Está en la página 1de 5

GEOPHYSICAL RESEARCH LETTERS, VOL. 40, 1–5, doi:10.1002/grl.

50942, 2013

Density-driven transport in the umbrella region of volcanic clouds:


Implications for tephra dispersion models
Antonio Costa,1 Arnau Folch,2 and Giovanni Macedonio3
Received 17 July 2013; revised 5 September 2013; accepted 6 September 2013.

[1] Large explosive volcanic eruptions can generate ash [3] Most Tephra Transport and Dispersal Models
clouds from rising plumes that spread in the atmosphere (TTDM) [Folch, 2012] build on the assumption of passive
around a Neutral Buoyancy Level (NBL). These ash clouds transport, i.e., they assume that the dispersion and sedimen-
spread as inertial intrusions and are advected by atmospheric tation of tephra particles in the atmosphere are governed
winds. For low mass flow rates, tephra transport is mainly by wind advection, atmospheric turbulent diffusion, and
dictated by wind advection, because ash cloud spreading due settling of particles by gravity (e.g., Advection-Diffusion-
to gravity current effects is negligible (passive transport). Sedimentation (ADS) models) [e.g., Armienti et al., 1988;
For large mass flow rates, gravity-driven transport at the Folch et al., 2009]. Conditions under which this passive
NBL can be the dominant transport mechanism. Conditions transport assumption is valid have not been critically exam-
under which the passive transport assumption is valid have ined previously, although it has been shown that ADS
not yet been critically studied. We analyze the conditions models capture the features of past deposits and reproduce
when gravity-driven transport is dominant in terms of the correct order of magnitude of observations, such as
the cloud Richardson number. Moreover, we couple an accumulated tephra thickness, grain size distribution, and
analytical model that describes cloud spreading as a gravity concentration in the atmosphere. Good performance of pas-
current with an advection-diffusion model. This coupled sive TTDMs (see Folch [2012] for details) is in apparent
model is used to simulate the evolution of the volcanic contradiction to gravity current models for volcanic plumes,
cloud during the climatic phase of the 1991 Pinatubo especially for high-intensity eruptions and for proximal
eruption. Citation: Costa, A., A. Folch, and G. Macedonio (2013), transport and proximal depositional facies. One of the rea-
Density-driven transport in the umbrella region of volcanic clouds: sons even simple analytical ADS models can model many
Implications for tephra dispersion models, Geophys. Res. Lett., 40, tephra deposits is that these models use an effective diffu-
doi:10.1002/grl.50942. sion coefficient (commonly calibrated from past deposits)
that is larger than the actual atmospheric turbulent diffusion,
1. Introduction although the actual spreading due to gravity-driven transport
[2] Atmospheric transport of tephra released during follows a different time evolution and should be localized
explosive volcanic eruptions is significantly affected by the around the source. In this way, even simpler ADS models
interaction of the volcanic plume and the atmospheric wind are able to mimic the effective increase of the cloud area at
field. Weak volcanic plumes develop in the troposphere and the NBL observed in large-magnitude eruptions.
follow bent-over trajectories as a result of the wind advec- [4] Here we describe the conditions when gravity-driven
tion [e.g., Carey and Sparks, 1986; Bonadonna and Phillips, transport is dominant and when it is negligible in terms
2003]. Strong plumes typically rise above the tropopause, of the cloud Richardson number, using an analytical model
developing a vertical eruption column that, on reaching the describing the radial growth of the cloud. Then we cou-
Neutral Buoyancy Level (NBL), spreads laterally as a tur- ple this analytical gravity-driven model with the FALL3D
bulent gravity current. Under these conditions, the vertical [Costa et al., 2006; Folch et al., 2009] ADS model in order
plume rises above the NBL due to its excess momentum and, to evaluate the relative importance of gravity current effects.
once its vertical velocity approaches zero, subsides feeding Finally, the coupled model is used to simulate the evolu-
the current that spreads as a lateral intrusion [e.g., Woods tion of the volcanic cloud during the climatic phase of the
and Kienle, 1994; Sparks et al., 1997]. The larger and higher 1991 Pinatubo eruption, for which satellite observations are
the intensity of the eruption, the more this transport mech- available and previous studies can help to constrain the key
anism dominates over passive wind advection at distances eruption source parameters [e.g., Koyaguchi, 1996; Holasek
from tens to several hundreds of kilometers from the source et al., 1996a; Suzuki and Koyaguchi, 2009].
[Baines and Sparks, 2005].
2. Density-Driven Cloud Model
1
Istituto Nazionale di Geofisica e Vulcanologia, Sezione di Bologna,
Bologna, Italy.
[5] In the initial phase of a Plinian eruption, volcanic
2
Barcelona Supercomputing Center, Barcelona, Spain. clouds spread as a gravity current around the NBL [e.g.,
3
Istituto Nazionale di Geofisica e Vulcanologia, Osservatorio Woods and Kienle, 1994; Sparks et al., 1997]. The radius of
Vesuviano, Napoli, Italy. the umbrella cloud, R, can be written as a function of time as
Corresponding author: A. Costa, Istituto Nazionale di Geofisica e
[e.g., Woods and Kienle, 1994; Sparks et al., 1997]
Vulcanologia, Sezione di Bologna, Via Donato Creti 12, I-40128 Bologna,
Italy. (antonio.costa@bo.ingv.it)  1/3
3Nq
©2013. American Geophysical Union. All Rights Reserved. R= t2/3 (1)
0094-8276/13/10.1002/grl.50942 2

