Download H.Albert et al.

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Axial Seamount wikipedia , lookup

Llullaillaco wikipedia , lookup

Mono–Inyo Craters wikipedia , lookup

Itcha Range wikipedia , lookup

Licancabur wikipedia , lookup

Mauna Loa wikipedia , lookup

Level Mountain wikipedia , lookup

Mount Garibaldi wikipedia , lookup

Teide wikipedia , lookup

Mount Pleasant Caldera wikipedia , lookup

Lastarria wikipedia , lookup

Mount Edziza volcanic complex wikipedia , lookup

Mount Meager massif wikipedia , lookup

Shield volcano wikipedia , lookup

Mount Rinjani wikipedia , lookup

Surtsey wikipedia , lookup

Santorini wikipedia , lookup

Large igneous province wikipedia , lookup

Mount St. Helens wikipedia , lookup

Wells Gray-Clearwater volcanic field wikipedia , lookup

Lascar (volcano) wikipedia , lookup

Cascade Volcanoes wikipedia , lookup

Cerro Blanco (volcano) wikipedia , lookup

Volcanology of Io wikipedia , lookup

Krakatoa wikipedia , lookup

Mayon wikipedia , lookup

1257 Samalas eruption wikipedia , lookup

Mount Pinatubo wikipedia , lookup

Volcano wikipedia , lookup

Mount Etna wikipedia , lookup

Silverthrone Caldera wikipedia , lookup

Nevado del Ruiz wikipedia , lookup

Mount Vesuvius wikipedia , lookup

Mount Pelée wikipedia , lookup

Cerro Azul (Chile volcano) wikipedia , lookup

Transcript
1
Years to weeks of seismic unrest and magmatic intrusions
2
precede monogenetic eruptions
3
Helena Albert1, Fidel Costa2, and Joan Martí3
4
1
Central Geophysical Observatory, Spanish Geographic Institute (IGN), 28014, Madrid, Spain
5
2
Earth Observatory of Singapore, Nanyang Technological University, 639798, Singapore, Singapore
6
3
Institute of Earth Sciences Jaume Almera, CSIC, 08028, Barcelona, Spain
7
ABSTRACT
8
9
Seismic, deformation, and gas activity (unrest) typically precedes volcanic eruptions.
Tracking the changes of this activity with monitoring data is increasingly possible to successfully
10
forecast eruptions from stratovolcanoes. However, this is not the case for monogenetic
11
volcanoes. Eruptions from these volcanoes tend to be small but are particularly difficult to
12
anticipate since they occur at unexpected locations (e.g. Paricutin, 1943), and there is very
13
limited instrumental monitoring data. Many monogenetic volcanic fields occur in highly touristy
14
or populated areas (e.g. Canary Islands, Auckland City, Mexico City, Izu-Tobu volcanic field)
15
and thus even a small eruption can have a major economic and societal impact. We have
16
gathered the available instrumental data for unrest and combined it with new historical factual
17
accounts of seismicity. We find that there is a commonality in the seismic activity preceding
18
these eruptions, with clusters at around one or two years, two or three months, and one or two
19
weeks. The petrological and geochemical characteristics of these eruptions show that multiple
20
magma batches interacted in a subvolcanic reservoir, and multiple intrusions occurred on a
21
similar time scales to the seismicity. We propose a general model where early dike intrusions in
22
the crust do not erupt and create small plumbing systems (e.g. stalled intrusions), but they
23
probably are instrumental in creating a thermal and rheological pathway for later dikes to be able
24
to reach the surface. These observations provide a conceptual framework for better anticipating
25
monogenetic eruptions and should lead to improved strategies for mitigation of their associated
26
hazards and risks.
27
SEISMICITY ASSOCIATED WITH MONOGENETIC ERUPTIONS
28
One of the main problems to quantify the probability of eruption in a monogenetic
29
volcanic field is the lack of data. Monogenetic fields can be active during millions of year, but
30
