ArticlePDF Available

Comparison of seismic reflection data to a synthetic seismogram in a volcanic apron at Site 953

Authors:

Abstract

The volcanic apron of Gran Canaria at Site 953 is characterized by numerous, closely spaced reflectors, allowing a high-resolution stratigraphic correlation. The calibration of the presite survey seismic data (during the Meteor Cruise 24) with regard to the lithology and stratigraphy found at the drill site was achieved by computing a synthetic seismogram serving as the link between seismic and borehole data. Because logging data were available for only 53% of the hole, velocity and density mea-surements taken from the recovered cores were used in the missing intervals to obtain a complete synthetic seismogram. Most reflectors in the upper ~900 m of the sequence (lithologic Units I–V) turned out to be thin volcaniclastic layers intercalated to the nonvolcanic background sediments. Their thicknesses are generally <2 m, and the reflections from their tops and bases overlap, forming a single reflection. The limit of the seismic detection of such interbeds is on the order of several decimeters and thus requires special care for the processing of the velocity and density data to avoid destruction of the signal from these thin layers.
Weaver, P.P.E., Schmincke, H.-U., Firth, J.V., and Duffield, W. (Eds.), 1998
Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 157
3
1. COMPARISON OF SEISMIC REFLECTION DATA TO A SYNTHETIC SEISMOGRAM
IN A VOLCANIC APRON AT SITE 9531
Thomas Funck2 and Holger Lykke-Andersen3
ABSTRACT
The volcanic apron of Gran Canaria at Site 953 is characterized by numerous, closely spaced reflectors, allowing a high-
resolution stratigraphic correlation. The calibration of the presite survey seismic data (during the Meteor Cruise 24) with regard
to the lithology and stratigraphy found at the drill site was achieved by computing a synthetic seismogram serving as the link
between seismic and borehole data. Because logging data were available for only 53% of the hole, velocity and density mea-
surements taken from the recovered cores were used in the missing intervals to obtain a complete synthetic seismogram. Most
reflectors in the upper ~900 m of the sequence (lithologic Units I–V) turned out to be thin volcaniclastic layers intercalated to
the nonvolcanic background sediments. Their thicknesses are generally <2 m, and the reflections from their tops and bases
overlap, forming a single reflection. The limit of the seismic detection of such interbeds is on the order of several decimeters
and thus requires special care for the processing of the velocity and density data to avoid destruction of the signal from these
thin layers.
INTRODUCTION
The volcanic apron of Gran Canaria consists of volcaniclastic de-
posits with generally high P-wave velocities and densities intercalat-
ed to the hemipelagic background sediments with lower velocities
and densities (Schmincke, Weaver, Firth, et al., 1995). Although the
volcaniclastic layers are generally <1 m thick, they are conspicuous
in the seismic record by their continuity outside the chaotic or discon-
tinuous slope facies proximal to the island. Because prominent re-
flectors can be used as marker horizons for the mapping of the apron,
it is worthwhile to calculate synthetic seismograms for correlating the
seismic data with the lithology and stratigraphy at the drill sites.
Site 953 (Fig. 1) is the deepest site around Gran Canaria (maxi-
mum penetration 1159 meters below seafloor [mbsf]), reflecting the
entire volcanic and erosional evolution of the island. It thus repre-
sents a key site for the calibration of seismic data. The computed syn-
thetic seismogram is compared with the high-resolution seismic re-
flection Profile 134 (Fig. 1) where Site 953 is located. This profile
was collected during the presite survey of the vessel Meteor Cruise
24 (Schmincke and Rihm, 1994).
The objective of this study was to calculate a complete synthetic
seismogram to benefit as much as possible from the drilling results
for the seismic interpretation. Special care had to be taken in process-
ing the density and velocity data to preserve the signal of the thin vol-
caniclastic interbeds.
DATA PREPARATION
The standard method of obtaining synthetic seismograms is to use
impedance logs from downhole measurements. Impedance is the
product of P-wave velocity and bulk density. However, because Site
953 was not logged intensely (only 53% of the hole was logged), the
logging data set alone was not sufficient to compute a synthetic seis-
mogram for the entire hole. Gaps in the logging data therefore were
filled with laboratory measurements carried out on the recovered sed-
iments aboard ship.
At Site 953, a huge amount of overlapping velocity and density
measurements was available. Velocity data were provided by the P-
wave logger of the multisensor track (MST; depth interval 0192
mbsf), digital sound velocimeter (DSV; 076 mbsf), Hamilton frame
(1871159 mbsf), and sonic log (372963 mbsf). Density data were
available from the gamma-ray attenuation porosity evaluator
(GRAPE) sensor of the MST (01159 mbsf), the index properties (0
1159 mbsf), and the density log (372987 mbsf). These data had to
be processed and merged in a trial and error process to find the best
correlation between seismic Line 134 and the computed synthetic
seismogram.
1Weaver, P.P.E., Schmincke, H.-U., Firth, J.V., and Duffield, W. (Eds.), 1998. Proc.
ODP, Sci. Results, 157: College Station, TX (Ocean Drilling Program).
2Graduiertenkolleg, GEOMAR, Wischhofstr. 1-3, 24148 Kiel, Federal Republic of
Germany. (Present address: Department of Oceanography, Dalhousie University, Hali-
fax, Nova Scotia, B3H 4J1, Canada. tfunck@is.dal.ca).
3Department of Earth Sciences, University of Århus, Finlandsgade 8, 8200 Århus
N, Denmark.
15˚30
´
W 15˚00
´
W 14˚30
´
W
28˚00
´
N
28˚30
´
N
29˚00
´
N
P134
SITE 953
Gran Canaria Fuerteventura
50 km
1000
2000
3000
Figure 1. Generalized bathymetry chart of the volcanic apron north of Gran
Canaria (contours at 500-m intervals), showing the location of Site 953 and
the crossing reflection seismic Line P134 of the Meteor Cruise 24.
T. FUNCK, H. LYKKE-ANDERSEN
4
In the end, the MST data (0192 mbsf) were used for the first part
of the final velocity function (Fig. 2), because the data density of the
overlapping DSV data was only ~1 per 5 m. The erroneous values at
the ends of each measured core section (1.5 m length) were deleted.
Furthermore, a median filter (filter length 0.25 m) had to be applied
because the broad scatter in the data introduced artificial reflections.
