeling. Tectonophysics 426:135–152

Thomson RE, Rabinovich AB, Krassovski MV (2007) Double jeopardy: concurrent arrival of the 2004Sumatra tsunami and storm-generated waves on the Atlantic coast of the United States and Canada.Geophys Res Lett 34:L15607. doi:10.1029/2007GL030685

Tocque P (1846) Wandering thoughts, or solitary hours. Thomas Richardson and Son, London

Nat Hazards (2014) 70:237–274 273

123

van Zeyl DP (2009) Evaluation of subaerial landslide hazards in Knight Inlet and Howe Sound, BritishColumbia. MSc. thesis, Simon Fraser University, Vancouver, Canada. 199 pp

Vernant P, Fadil A, Mourabit T, Ouazar D, Koulali A, Davila JM, Garate J, McClusky S, Reilinger R (2010)Geodetic constraints on active tectonics of the Western Mediterranean: implications for the kinematicsand dynamics of the Nubia-Eurasia plate boundary zone. J Geodyn 49:123–129

Ward SN (2001) Landslide tsunami. J Geophys Res 106(6):11201–11215Ward SN, Day S (2001) Cumbre Vieja Volcano—potential collapse and tsunami at La Palma,Canary

Islands. Geophys Res Lett 28(17):3397–3400Ward SN, Day S (2005) Tsunami thoughts. CSEG Recorder, December 2005. pp 38–44Watts P (2004) Probabilistic predictions of landslides tsunamis off southern California. Mar Geol

203(3–4):281–301Waythomas CF, Watts P, Shi F, Kirby JT (2009) Pacific Basin tsunami hazards associated with mass flows

in the Aleutian arc of Alaska. Quat Sci Rev 28(11–12):1006–1019WCATWC (West Coast and Alaska Tsunami Warning Center) (2009a) Amplitudes for the September 29,

2009 Samoa tsunami. http://wcatwc.arh.noaa.gov/previous.events/09-29-09-Samoa/09-29-09.htmWCATWC (West Coast and Alaska Tsunami Warning Center) (2009b) Amplitudes for the October 7, 2009

Torres Islands, Vanuatu tsunami. http://wcatwc.arh.noaa.gov/previous.events/10-07-09/10-07-09.htmWessel P, Smith WHF (1995) New version of the generic mapping tools released. EOS Trans Am Geophys

Union 76(33):329. doi:10.1029/95EO00198Whitmore PM (1993) Expected tsunami amplitudes and currents along the North American coast for

Cascadia subduction zone earthquakes. Nat Hazards 8:59–73Whitmore P, Benz H, Bolton M et al (2008) NOAA/West Coast and Alaska tsunami warning center Pacific

Ocean response criteria. Sci Tsunami Hazards 27(1):1–21Wigen SO (1983) Historical study of tsunamis at Tofino, Canada. In: Iida K, Iwasaki T (eds) Tsunamis—

their science and engineering. Terra Scientific Publishing, Tokyo, pp 105–119Willoughby EC, Hyndman RD (2005) Earthquake rate, slip rate, and the effective seismic thickness for

oceanic transform faults of the Juan de Fuca plate system. Geophys J Int 160:855–868Witter RC, Zhang Y, Wang K, Priest GR, Goldfinger C, Stimely LL, English JT, Ferro PA (2011) Simu-

lating tsunami inundation at Bandon, Coos County, Oregon, using hypothetical Cascadia and Alaskaearthquake scenarios. Oregon Dept of Geology and Mineral Industries Special Paper 43, 57 pp

Wynn RB, Masson DG (2003) Canary Islands landslides and tsunami generation: can we use turbiditedeposits to interpret landslide processes. In: Locat J, Mienert J (eds) Submarine mass movements andtheir consequences. Advances in Natural and Technological Hazards Research, vol 19. Kluwer, TheNetherlands, pp 325–332

Xu Z (2007) The all-source Green’s function and its application to tsunami problems. Sci Tsunami Hazards26(1):59–69

274 Nat Hazards (2014) 70:237–274

123