the magmatic processes and unrest times are very short compared with the inactivity periods of a
31
given volcanic field. Many of these eruptions occurred before there was any instrumental
32
monitoring data, and our knowledge is currently based on a few factual accounts of historical
33
eruptions (Baker, 1946; Romero, 1991; De la Cruz-Reyna et al., 2011; Sánchez, 2014). We have
34
done an exhaustive revision and compilation of the unrest activity (mainly seismicity) of all the
35
historical monogenetic eruptions for which we have had access (ten eruptions in total; Table 1).
36
This includes five events in the Canary Islands, two in the Michoacan-Guanajuato (Mexico), and
37
one event for those of the Higashi-Izu (Japan), the Owen Stanley Range (Papua), and Iceland
38
(Table 1). Instrumental monitoring data are sparse and available in four cases, but only for the
39
eruption of El Hierro 2011 the data are of high quality (López et al., 2012). For the rest eruptions
40
we did a revision of the factual accounts available in historical documents (see supplemental
41
material for details). Some historical documents are very detailed and give the number of seismic
42
events and the effects on people, furniture and buildings, and sometimes an intensity value (e.g.
43
Mercalli scale). Other reports are not detailed enough to discern between intensities or establish a
44
detailed time series of the number of seismic events. It is important to note that the lack of data
45
in some periods for some eruptions could be due to the lack of historical reports, not to the lack
46
of earthquakes. The details of our analysis of seismic activity for each eruption can be found in
47
supplemental material.
48
We compared the time frames and intensity of seismic activity between different events
49
using a normalization of times and number of events (Fig 1). There are some common features
50
shown by several eruptions. We see that in some cases there are seismic crises interspersed by
51
calm periods that occur between a year to a few months prior to eruption. Many of the
52
considered monogenetic volcanoes display an exponential behaviour of the seismic activity from
53
two or three months before the eruption. These seismic crises might reflect the repetitive
54
intrusions of magma in the crust, and thus probably correspond to mid crustal stalled intrusions
55
(Moran et al., 2011). These intrusions would stall and start to create a small plumbing system.
56
The depth at which these intrusions stall is difficult to constrain but some seismic and
57
petrological data suggest 5-15 km (Klügel et al., 1997; Johnson et al., 2008; Domínguez Cerdeña
58
et al., 2014) below the volcano. Virtually all the eruptions we have studied show a sharp increase
59
in the seismic activity about two weeks to two days before eruption. This short time might be
60
related to magma migration towards the surface (Johnson et al., 2008).
61
PETROLOGICAL EVIDENCE FOR MAGMA INTERACTIONS AND THEIR TIME
62
SCALES
63
The temporal analyses of seismic activity that we report below already provide a
64
framework for anticipating the monitoring data that can be expected for monogenetic volcanoes.
65
However, it is necessary to have a conceptual model of the processes that are occurring for
66
making knowledge-based forecast. Given the large number of variables and parameters that play
67
a role for a magma being able to erupt or not it is important to have deeper knowledge of what is
68
occurring at depth. This allows adapting our interpretations of monitoring data for each case
69
when the local situations are different from one volcanic field to another.
70
A set of complementary data that is readily available for many monogenetic volcanoes and some
71
historical eruptions are the petrological and geochemical characteristic of the erupted rocks. The