The second part of the velocity function consists of Hamilton frame
data (192372 mbsf) as no other data were available. The third part
was provided by the sonic log (372963 mbsf). Because the main part
of reflections is produced by only thin (meter-sized or even less) in-
terbeds of volcaniclastic material, the best results were achieved us-
ing raw data, although some spikes were probably erroneous; in fact,
filtering the logging data seemed to overly flatten the characteristics
(amplitudes became too low). The last part of the velocity function
was made again using the Hamilton frame (9631157 mbsf); howev-
er, a velocity correction had to be applied because the synthetic seis-
mogram did not fit well with the measured data in the lowermost part.
A systematic offset was detected below ~720 mbsf, where the Hamil-
ton frame velocities were systematically higher than the logging ve-
locities. The reason for this deviation is unclear; possible explana-
tions include
1. Systematic error in the Hamilton frame measuring instrument
(improbable);
2. Selective coring, (i.e., softer sediments with a lower velocity
tend to be washed out of the core);
3. Selective measuring, that is, the measured points are not repre-
sentative for the cored sediments; and
4. Anisotropy, because the velocity measured using the Hamilton
frame was made perpendicular to the core axis (i.e., in the bed-
ding plane and not along the core axis).
Nevertheless, the average velocity for the interval 720963 mbsf
was 306 m/s higher for the Hamilton frame than for the sonic log so
that 306 m/s was applied as the velocity correction below 963 mbsf.
The final density function contains only index properties mea-
surements. The GRAPE density was eliminated because the neces-
sary filtering of the values for the upper interval (0192 mbsf), in
combination with the smooth-filtered MST velocities, did not pro-
duce the necessary impedance variations to fit the observed reflec-
tions. In the lower interval down to 1159 mbsf, the GRAPE densities
are erroneous because the applied rotary drilling technique resulted
in core disturbances and variable core diameters. The decision
whether to use index properties or the density log was made on the
basis of a comparison of the correlation with the seismic data. The in-
dex properties proved to be slightly better.
The average density and velocity increase with depth because of
the increasing consolidation and compaction of the hemipelagic
background sediments. Volcaniclastic interbeds show generally pro-
nounced deviations from this average, with a maximum of 3.1 g/cm3
in density and 5.3 km/s in velocity. Figure 2 shows that peaks in ve-
locity do not always correlate with peaks in density, because such in-
tervals are not sampled by both data sets, especially where core re-
covery is low.
SYNTHETIC SEISMOGRAM
The depths of velocity and density function derived from merging
the different data sources were converted to time using the time-depth
relation calculated from the velocity. The data were then sampled at
1 m/s intervals, and the reflectivity was derived from two successive
samples (Fig. 2). The reflectivity was then convolved with the source
0
100
200
300
400
500
600
700
800
900
1000
11001.0 1.5 2.0 2.5 3.0
Density (g/cm
3
)
TWT (ms)
1 2 3 4 5
Hamilton frame
Hamilton
frame
MST
Sonic log
Velocity (km/s)
-0.5 0.0 0.5
Reflectivity Synthetic
Figure 2. Computation of the synthetic seismogram
at Site 953. The bulk density is derived from the
index properties. The P-wave velocity is composed
of multisensor track, Hamilton frame, and sonic log
data. The vertical axis shows the TWT below
seafloor. The synthetic seismogram results from the
reflectivity convolved with the source wavelet used
during the Meteor Cruise 24 presite survey. The
AGC is applied to the synthetic traces.
COMPARISON OF SEISMIC REFLECTION DATA
5
wavelet used during the reflection seismic survey. The wavelet was
computed from stacking 100 direct arrivals using the same sleevegun
cluster (4 × 0.65l) as during the Meteor Cruise 24 (Larsen, Saunders,
Clift, et al., 1994). The computed synthetic seismogram does not
comprise internal multiples, even though some peg-leg multiples be-
tween the volcaniclastic horizons are likely. However, because gaps
in the velocity and density data sets occurred preferentially in the
coarse-grained volcaniclastic intervals, a number of erroneous multi-
ple reflections would be introduced to the synthetic seismogram.
To match the real and synthetic data as close as possible, the same
time-variant band-pass-filter and automatic gain control (AGC), as
used in the processing of the real data, was applied to the synthetic
data (Fig. 3). Furthermore, random noise was added to the synthetic
data. Differences between the synthetic seismogram and the seismic
profile occur where the data density was not adequate to detect thin
interbeds. Discrepancies shallower than 5.0 s two-way traveltime
(TWT) are, for example, caused by an insufficient sampling rate for
the index properties measurements, caused by the rapid flow of cores
in the laboratory. Further downcore (down to ~5.15 s and below 5.65
s TWT), many intervals had low core recovery, probably masking
some reflectors. Nevertheless, the synthetic seismogram fits the seis-
mic data surprisingly well considering that only half of the site was
logged.
CORRELATION OF SEISMIC DATA
WITH LITHOLOGY
By means of the synthetic seismogram computed for Site 953, a
number of reflectors could be assigned to specific lithologies, mainly
volcaniclastic interbeds in the hemipelagic background sediments.
Reflectors identified in the seismic data and in the synthetic seismo-
gram are listed in Table 1, and, when possible, the reflecting lithology
is given. Several of these lithologies are labeled with question marks,
indicating that the reflection could not be determined unequivocally.
For example, this might be the case where amplitudes did not fit very
well or where lithologies in the vicinity were not sampled but could
have caused the reflection. Where only one lithology is given, the re-
flection is caused by the interference of the top and bottom reflection
of a thin interbed surrounded by sediments with a different imped-
ance. These layers are generally thinner than 2 m, often <1 m.