1
COSTA ET AL.: DENSITY-DRIVEN TRANSPORT OF TEPHRA

where t is time,  is an empirical constant, N is the fre- and McQuaid [1988]. Based on these observations, we con-
quency of Brunt-Väisälä due to the ambient stratification of sider that when Ri > 1, the transport is mainly density driven,
the atmosphere, and q is the volumetric flow rate into the whereas for Ri < 0.25, transport is substantially passive.
umbrella region. The value of  was first estimated to be [6] From (7), we can estimate the critical time scales
in the range  = 0.1 – 0.6 from laboratory experiments and characterizing density-driven (tb ) and passive (tp ) transport
satellite observations [e.g., Holasek et al., 1996a; Holasek processes:
et al., 1996b] and then constrained to   0.2 from Direct 4Nq
tb = (8)
Numerical Simulations (DNS) [e.g., Suzuki and Koyaguchi, 9u3w
2009]. The volumetric flow rate into the umbrella region can
be estimated as a function of the efficiency of air entrain- 32Nq
tp = (9)
ment, k, and the mass eruption rate MP as [Morton et al., 1956; 9u3w
Suzuki and Koyaguchi, 2009]
Note that the classic passive transport assumption of ADS
p MP 3/4 models for describing tephra dispersal should not be consid-
q = C k 5/8 (2) ered valid for timescales t < tb , whereas for tb < t < tp ,
N
both transport mechanisms are relevant. However, as dis-
where, from the results of Suzuki and Koyaguchi [2009] cussed previously, some ADS models appear to work well
 even for timescales t < tp because they use an effective diffu-
0.5  104 m3 kg–3/4 s–7/8 for tropical eruptions
C sion coefficient calibrated from observations rather than the
1.0  104 m3 kg–3/4 s–7/8 for midlatitude and polar eruptions
(3) actual atmospheric turbulent diffusion.
The relationship (1) derives from the conservation equation [7] Considering equations (4) and (5) and solving the
ordinary differential equation for ub (R) yields the following
d relationship between the front velocity and the radius of the
(R2 h) = q (4)
dt cloud:
 1/2
with the assumption that the velocity of the leading edge 2Nq 1
ub (R) = p (10)
of the spreading current, ub , scales with the average cloud 3 R
thickness, h, as
For practical purposes, this radial velocity field is considered
dR
ub = = Nh (5) only at distances less than Rp = R(tp ) and for heights between
dt H – h/2 and H + h/2, where H is the NBL height and h the
Combining (1) and (5), one obtains the radial velocity of the umbrella thickness. Within this region, the variation of the
umbrella spreading as a function of time: velocity field with distance r is given by
 1/3  
2 3Nq 3 R 1 r2
ub = t–1/3 (6) ub (r) = ub (R) 1+ (0 < r  R) (11)
3 2 4 r 3 R2
As the cloud entrains air, it becomes more dilute and, at a consistent with the equations above. Our strategy, therefore,
certain distance, atmospheric turbulence and wind advection consists of using equations (10) and (11) to compute a time-
transport mechanisms dominate cloud transport. In order dependent radial velocity field centered above the vent in the
to estimate the radial distance at which this critical transi- umbrella region. This radial velocity is added to the wind
tion between density-driven and passive transport occurs, we field furnished by the Numerical Weather Prediction Model
compare the umbrella front velocity ub with the mean wind to estimate the combined windfield for the ADS model
velocity uw at the NBL estimating the Richardson number: (FALL3D in our case). By doing this, ADS models can
 2/3 run with an “effective” velocity field accounting for contri-
u2b 4 3Nq
Ri = = t–2/3 (7) butions from both passive and density-driven mechanisms.
u2w 9u2w 2
Note that, depending on the balance between wind inten-
Atmospheric studies [Zilitinkevich et al., 2008] have shown sity and volumetric flow rate at the NBL, the added radial
that the interval 0.25 < Ri < 1 separates two different tur- velocity field can change the wind advection significantly.
bulent regimes (associated with strong and weak mixing, The box model given by equations (10) and (11) provides
respectively) rather than the turbulent and laminar regimes a reasonable approximation about the position of the flow
usually assumed for low and high Ri numbers. That does not front. Because it is based on the assumption of uniform cloud
support the existence of a critical Richardson number above thickness, the velocity field inside the cloud predicted by
which turbulent mixing is inhibited. Actually, experimental equation (11), although consistent with equations (4) and
and observational data indicate that turbulence survives for (5), should be viewed as a crude approximation of the real
Ri  1 [Galperin et al., 2007]. For example, in the free velocity field. Moreover, because the radial velocity (11)
atmosphere, where Ri typically varies from 1 to 100, signif- diverges as r ! 0, numerical calculationsp are truncated
icant turbulence has been observed at all levels [Lawrence assuming a minimum radius rmin = x2 + y2 , being
et al., 2004]. Similar observations are valid for the deep x and y the horizontal computational grid sizes. The
ocean [Galperin et al., 2007] and for dispersion of dense method of superposition of the radial flow of density cur-
CO2 clouds, where Cortis and Oldenburg [2009] found rent and the ambient wind field is suitable for large steady
that the threshold between passive and density-dominated eruption columns for which the wind velocity near the vent
regimes appears around values of Ri ' 0.25 (note that their is typically much smaller than the radial density current
Richardson number corresponds to the square root of the one velocity and the movement of the source position can be
used here), much larger than the value suggested by Britter assumed negligible.
2
COSTA ET AL.: DENSITY-DRIVEN TRANSPORT OF TEPHRA