72
eruptions we have studied (Table 1) and many others for which there is no knowledge of unrest
73
data show that the erupted magmas were affected by open-system processes involving multiple
74
magmas (Klügel et al., 2000; Johnson et al., 2008; Rowe et al., 2011; Martí et al., 2013; Longpré
75
et al., 2014; Albert et al., 2015). Mixing between mafic magmas has been reported in seven of
76
the ten considered cases (Table 1). In the cases of Jorullo and Paricutin, in addition to the mixing
77
between similar magmas, upper crustal assimilation has also been identified, and implies stalling
78
magma batches at crustal levels. Thus, these petrological studies also show that the monogenetic
79
eruptions are not driven simply by dikes that travel from the mantle to the surface, but support a
80
more complex scenario of magma interactions and assimilation in subvolcanic reservoirs
81
(Johnson et al., 2008; Rowe et al., 2011; Martí et al., 2013; Longpré et al., 2014; Albert et al.,
82
2015). Thus, the seismic and petrological studies imply that there is an incubation period of the
83
early intrusions at mid crustal depths before eruption.
84
Studies of the zoning pattern of crystals from stratovolcanoes have been used to provide a good
85
framework for the interpretation of unrest data in stratovolcanoes like Mt. Etna (Kahl et al.,
86
2011), Vesuvius (Morgan et al., 2006) and Mt.St Helens (Saunders et al., 2012). A few studies
87
have also been done in monogenetic volcanoes. For example, the zoning patterns of olivine
88
crystals of the Siete Fuentes, Fasnia and Arafo eruptions (Albert et al., 2015) shows that there
89
were several magma mixing events that occurred about a year, two months and two weeks before
90
eruption. Similar time frames from crystals have been found in the other studied eruptions of San
91
Juan, Jorullo and El Hierro. The time frames of crystals and seismicity are very similar (Fig 1).
92
This means that we can associate the changes in seismicity with intrusion that might erupt or not,
93
but it typically seems to involve at least two distinct periods of intrusion before eruption. Some
94
monogenetic eruptions carry mantle xenoliths which may be interpreted as direct magma transfer
95
to the surface (Bolós et al., 2015), but detailed petrological studies have shown stalling of the
96
magmas at shallower depths and thus the existence of subvolcanic reservoirs and multiple
97
intrusions (Klügel et al., 1997; Klügel et al., 2001).
98
THE MAGMA PLUMBING SYSTEM AND PROCESSES LEADING TO
99
MONOGENTIC ERUPTIONS
100
An intriguing aspect of the petrological and geochemical data for these eruptions is that
101
open system and mixing seem are prevalent and imply that magmas coming from depth
102
commonly intercepts a shallower reservoir. Although this might can be expected in areas with
103
high volcanic fluxes such as Iceland, it is not to be expected that for example in the Canaries
104
there are numerous melt lenses at shallow depth and ‘waiting to erupt’. The commonality of
105
open-system processes for many of these events may rather be indicative that the early seismicity
106
associated with intrusions are stalled intrusions. In other words, it seems that magmas coming
107
from depth in dikes are rarely able to go straight to the surface, but stall at some intermediate
108
depth (Fig 2a and b). There are many parameters that control whether mafic dikes from the
109
mantle will be able to reach the surface, including magma buoyancy, thermal survival, tectonic
110
stress or pre-existing crustal discontinuities (Rubin 1995; Valentine and Gregg, 2008; Le Corvec
111
et al., 2013; Rivalta et al., 2015). Our dataset doesn’t allow us to identify a precise control on
112