The reflectors are numbered from the seafloor (No. 0) downcore
(Fig. 4). The age of the reflectors is taken from biostratigraphic and
paleomagnetic data (Brunner et al., Chap. 9, this volume). Some re-
flectors, which are suitable for regional mapping because of their
continuity and amplitude, are labeled with letters, corresponding to
volcanic phases on Gran Canaria (Schmincke, 1976, 1982, 1994; Ho-
ernle and Schmincke, 1993a, 1993b) and the neighboring island Ten-
erife in the west (Ancochea et al., 1990). The sediments causing the
reflectors M and F were deposited during the Mogan and Fataga
phase of the Miocene volcanism on Gran Canaria. Reflector H repre-
sents the transition to the volcanic hiatus on Gran Canaria, reflector
T is generated by sediments deposited during the shield stage of Ten-
erife, reflector RN corresponds to the Pliocene Roque Nublo volcan-
ism on Gran Canaria, and the sediments causing reflector Q were de-
posited in the Quaternary with volcanic activity both on Gran Canaria
and Tenerife. This reflector nomenclature was introduced by Funck
(1996). Comparison between reflectors found at the Deep Sea Drill-
ing Project (DSDP) Site 397 (Wissmann, 1979; von Rad, Ryan, et al.,
1979) or Ocean Drilling Program Sites 955 and 956 (Schmincke,
Weaver, Firth, et al., 1995) south of Gran Canaria is hampered be-
cause of the influence of the continental slope, whereas the northern
basin is more or less shielded from continental input by the East Ca-
nary Ridge. Furthermore, no direct seismic correlation from south to
north is possible across the channel between Gran Canaria and Fu-
erteventura/Tenerife, because the reflectors onlap the volcanic base-
ment and/or reflection patterns become chaotic in the proximity of
the islands. This explains why reflector names introduced for the
southern basin at DSDP Site 397 are not suitable in the north. How-
ever, the prominent reflector bands R7 and R3 at DSDP Site 397 (von
Rad, Ryan, et al., 1979) correspond to the reflector bands around M
and RN, respectively, based on their depositional age and seismic fa-
cies (Funck et al., 1996).
DISCUSSION AND CONCLUSIONS
The high-resolution seismic data of the Meteor Cruise 24 revealed
a large number of reflectors from which 55 (Table 1) could be iden-
tified by means of the synthetic seismogram at Site 953. The deposi-
tional setting of the volcanic apron of Gran Canaria with its thin
volcaniclastic layers intercalated to the nonvolcanic background sed-
iments put great demands on the preparation of the velocity and den-
sity data. For example, filtering can remove unrealistic spikes but
also can eliminate the signal of thin layers. Another difficulty that oc-
curred at Site 953 was the incomplete logging data set resulting in the
use of discrete measurements carried out on the cores. Apart from the
fact that these data were only available for the cored intervals, two
problems occurred in the case of the investigated volcanic apron:
1. The use of the closely sampled (0.53.0 cm) MST data with its
broad scatter resulted in artificial reflections, whereas filtering
destroyed almost all real signals from thin layers.
2. The typical sampling rate of one per section (1.5 m) for the in-
dex properties, the Hamilton frame, and DSV measurements
often proved to be insufficient to record all reflectors.
Nevertheless, the careful composition of the velocity and density
function allowed identification of most of the reflectors. The reflec-
tors in lithologic Units I–V (Fig. 4) represent thin layers, generally <2
m thick. This raises the question of the vertical resolution of the seis-
mic data. Badley (1985) states that an interbedded layer has to be
thicker than half the seismic wavelength to enable resolution of its
top and base. Thinner layers appear as one single reflector with a
maximum amplitude at one-quarter wavelength—the tuning thick-
ness. For thicknesses below one-quarter wavelength, the reflection
remains the same shape, but decreases in amplitude. Once the thick-
ness is about one-thirtieth wavelength or less, there is no detectable
response. This explains why thin volcaniclastic beds can be detected
as single reflectors. Applying this knowledge to the data reported
here, the signal contains frequencies of up to 230 Hz, and the velocity
in the volcaniclastic layers is typically between 1700 and 2000 m/s in
the upper 100 m. One-thirtieth of a wavelength thus corresponds to
only 2530 cm, which is the minimum thickness for seismic detec-
tion of such layers. Frequencies around the maximum energy of the
source wavelet (~80100 Hz) correspond to a detection limit in the
order of 5080 cm. Weak and discontinuous reflectors seem to rep-
resent thin layers with a thickness on the order of one-thirtieth of the
wavelength. This is the case for the thin sand layers that are recogniz-
able in the upper 50 m of the seismic record in Fig. 3. Their thickness-
es at Site 953 are ~80 cm (Schmincke, Weaver, Firth, et al., 1995).
Another point to discuss here briefly is the boundaries of the litho-
logic units and their seismic expression. Most of the lithologic
boundaries are not characterized by reflections, at least not by reflec-
tions with a high continuity and amplitude (Fig. 4). The reason for
this is simply the fact that the main criterion to subdivide the se-
quence into lithologic units was the variation of the flux of coarse ma-
terial to Site 953 (Schmincke, Weaver, Firth, et al., 1995). This flux
was averaged for larger intervals (core length), and therefore the
lithologic boundaries typically do not correspond to a volcaniclastic
interbed and its associated reflection. The boundary between litho-
logic Units VI and V, separating the massive basaltic pedestal of
Gran Canaria from the overlying sediments, is characterized by a
high amplitude reflection and represents an exception to the other-
T. FUNCK, H. LYKKE-ANDERSEN
6
Figure 3. Part of seismic Line P134 is plotted together with the synthetic seismogram at Site 953. The seismic data are stacked, scaled with an AGC (50200
ms), and band-pass filtered (30230 Hz at seafloor, 30160 Hz at 6 s TWT). Random noise was added to the 10 synthetic seismic traces. One common depth
point (CDP) corresponds to a horizontal distance of 3.125 m, and the TWT is given in seconds below sea level.
COMPARISON OF SEISMIC REFLECTION DATA
7
wise low correlation between lithologic boundaries and their seismic
expression.
The narrow spacing of the volcaniclastic deposits in conjunction
with the drilling allows a high temporal resolution of the apron. The
nonvolcanic background sediments represent an ideal contrast medi-
um for seismic detection of thin volcaniclastic layers. Hence, a seis-
mic investigation in a volcanic apron benefits from high background
sedimentation rates like around the Canary Islands.
ACKNOWLEDGMENTS
We are grateful to the crew and scientific staff of the research ves-
sels Meteor and JOIDES Resolution for collecting the data in a pro-
fessional manner. Some figures were generated with software provid-
ed by P. Wessel and W.F.H. Smith (1991). The work on this paper
was supported by the Deutsche Forschungsgemeinschaft (DFG-
Schm250/49 III GK, DFG-Schm250/54), the Bundesministerium für
Forschung und Technologie, and the European Union (EPOCH,
EVSV-CT93-0283, MAS2-CT94-0083).
REFERENCES
Ancochea, E., Fúster, J.M., Ibarrola, E., Cendrero, A., Coello, J., Hernán, F.,
Cantagrel, J.M., and Jamond, C., 1990. Volcanic evolution of the island
of Tenerife (Canary Islands) in the light of new K-Ar data. J. Volcanol.
Geotherm. Res., 44:231249.
Badley, M.E., 1985. Practical Seismic Interpretation: Englewood Cliffs, NJ
(Prentice Hall).
Funck, T., 1996. Structure of the volcanic apron north of Gran Canaria
deduced from reflection seismic, bathymetric and borehole data [Ph.D.
dissert.]. Univ. Kiel.
Funck, T., Dickmann, T., Rihm, R., Krastel, S., Lykke-Andersen, H., and
Schmincke, H.-U., 1996. Reflection seismic investigations in the volca-
niclastic apron of Gran Canaria and implications for its volcanic evolu-
tion. Geophys. J. Int., 125: 519536.
Hoernle, K., and Schmincke, H.-U., 1993a. The role of partial melting in the
15-Ma geochemical evolution of Gran Canaria: a blob model for the
Canary Hotspot. J. Petrol., 34:599627.
————, 1993b. The petrology of the tholeiites through melilite nephelin-
ites on Gran Canaria, Canary Islands: crystal fractionation, accumulation,
and depths of melting. J. Petrol., 34:573597.
Larsen, H.C., Saunders, A.D., Clift, P.D., et al., 1994. Proc. ODP, Init.
Repts., 152: College Station, TX (Ocean Drilling Program).
Table 1. Correlated reflectors at Site 953.
Notes: ? = reflection not fully determined. Reflector numbers are shown in Figure 4. Unless specified, the reflections are caused by thin layers of the given lithology intercalated to
background sediments with a different impedance.
Reflector Depth (mbsf) TWT (ms) Age (Ma) Reflection causing lithologies (and comments)
0 0.00 4745 0.00 Seafloor
1 11.45 4760 0.36 Clayey medium to coarse-grained silty pumice sand
2 20.60 4772 0.48 Silty fine medium-grained foraminifer pumice sand
3 23.90 4777 0.53 Calcareous sand
4 30.17 4785 0.61 Basaltic sand
5 45.85 4805 0.82 Silty crystal lithic sand (amplitude difference due to low recovery)
6 55.49 4818 0.95 Clayey nannofossil ooze (weak reflector)
7 60.77 4826 1.03 Crystal lithic sand
8 72.61 4840 1.19 Calcareous sand with volcanic lithics
9 96.40 4871 1.51 Foraminifer lithic sand
10 130.73 4916 1.97 Massive lithic to calcareous sand
11 149.74 4940 2.23 Massive foraminifer sand with lithics (weak reflector, low continuity)
12 170.53 4967 2.62 Foraminifer silt with lithics and crystals
13 176.00 4974 2.70 Clayey nannofossil ooze?
14 206.23 5010 3.12 Foraminifer sandstone
15 225.70 5030 3.39 Foraminifer lithic sandstone
16 237.59 5041 3.56 Lapillistone
17 257.54 5058 3.83 Lapillistone (weak reflector, low continuity)
18 267.67 5070 4.24 Silty, foraminifer nannofossil chalk (weak reflector)
19 278.07 5081 4.48 Nannofossil chalk (weak, discontinuous reflector)
20 296.19 5099 4.91 Density/velocity increase in a nannofossil chalk sequence
21 313.12 5118 5.32 Lithic crystal sand
22 325.53 5130 5.61 Indurated nannofossil ooze (weak reflector)
23 337.50 5144 5.90 Clay with nannofossils (grading to lithic crystal sand), weak reflector
24 350.81 5157 6.21 Slump unit consisting of ooze, clay, silt, mixed rock
25 361.35 5167 6.21 ?
26 385.85 5185 7.60 Foraminifer lithic crystal sand (weak, discontinuous reflector)
27 399.73 5200 8.28 Indurated clay with nannofossil (weak and discontinuous reflector)
28 409.80 5210 8.46 Calcareous sandstone?
29 417.79 5219 8.60 Indurated clay?
30 442.71 5242 9.03 Silty claystone
31 455.00 5254 9.25 Claystone?
32 470.39 5271 9.52 Nannofossil chalk?
33 495.30 5293 9.96 Lithic crystal sandstone?
34 516.44 5312 10.33 Nannofossil claystone
35 556.83 5350 11.04 Lithic crystal sandstone
36 594.71 5385 11.71 Lithic crystal sandstone
37 601.86 5390 11.82 Lapillistone
38 621.49 5408 11.82 Lithic crystal sandstone?
39 633.90 5419 11.98 ? (logged but not recovered)
40 689.14 5467 12.71 Siltstone
41 710.40 5485 12.99 Siltstone? (weak and discontinuous reflector)
42 749.72 5519 13.37 Nannofossil claystone, graded to crystal siltstone
43 755.34 5524 13.41 Claystone? (close to not sampled sandstone)
44 768.32 5535 13.50 Vitric tuff
45 787.20 5552 13.63 Calcareous vitric siltstone? (weak reflector)
46 794.45 5559 13.68 Lithic crystal vitric tuff
47 822.44 5584 13.90 Vitric rich claystone with sandy vitric tuff
48 842.17 5602 14.25 Nannofossil claystone grading to lithic crystal sandstone with pumice
49 863.22 5618 14.61 Claystone grading down to lithic crystal sandstone (weak reflector)
50 920.42 5660 14.80-15.80 Lapillistone with basaltic breccia
51 988.44 5709 14.80-15.80 Pebble- and granule-sized, fine-grained hyaloclastite tuff
52 1029.97 5737 >15.80 High velocity unit in hyaloclastite lapillistone
53 1066.83 5756 >15.80 Transition from hyaloclastite lapillistone breccia to hyaloclastite tuff
54 1110.61 5781 >15.80 Transition from basaltic hyaloclastite breccia to hyaloclastite tuff
55 1139.86 5797 >15.80 Transition from hyaloclastite breccia to hyaloclastite lapillistone
T. FUNCK, H. LYKKE-ANDERSEN
8
Schmincke, H.-U., 1976. The geology of the Canary Islands. In Kunkel, G.
(Ed.), Biogeography and Ecology in the Canary Islands: The Hague (W.
Junk), 67184.
————, 1982. Volcanic and chemical evolution of the Canary Islands. In
von Rad, U., Hinz, K., Sarnthein, M., and Seibold, E. (Eds.), Geology of
the Northwest African Continental Margin: Berlin (Springer), 273306.
————, 1994. Geological Field Guide: Gran Canaria (7th ed.): Kiel, Ger-
many (Pluto Press).
Schmincke, H.-U., and Rihm, R., 1994. Ozeanvulkan 1993, Cruise No. 24,
15 April9 May 1993. METEOR-Berichte, Univ. Hamburg, 94-2.
Schmincke, H.-U., Weaver, P.P.E., Firth, J.V., et al., 1995. Proc. ODP, Init.