each of the eruptions, but the existence of a shallow plumbing system suggest that dikes
113
separated from their sources and travelled as small batches. Once at shallow depths, magmas can
114
cool, probably degas and evolve toward more differentiated compositions. Repetitive intrusions
115
of small magma batches on the same location are probably able to modify the thermal and
116
rheological state of the crust through which they pass and reside, and thus when renewed dike
117
intrusions from depth occur they find an increasingly easier path, and lower energy required to
118
reach the surface and erupt (Fig. 2c) (Strong and Wolff, 2003). Our conceptual model fits well
119
with our new compilation of seismic unrest and petrology of the erupted magmas and thus
120
should lead to more informed and process based anticipating of monogenetic eruption. More
121
experiments and numerical models of dike migration should be able to test the importance of
122
repetitive intrusions in allowing mafic magmas from monogenetic eruption to reach the surface.
123
124
125
ACKNOWLEDGMENTS
We are grateful for the help and discussions about seismicity and deformation with C. del
126
Fresno and L. García. T. Girona is thanked for discussion about dikes. We also thank the Teide
127
National Park for giving us permission to access the studied sites. H. Albert was funded by the
128
Spanish Geographic Institute (IGN). This research was partially funded by the Earth Observatory
129
of Singapore (EOS) “Magma Plumbing Project”.
130
131
REFERENCES CITED
132
133
134
Albert, H., Costa, F., and Martí, J., 2015, Timing of magmatic processes and unrest associated
with mafic historical monogenetic eruptions in Tenerife Island: Journal of Petrology,
accepted after major revisions.
135
136
137
Araña, V., and Ibarrola, E., 1973, Rhyolitic pumice in the basaltic pyroclasts from the 1971
eruption of Teneguía volcano, Canary Islands: Lithos, v. 6, no. 3, p. 273–278, doi:
10.1016/0024-4937(73)90088-1.
138
139
140
Baker, G., 1946, Preliminary note on volcanic eruptions in the Goropu Mountains, Southeastern
Papua, during the period December, 1943, to August, 1944: The Journal of Geology, v. 54,
no. 1, p. 19–31.
141
142
143
Bolós, X., Martí, J., Becerril, L., Planagumà, L., Grosse, P., and Barde-Cabusson, S., 2015,
Volcano-structural analysis of La Garrotxa Volcanic Field (NE Iberia): Implications for the
plumbing system: Tectonophysics, v. 642, p. 58–70, doi: 10.1016/j.tecto.2014.12.013.
144
145
Carreón Nieto, M.C., 2002, Un castigo divino: el volcán de Jorullo: Tzintzun, Revista de
Estudios Históricos, v. 35, p. 37-64.
146
147
De la Cruz-Reyna, S., and Yokoyama, I., 2011, A geophysical characterization of monogenetic
volcanism: Geofisica Internacional, v. 50, p. 465–484.
148
149
150
151
Le Corvec, N., Bebbington, M.S., Lindsay, J.M., and McGee, L.E., 2013, Age, distance, and
geochemical evolution within a monogenetic volcanic field: Analyzing patterns in the
Auckland Volcanic Field eruption sequence: Geochemistry, Geophysics, Geosystems, v. 14,
no. 9, p. 3648–3665, doi: 10.1002/ggge.20223.
152
153
154
Domínguez Cerdeña, I., del Fresno, C., and Gomis Moreno, A., 2014, Seismicity patterns prior
to the 2011 El Hierro eruption: Bulletin of the Seismological Society of America, v. 104,
no. 1, p. 567–575, doi: 10.1785/0120130200.
155
156
157
158
Higgins, M.D., and Roberge, J., 2007, Three magmatic components in the 1973 eruption of
Eldfell volcano, Iceland: Evidence from plagioclase crystal size distribution (CSD) and
geochemistry: Journal of Volcanology and Geothermal Research, v. 161, no. 3, p. 247–260,
doi: 10.1016/j.jvolgeores.2006.12.002.
159
IGN, Seismic catalogue: http://www.ign.es.
160
161
162
163
164
Johnson, E.R., Wallace, P.J., Cashman, K. V., Granados, H.D., and Kent, A.J.R., 2008,