Repts., 157: College Station, TX (Ocean Drilling Program).
von Rad, U., Ryan, W.B.F., et al., 1979. Init. Repts. DSDP, 47 (Pt. 1): Wash-
ington (U.S. Govt. Printing Office).
Wessel, P., and Smith, W.H.F., 1991. Free software helps map and display
data. Eos, 72:441, 445446.
Wissmann, G., 1979. Cape Bojador slope, an example for potential pitfalls in
seismic interpretation without the information of outer margin drilling. In
von Rad, U., Ryan, W.B.F., et al., Init. Repts. DSDP, 47 (Pt. 1): Washing-
ton (U.S. Govt. Printing Office), 491499.
Date of acceptance: 6 January 1997
Date of initial receipt: 24 June 1996
Ms 157SR-100
COMPARISON OF SEISMIC REFLECTION DATA
9
4.8
5.0
5.2
5.4
5.6
5.8
I
II
III
IV
V
VI
VII
A
B
C
Lithologic UnitM24-LINE 134 M24-LINE 134 Reflector
Synthetic
seismogr.
TWT (s)
Water
S e d i m e n t s
(volcanic / non-volcanic)
Basaltic
pedestal
Gran Canaria
Hyaloclastite
debris flows
1159 mbsf
maximum
penetration
0
1
2
4
5 (Q)
6
8
9
10
11
12
14
15
16 (RN)
17
18
19
20
21 (T)
22
23
24
25
26
27
28
29
30
31
32 (H)
33
34
35
36
38 (F)
39
40
41
42
44
45
47 (M)
48
49
50
51
52
53
54
55
Age (Ma)
01020
age
Figure 4. Part of seismic Line P134 with synthetic seismogram at Site 953 and the lithologic units converted to TWT (in seconds below sea level). The two age
lines in lithologic Unit VII give minimum and maximum ages. The small numbers to the right of the seismic data refer to individual reflectors as listed in Table
1.
... Next, for seismic modeling, 'Wavelet estimation through matching' method as described in White and Simm (2003) was followed. Reflectivity series was generated from log data in two-way traveltime (TWT) and convolved (Funck et al, 1998) with Ricker wavelet. This was done for original wells within the 3D seismic survey area and the two pseudo-wells created. ...
Article
Full-text available
One of the most common problems in the North Sea is the occurrence of salt (solid) in the pores of Triassic sandstones. Many wells have failed due to interpretation errors based conventional substitution as described by the Gassmann equation. A way forward is to device a means to model and characterize the salt-plugging scenarios. Modelling the effects of fluid and solids on rock velocity and density will ascertain the influence of pore material types on seismic data. In this study, two different rock physics modelling approaches are adopted in solid-fluid substitution, namely the extended Gassmann theory and multi-mineral mixing modelling. Using the modified new Gassmann equation, solid-and-fluid substitutions were performed from gas or water filling in the hydrocarbon reservoirs to salt materials being the pore-filling. Inverse substitutions were also performed from salt-filled case to gas- and water-filled scenarios. The modelling results show very consistent results - Salt-plugged wells clearly showing different elastic parameters when compared with gas- and water-bearing wells. While the Gassmann equation-based modelling was used to discretely compute effective bulk and shear moduli of the salt plugs, the algorithm based on the mineral-mixing (Hashin-Shtrikman) can only predict elastic moduli in a narrow range. Thus, inasmuch as both of these methods can be used to model elastic parameters and characterize pore-fill scenarios, the New Gassmann-based algorithm, which is capable of precisely predicting the elastic parameters, is recommended for use in forward seismic modelling and characterization of this reservoir and other reservoir types. This will significantly help in reducing seismic interpretation errors.
... We thus present depth-converted maps ( Figure S1 in the supporting information) using a range of seismic velocities from 2,000 m/s to 4,000 m/s within the sediment. A seismic velocity of 2,000 m/s is comparable to the P wave velocities measured at Site 953 in the volcaniclastic apron offshore Gran Canaria (e.g., Funck & Lykke-Andersen, 1998) and in the Lesser Antilles (e.g., Le Friant et al., 2015). For the thickness of the seismostratigraphic units, there is a linear relationship between the TWT time and the figure in meters. ...
Article
Full-text available
High-resolution seismic reflection profiles gathered in 2006 on La Réunion submarine flanks and surrounding abyssal plain, enabled characterization of the seismostratigraphy architecture of the volcaniclastic apron. Four seismic units are defined beyond the edifice base: (1) a basal unit, interpreted as pelagic sediment predating La Réunion volcanism; (2) a second unit showing low- to medium-amplitude reflections, related to La Réunion emergence including the submarine explosive phase; (3) a high-amplitude seismic unit, associated with subaerial volcanic activity (i.e., mature island stage); and (4) an acoustically transparent unit, ascribed to erosion that currently affects the volcanic complex. Two prominent horizons delineate the base of the units II and III marking, respectively, the onset of La Réunion seamount explosive activity and the Piton des Neiges volcanic activity. Related isopach maps demonstrate: (1) the existence of a large proto-Piton des Neiges volcano during the first building phase of the volcanic complex, and (2) the central role of the Piton des Neiges volcano during the second phase. Shield growth stage of the Piton de la Fournaise volcano is also captured in the upper part of the volcaniclastic apron, attesting to its recent contribution. Seismic facies identified in the apron highlight a prevalence of sedimentary and reworking processes since the onset of the volcanism compared to catastrophic flank collapses. We present here a new model of evolution for La Réunion volcanic complex since the onset of the volcanism and argue a major proto-Piton des Neiges – Piton des Neiges volcanic complex controls La Réunion present-day morphology.
... Otherwise conditions have been relatively quiet on the rise since the Oligocene with oceanographic upwelling resulting in high accumulation rates of calcareous hemipelagic sediments. Geophysical logging at DSDP site 397 showed a significant acoustic impedance contrast between volcanic components and both the margin (generally quartz and feldspar) and biogenic deposits (Funck & Lykke Anderson 1998). Discrete volcanic layers are therefore potentially good seismic reflectors provided they are thick enough (as a rule of thumb a thin layer needs to be of the order of one thirtieth of a wavelength to be detected for 10 20 Hz typical of deep seismic data this is equivalent to a few metres). ...
Article
We use multichannel seismic reflection profiles to determine the seismic stratigraphy of the flexural moat that flanks the Canary Islands. The moat stratigraphy has been divided into 5 units on the basis of internal character and correlation of distinctive reflections. The deepest units, I and II, which well-ties indicate are Eocene and older, thicken towards the east suggesting they are the consequence of sediment loading at the Moroccan continental margin. Units III, IV and V, which are Oligocene and younger and highly reflective, thicken concentrically around individual islands suggesting they are dominantly the result of volcanic loading. Distinct stratigraphic patterns of onlap at the base and offlap at the top of individual flexural units are seen on the across-moat profiles but they were not easily identified on our limited along-moat profiles. The thickness of the upper three units is in accord with the predictions of flexural loading models. Moreover, a model in which the volcanoes that make up the Canary Islands progressively load the underlying lithosphere from east to west generally accounts for the thickness variations that are observed in the region of individual islands. We date the shield building stages of the Fuerteventura, Gran Canaria and La Gomera as Oligocene to Early Miocene, that of Tenerife as Middle Miocene to Late Miocene and those of La Palma and El Hierro as Pliocene to Quaternary. The best overall fit to stratigraphic data in the northern moat is for an elastic thickness of the lithosphere, Te, of 35 km, which is similar to the 30–40 km which would be expected for Oligocene and Neogene loading of Jurassic oceanic lithosphere. There is evidence that a contribution from the margin is required to explain the divergence of Units III, IV and V along the Moroccan margin. Detailed modelling of an along-strike seismic profile of the moat north of Tenerife and Gran Canaria, however, suggests that flexure due to island loading fully explains the stratigraphic patterns that are observed and does not require an additional contribution from the margin. The most likely explanation for this observation is that a ‘barrier’ had developed by the Oligocene, along the present trend of Fuerteventura and Lanzarote, which prevented sediments from the Moroccan margin infilling the northern parts of the moats caused by volcanic loading. Furthermore, there is evidence from differences in the thickness of Units I and II that a barrier may also have existed prior to the Oligocene which protected the northern basin from corrosive bottom currents that removed large amounts of late Cretaceous and Palaeogene age material from the southern basin.
... The reflectivity is generally higher in the northern basin. Funck and Lykke-Andersen (1998a) showed that most reflectors in the northern basin are the result of volcaniclastic material derived from the islands showing a large impedance contrast to the hemipelagic background sedimentation. The lower reflectivity in the southern basin indicates a lower amount of volcaniclastic material in this basin, an interpretation supported by drilling (Schmincke et al. 1995). ...
Article
Full-text available
Seismic, sidescan sonar, bathymetric multibeam and ODP (Ocean Drilling Program) data obtained in the submarine channel between the volcanic islands of Gran Canaria and Tenerife allow to identify constructive features and destructive events during the evolution of both islands. The most prominent constructive features are the submarine island flanks being the acoustic basement of the seismic images. The build-up of Tenerife started following the submarine stage of Gran Canaria because the submarine island flank of Tenerife onlaps the steeper flank of Gran Canaria. The overlying sediments in the channel between Gran Canaria and Tenerife are chaotic, consisting of slumps, debris flow deposits, syn-ignimbrite turbidites, ash layers, and other volcaniclastic rocks generated by eruptions, erosion, and flank collapse of the volcanoes. Volcanic cones on the submarine island flanks reflect ongoing submarine volcanic activity. The construction of the islands is interrupted by large destructive events, especially by flank collapses resulting in giant landslides. Several Miocene flank collapses (e.g., the formation of the Horgazales basin) were identified by combining seismic and drilling data whereas young giant landslides (e.g., the Gimar debris avalanche) are documented by sidescan, bathymetric and drilling data. Sediments are also transported through numerous submarine canyons from the islands into the volcaniclastic apron. Seismic profiles across the channel do not show a major offset of reflectors. The existence of a repeatedly postulated major NE-SW-trending fault zone between Gran Canaria and Tenerife is thus in doubt. The sporadic earthquake activity in this area may be related to the regional stress field or the submarine volcanic activity in this area. Seismic reflectors cannot be correlated through the channel between the sedimentary basins north and south of Gran Canaria because the channel acts as sediment barrier. The sedimentary basins to the north and south evolved differently following the submarine growth of Gran Canaria and Tenerife in the Miocene.
Article
The morphology and structure of the submarine flanks of Gran Canaria have been mapped using Hydrosweep swath bathymetry and high-resolution reflection seismic data. The growth and destruction of the island has been characterized previously by three major periods of volcanic activity (16-9 Ma, 4.5-3.5 Ma, younger than 3 Ma) separated by erosional intervals. Two major sector collapses along the west coast, inferred from the coastal morphology, are believed to have formed at the end of the shield-building phase. One is characterized by a 19-km-wide reentrant along the northwestern coast that may have formed synchronously with the formation of the 20-km-diameter Miocene Tejeda Caldera. High sedimentation rates around Gran Canaria (>50 m/Myr) tend to cover and bury major landslide blocks. SSW off the island, several canyons continue seaward into a major sediment fan. A 9.5-km-wide volcaniclastic ridge in this fan is interpreted to represent deposits of the Pliocene subaerial Roque Nublo debris avalanche. We tentatively interpret the slope break at a depth of 600-800 m as the transition between subaerial and subaqueous chilled lavas at the end of the shield-building phase. The subsidence caused by the volcanic load (30,000 km3) on the lithosphere may thus amount to no more than 800 m. Several canyons on the island can be traced down the submarine flanks to a depth of 3.5 km, indicating that at least deeper portions below the level of subsidence were eroded by mass flows continuing seaward from the subaerial canyons. Four submarine volcanoes were identified west and northeast of the island.
Article
The fault zone of a mature large-displacement fault may be idealized as a nested hierarchical structure consisting of a core of extremely fine grained material surrounded by coarser granulated gouge and breccia which is in turn bordered by fracture-damaged wall rock in which the fracture density decreases with distance to a regional background level. While there are significant variations in the symmetry of this structure, virtually all fault zones have a core of deformed granular rock within which most of the displacement appears to have occurred, often on an extremely narrow prominent slip surface. In many faults, the gouge and breccia layer is missing from one or both sides of the fault zone. For strike-slip faults, this appears to be associated with variations in the lithology of the wall rock. For dip-slip faults it is most likely a consequence of the exhumation of only one wall by the fault motion. In normal faults, the layered structure appears on the footwall while the hanging wall shows almost no damage. For reverse faults it is the hanging wall which contains the layered structure. Mechanisms proposed to explain the formation of these fault zone structures are reviewed with an eye toward whether they shed any light on the earthquake process.