Magmatic volatile contents and degassing-induced crystallization at Volcán Jorullo,
Mexico: Implications for melt evolution and the plumbing systems of monogenetic
volcanoes: Earth and Planetary Science Letters, v. 269, no. 3-4, p. 477–486, doi:
10.1016/j.epsl.2008.03.004.
165
166
167
168
Kahl, M., Chakraborty, S., Costa, F., and Pompilio, M., 2011, Dynamic plumbing system
beneath volcanoes revealed by kinetic modeling, and the connection to monitoring data: An
example from Mt. Etna: Earth and Planetary Science Letters, v. 308, no. 1-2, p. 11–22, doi:
10.1016/j.epsl.2011.05.008.
169
170
171
172
Klügel, A., 2001, Prolonged reactions between harzburgite xenoliths and silica-undersaturated
melt: implications for dissolution and Fe-Mg interdiffusion rates of orthopyroxene:
Contributions to Mineralogy and Petrology, v. 141, no. 1, p. 1–14, doi:
10.1007/s004100000222.
173
174
Klügel, A., Hansteen, T.H., and Schmincke, H.U., 1997, Rates of magma ascent and depths of
magma reservoirs beneath La Palma (Canary Islands): Terra Nova, v. 9, p. 117–121.
175
176
177
178
Klügel, A., Hoernle, K.A., Schmincke, H.-U., and White, J.D.L., 2000, The chemically zoned
1949 eruption on La Palma (Canary Islands): Petrologic evolution and magma supply
dynamics of a rift zone eruption: Journal of Geophysical Research, v. 105, no. B3, p. 5997–
6016.
179
180
181
Longpré, M.A., Klügel, A., Diehl, A., and Stix, J., 2014, Mixing in mantle magma reservoirs
prior to and during the 2011-2012 eruption at El Hierro, Canary Islands: Geology, v. 42, no.
4, p. 315–318, doi: 10.1130/G35165.1.
182
183
184
185
186
López, C., Blanco, M.J., Abella, R., Brenes, B., Cabrera Rodríguez, V.M., Casas, B., Domínguez
Cerdea, I., Felpeto, a., De Villalta, M.F., Del Fresno, C., García, O., García-Arias, M.J.,
García-Cañada, L., Gomis Moreno, a., et al., 2012, Monitoring the volcanic unrest of El
Hierro (Canary Islands) before the onset of the 2011-2012 submarine eruption: Geophysical
Research Letters, v. 39, no. 13, p. 1–7, doi: 10.1029/2012GL051846.
187
188
189
190
Martí, J., Castro, A., Rodríguez, C., Costa, F., Carrasquilla, S., Pedreira, R., and Bolos, X., 2013,
Correlation of magma evolution and geophysical monitoring during the 2011-2012 El
Hierro (Canary Islands) submarine eruption: Journal of Petrology, v. 54, no. 7, p. 1349–
1373, doi: 10.1093/petrology/egt014.
191
192
193
194
Mattsson, H.B., and Oskarsson, N., 2005, Petrogenesis of alkaline basalts at the tip of a
propagating rift: Evidence from the Heimaey volcanic centre, south Iceland: Journal of
Volcanology and Geothermal Research, v. 147, no. 3-4, p. 245–267, doi:
10.1016/j.jvolgeores.2005.04.004.
195
196
197
Moran, S.C., Newhall, C., and Roman, D.C., 2011, Failed magmatic eruptions: Late-stage
cessation of magma ascent: Bulletin of Volcanology, v. 73, no. 2, p. 115–122, doi:
10.1007/s00445-010-0444-x.
198
199
200
Morgan, D.J., Blake, S., Rogers, N.M., De Vivo, B., Rolandi, G., and Davidson, J.P., 2006,
Magma chamber recharge at Vesuvius in the century prior to the eruption of A.D. 79:
Geology, v. 34, no. 10, p. 845–848, doi: 10.1130/G22604.1.
201
202
203
Rivalta, E., Taisne, B., Bunger, a P., and Katz, R.F., 2015, A review of mechanical models of
dike propagation : Schools of thought , results and future directions: Tectonophysics, v. 638,
no. OCTOBER, p. 1–42, doi: 10.1016/j.tecto.2014.10.003.
204
205
Romero Ruiz, C., 1991, Las Manifestaciones Volcánicas Históricas del Archipiélago Canario:
Universidad de La Laguna and Gobierno Canarias, Santa Cruz de Tenerife.
206
207
208
Rowe, M.C., Peate, D.W., and Peate, I.U., 2011, An investigation into the nature of the
magmatic plumbing system at Paricutin Volcano, mexico: Journal of Petrology, v. 52, no.
11, p. 2187–2220, doi: 10.1093/petrology/egr044.
209
210
Rubin, A.M., 1995, Propagation of magma filled-cracks: Annual Review of Earth and Planetary