Article
Keating and McGuire (2000) [Island edifice failures and associated tsunami hazards. In “Landslides and Tsunamis”, Birkhauser, Boston, pp. 899–956] presented and examined evidence for ubiquitous island edifice failures and their associated tsunami hazards. In this follow-up review, we examine the status of landslide, debris flow and tsunami research and find that significant progress has been made in placing constraints on physical parameters that will facilitate numerical modeling of tsunami, landslide and debris flow movements. Similarly, physical modeling has provided an important contribution to our understanding of slope failure and debris transport, with many features generated in laboratory experiments clearly identifiable in sonar images of mass wasting events.
Article
Full-text available
When creating camera‐ready figures, most scientists are familiar with the sequence of raw data → processing → final illustration and with the spending of large sums of money to finalize papers for submission to scientific journals, prepare proposals, and create overheads and slides for various presentations. This process can be tedious and is often done manually, since available commercial or in‐house software usually can do only part of the job. To expedite this process, we introduce the Generic Mapping Tools (GMT), which is a free, public domain software package that can be used to manipulate columns of tabular data, time series, and gridded data sets and to display these data in a variety of forms ranging from simple x‐y plots to maps and color, perspective, and shaded‐relief illustrations. GMT uses the PostScript page description language, which can create arbitrarily complex images in gray tones or 24‐bit true color by superimposing multiple plot files. Line drawings, bitmapped images, and text can be easily combined in one illustration. PostScript plot files are device‐independent, meaning the same file can be printed at 300 dots per inch (dpi) on an ordinary laserwriter or at 2470 dpi on a phototypesetter when ultimate quality is needed. GMT software is written as a set of UNIX tools and is totally self contained and fully documented. The system is offered free of charge to federal agencies and nonprofit educational organizations worldwide and is distributed over the computer network Internet.
Chapter
The Canary Islands, a group of seven major volcanic islands, extends for almost 500 km roughly east-west 100 km off Northwest Africa. The islands formed chiefly during the last 20 Ma, although volcanic activity started during the Oligocene and possibly Eocene in the eastern island of Fuerteventura. Ages of the rapidly formed sub-Canarian mantle are presently active across the entire belt. Total volumes of individual islands are about 10 to 20 x 106 km3 of which the subaerial part generally makes up less than 10%. turated to moderately undersaturated alkali basalt with local tholeiite. Low pressure fractionation of olivine, clino- pyroxene, and plagioclase was generally moderate, owing to rapid replenishment of the fast upward growth of the shield volcanoes and their magma chambers. Highly differentiated magma columns developed chiefly during the waning stages resulting in minor (quartz)-trachyte in the eastern and phonolitic plugs in the central and western islands. Major differentiated magma reservoirs on Gran Canaria and Tenerife culminated in large caldera-forming ash flow eruptions. Surface eruption of basalt magmas was generally inhibited during evolution and periodic partial emptying of such large differentiated zoned magma columns. Late stage basanites to nephelinites are locally nodule-bearing, of small volume, and are only slightly fractionated. High Ca/Al ratios and variable K-contents of these primitive magmas suggest garnet and phlogopite as residual phases during very low degrees of partial melting. Multiphase episodic magmatic evolution consisting of two or more magmatic phases is characteristic of most Canary Islands and is best developed on Gran Canaria where two major multiphase cycles are distinguished. Multiphase magmatic evolution is common on other islands in the Central North Atlantic with alkali basalt shield magmas being broad-ly similar. It is less well developed on smaller islands and those close to the Mid-Atlantic Ridge. Highly alkalic, mafic, under- saturated magmas appear to be restricted to (large volume?) islands on thicker lithosphere (Canaries and Cape Verde Islands), presumably due to low heat flow and thus small degrees of partial melting at greater depth. Intra-archipelago differences in melting conditions and mantle composition are reflected by consistently higher alkalinity and different trace element ratios between the western and central islands contrasted with Lanzarote and Fuerteventura to the east. Canary Island magmas on the whole are richer in Ti, Fe, and Zr and lower in A1 than Azorean and Madeira magmas. Canary Island magmas may be derived from garnet-bearing manle leaving residual garnet. The mantle beneath the Canaries is not very radiogenic with respect to 87Sr/86Sr as is characteristic for the eastern central Atlantic en-compassing the Cape Verde Islands and Madeira. The mantle area south of about 30 to 35 N may be distinct from, and less heterogeneous than the mantle farther north. There is no geological or geochemical evidence for the existence of continental crust beneath any of the Canary Islands. The origin of the Canary Island melting domain is not adequately explained by (a) an oceanic fracture zone, (b) extension of the South Atlas fault, (c) mantle plume and (d) propagating fracture zone. Unspecified mantle instabilities along the boundary between oceanic and continental lithosphere may have been instrumental in generating the unusually long-lived mantle anomaly with west to east translation of the lithosphere leading to an irregular non-linear age progression. Age data presently available for island volcanism in the Eastern Central North Atlantic suggest episodes of high activity between about 18 and 10 Ma and 5 Ma to the present, separated by a period of lesser magmatic productivity.
Article
High-resolution reflection seismic data obtained around Gran Canaria allow a detailed and consistent correlation of seismic reflectors of the northern and southern Canary Basins with the lithology drilled by DSDP Leg 47A SSE of Gran Canaria, as well as with major phases of volcanic activity on Gran Canaria as mapped onshore. Two prominent reflectors were chosen as marker horizons and correlated with the drilled lithology. the results indicate that reflector R7 above the Miocene volcaniclastic debris flows V1-V3 reflects the shield-building phase of Gran Canaria. Reflector R3 is interpreted as corresponding with the Pliocene Roque Nublo formation. The top of the massive island flank of Gran Canaria, defined by seismically chaotic facies, extends 44 to 72 km off the coast of Gran Canaria. West of Gran Canaria the flank of Tenerife onlaps the steeper and older flank of Gran Canaria, which, in turn, is onlapping the older flank of Fuerteventura to the east in a similar way. Erosional channels, which can also be traced up to 50 km from the area between Gran Canaria and Fuerteventura into the deeper northern basin, have been identified in the bathymetry. The data presented provide new detailed information for modelling the submarine and subaerial evolution of the central Canary Islands of Gran Canaria and Tenerife, i.e. the timing of their shield-building phases and later stages of major volcanic activity, as reflected by the position of prominent seismic reflectors in the seismic stratigraphy.
Article
During its evolution, the continental slope is frequently scoured by erosional events which may not always show up as unconformities on the seismic record. Pitfalls await the seismic interpreter, especially in the lower slope region where many of these unconformable reflec- tors merge and meet with the seemingly simple, conformably bedded horizons of the upper rise. In this setting, the acoustostratigraphy should be checked by drilling and refined by logging. Examples of pre- and post-Site 397 interpretation of profiles across the Cape Bo- jador continental margin are presented. They prove that slumps not clearly resolved with low-frequency airgun records mimic the con- tinuation of reflectors from the lower slope to the rise and cause er- roneous acoustostratigraphic identifications. Information about the undrilled section and the regional setting is discussed together with refraction seismic data. This results in a warning not to stratigraph- ically correlate layers with the same refraction velocities in the zone of facies change on the continental slope and upper rise.
Article
New age determinations from Tenerife, together with those previously published (93 in all), provide a fairly comprehensive picture of the volcanic evolution of the island. The oldest volcanic series, with ages starting in the late Miocene, are formed mainly by basalts with some trachytes and phonolites which appear in Anaga, Teno and Roque del Conde massifs. In Anaga (NE), three volcanic cycles occurred: one older than 6.5 Ma, a second one between 6.5 and 4.5 Ma, with a possible gap between 5.4 and 4.8 Ma, and a late cycle around 3.6 Ma. In Teno (NW), after some undated units, the activity took place between 6.7 and 4.5 Ma, with two main series separated by a possible pause between 6.2 and 5.6 Ma. In the zone of Roque del Conde (S), the ages are scattered between 11.6 and 3.5 Ma. Between 3.3 and 1.9 Ma, the whole island underwent a period of volcanic quiescence and erosion.
Article
Thesis (doctoral)--Christian-Albrechts-Universität zu Kiel, 1996. Includes bibliographical references (p. 127-132).
Article
The subaerial portion of Gran Canaria, Canary Islands, was built by three cycles of volcanism: a Miocene Cycle (8⋅5–15 Ma), a Pliocene Cycle (1⋅8–6⋅0 Ma), and a Quaternary Cycle (1⋅8–0 Ma). Only the Pliocene Cycle is completely exposed on Gran Canaria; the early stages of the Miocene Cycle are submarine and the Quaternary Cycle is still in its initial stages. During the Miocene, SiO 2 saturation of the mafic volcanics decreased systematically from tholeiite to nephelinite. For the Pliocene Cycle, SiO 2 saturation increased and then decreased with decreasing age from nephelinite to tholeiite to nephelinite. SiO 2 saturation increased from nephelinite to basanite and alkali basalt during the Quaternary. In each of these cycles, increasing melt production rates, SiO 2 saturation, and concentrations of compatible elements, and decreasing concentrations of some incompatible elements are consistent with increasing degrees of partial melting in the sequence melilite nephelinite to tholeiite. The mafic volcanics from all three cycles were derived from CO 2 -rich garnet lherzolite sources. Phlogopite, ilmenite, sulfide, and a phase with high partition coefficients for the light rare earth elements (LREE), U, Th, Pb, Nb, and Zr, possibly zircon, were residual during melting to form the Miocene nephelinites through tholeiites; phlogopite, ilmenite, and sulfide were residual in the source of the Pliocene–Quaternary nephelinites through alkali basalts. Highly incompatible element ratios (e.g., Nb/U, Pb/Ce, K/U, Nb/Pb, Ba/Rb, Zr/Hf, La/Nb, Ba/Th, Rb/Nb, K/Nb, Zr/Nb, Th/Nb, Th/La, and Ba/La) exhibit extreme variations (in many cases larger than those reported for all other ocean island basalts), but these ratios correlate well with degree of melting. Survival of residual phases at higher degrees of melting during the Miocene Cycle and differences between major and trace element concentrations and melt production rates between the Miocene and Pliocene tholeiites suggest that the Miocene source was more fertile than the Pliocene–Quaternary source(s). We propose a blob model to explain the multi-cycle evolution of Canary volcanoes and the temporal variations in chemistry and melt production within cycles. Each cycle of volcanism represents decompression melting of a discrete blob of plume material. Small-degree nephelinitic and basanitic melts are derived from the cooler margins of the blobs, whereas the larger-degree tholeiitic and alkali basaltic melts are derived from the hotter centers of the blobs. The symmetrical sequence of mafic volcanism for a cycle, from highly undersaturated to saturated to highly undersaturated compositions, reflects melting of the blob during its ascent beneath an island in the sequence upper margin-corelower margin. Volcanic hiatuses between cycles and within cycles represent periods when residual blob or cooler entrained shallow mantle material fill the melting zone beneath an island.
Article
We report major and trace element X-ray fluorescence (XRF) data for mafic volcanics covering the 15-Ma evolution of Gran Canaria, Canary Islands. The Miocene (12–15 Ma) and Pliocene-Quaternary (0–6 Ma) mafic volcanics on Gran Canaria include picrites, tholeiites, alkali basalts, basanites, nephelinites, and melilite nephelinites. Olivine�clinopyroxene are the major fractionating or accumulating phases in the basalts. Plagioclase, Fe–Ti oxide, and apatite fractionation or accumulation may play a minor role in the derivation of the most evolved mafic volcanics. The crystallization of clinopyroxene after olivine and the absence of phenocrystic plagioclase in the Miocene tholeiites and in the Pliocene and Quaternary alkali basalts and basanites with MgO>6 suggests that fractionation occurred at moderate pressure, probably within the upper mantle. The presence of plagioclase phenocrysts and chemical evidence for plagioclase fractionation in the Miocene basalts with MgO<6 and in the Pliocene tholeiites is consistent with cooling and fractionation at shallow depth, probably during storage in lower-crustal reservoirs. Magma generation at pressures in excess of 3⋅0–3⋅5 GPa is suggested by (a) the inferred presence of residual garnet and phlogopite and (b) comparison of FeO1 cation mole percentages and the CIPW normative compositions of the mafic volcanics with results from high-pressure melting experiments. The Gran Canaria mafic magmas were probably formed by decompression melting in an upwelling column of asthenospheric material, which encountered a mechanical boundary layer at {small tilde}100-km depth.
  • U Von Rad
  • W B F Ryan
von Rad, U., Ryan, W.B.F., et al., 1979. Init. Repts. DSDP, 47 (Pt. 1): Washington (U.S. Govt. Printing Office).