Science, v. 23, p. 287–336.
211
212
Sánchez, C., 2014, Revisión del catálogo sísmico de las Islas Canarias: Universidad Politécnica
de Madrid.
213
214
215
Saunders, K., Blundy, J., Dohmen, R., and Cashman, K., 2012, Linking Petrology and
Seismology at an Active Volcano: Science, v. 336, no. 6084, p. 1023–1027, doi:
10.1126/science.1220066.
216
217
Strong, M., and Wolff, J., 2003, Compositional variations within scoria cones: Geology, v. 31,
no. 2, p. 143–146, doi: 10.1130/0091-7613(2003)031<0143:CVWSC>2.0.CO;2.
218
219
220
Thorarinsson, S., Steinthórsson, S., Einarsson, T., Kristmannsdóttir, H., and Oskarsson, N., 1973,
The Eruption on Heimaey, Iceland: Nature, v. 241, no. 5389, p. 372–375, doi:
10.1038/241372a0.
221
222
223
Ukawa, M., 1993, Excitation mechanism of large-amplitude volcanic tremor associated with the
1989 Ito-oki submarine eruption, central Japan: Journal of Volcanology and Geothermal
Research, v. 55, no. 1-2, p. 33–50, doi: 10.1016/0377-0273(93)90088-9.
224
225
226
Valentine, G.A., and Gregg, T.K.P., 2008, Continental basaltic volcanoes - Processes and
problems: Journal of Volcanology and Geothermal Research, v. 177, no. 4, p. 857–873, doi:
10.1016/j.jvolgeores.2008.01.050.
227
228
229
Yokoyama, I., and De la Cruz-Reyna, S., 1990, Precursory earthquakes of the 1943 eruption of
Paricutin volcano, Michoacan, Mexico: Journal of Volcanology and Geothermal Research,
v. 44, no. 3-4, p. 265–281, doi: 10.1016/0377-0273(90)90021-7.
230
231
FIGURE CAPTIONS
232
Figure 1. Pre-eruptive unrest in historical monogenetic volcanoes and calculated mixing times.
233
The eruption is represented by a black line at time = 0. Red dashed lines indicate 90 and 15 days
234
prior to the eruption. Grey shadow area corresponds to the 90 days before the eruption. A:
235
Comparison between factual account (triangles) and monitored data (dots) of the unrest period of
236
multiple eruptions. The plots represent the days with felt earthquakes. Generally the trend of the
237
activity changes around two or three months and one or two weeks before the eruption for the
238
studied monogenetic eruptions. The inset shows a zoom view of the grey shadow area. Siete
239
Fuentes, Fasnia and Arafo (light blue triangles); Chinyero (garnet triangles); San Juan (dark blue
240
triangles); Teneguía (purple triangles); El Hierro (orange circles); Jorullo (yellow triangles);
241
Paricutín (green dots); Ito-oki (grey dots). Details of the seismicity are given in the
242
supplementary data. B: Calculated mixing times from some of the considered eruptions. Times
243
from San Juan eruption are not exact (dashed lines).
244
245
Figure 2. Possible plumbing system configuration and evolution of events that may occur below
246
monogenetic volcanoes. This is schematic and not to scale. The depth of the subvolcanic system
247
may vary from system to system, but evidences suggest 5-15 km. The depth of the magma source
248
is also variable but is at least 20 km. A: Intrusion of magma about one or two years prior to the
249
eruption. Stalling of magma at 5-15 km due to the loss of buoyancy or freezing of dykes. Mixing
250
processes registered by the crystals. Crust assimilation in some cases. Seismic activity felt by the
251
population in some cases. B. Renew magma intrusion. Progressive opening of the path between
252
deep and shallow reservoirs. Mixing processes registered by the crystals. Crust assimilation in
253
some cases. The seismic activity is commonly felt by the population. C: Continued intrusion of
254
mafic magma leads to easier transfer from deep to shallow reservoirs and this allows the magma
255
to finally erupt. Mixing processes (and crust assimilation in some cases) registered by the
256
crystals. Seismicity felt by the population.