Table 1. Parameters Used for the Simulation of the Volcanic Plume Evolution of the Climatic Phase of 15 June 1991
Pinatubo Eruption
FALL3D Parameters Values Notes
ı
Computational domain ( ) 18 (lat)30 (lon) -
Horizontal resolution (ı ) 0.15 -
Vertical resolution (m) 1000 -
Bottom-left corner coord. (lat; lon) (3.0; 98.0) -
Vent coord. (lat; lon) (15.133; 120.350) -
Vent elevation (m) 1500 -
TGSD  = 1.5/6.5;  = 1.55/1.55 Assumed(a)
Duration (h) 2.5 (C1)–4.5 (C2)–3 (C3) From Holasek et al. [1996a]
Average column height a.v.l. (km) 37 (C1)–31.5 (C2)–26 (C3) Observed(b)
Mass flow rate (109 kg/s) 1.5 (C1)–1 (C2)–0.3 (C3) Estimated(b)
Mixture exit velocity (m/s) 275 (C1)–225 (C2)–100 (C3) Estimated(b)
Mixture exit temperature (K) 1053 From Suzuki and Koyaguchi [2009]
Mixture exit water fraction (%) 6 From Suzuki and Koyaguchi [2009]
Total mass (1013 kg) 1.3 (C1)–1.6 (C2)–0.3 (C3) Computed
Empirical constant  (-) 0.2 From Suzuki and Koyaguchi [2009]
Brunt-Väisälä frequency N (s–1 ) 0.02 From Suzuki and Koyaguchi [2009]
Empirical constant C (m3 kg3/4 s7/8 ) 0.5  104 From Suzuki and Koyaguchi [2009]
NBL volumetric flow rate (1010 m3 /s) 12.9 (C1)–9.5 (C2)–3.7 (C3) Computed (c)
Meteorological set ECMWF ERA-Interim Rotated and rescaled(d)
a
Assumed as Bi-Gaussian Distribution similar to TGSD of other Plinian eruptions [Folch et al., 2009; Costa et al., 2012]; reported
values refer to the two means and variances of the distribution.
b
MFR and mixture exit parameters at the vent were fixed in order to reproduce the average column heights estimated from Figure 1
of Holasek et al. [1996a] using the BPT model coupled with the wind field [Bursik, 2001; Folch et al., 2012].
c
Computed in accord to equation (3). For comparison, Suzuki and Koyaguchi [2009] estimated q in the range 0.5–1.51011 m3 /s.
d
Original ERA-Interim wind fields were rotated 30ı anticlockwise around the vent and their intensities halved.

3. Application to 1991 Pinatubo Eruption Carazzo et al. [2006] and the crosswind entrainment coef-
ficient is estimated on the basis of the local Richardson
[8] Mount Pinatubo erupted early in the afternoon of 15 number [Folch et al., 2012]. This model allowed us to com-
June 1991, generating a Plinian column of 37–39 km high pute the MFR and the vertical distribution of mass along the
[Holasek et al., 1996a]. Satellite images revealed a giant column as function of the column height and mixture proper-
disk-shaped umbrella cloud that expanded radially for about ties at the vent (exit temperature, velocity, water content, and
5 h. The cloud extended up to about 140 km in radius cover- TGSD). In order to account for ash aggregation processes
ing an area of about 60,000 km2 by 14:40 Philippine Local [Costa et al., 2010; Folch et al., 2010], a simple aggre-
Time (PLT = UTC + 8). One hour later, at 15:40 PLT, gation model similar to that of Cornell et al. [1983] was
the cloud radius further expanded up to 200 km, cov- used [Costa et al., 2012]. Concerning meteorological data,
ering an area larger than 120,000 km2 [Koyaguchi, 1996; we used the European Centre for Medium-Range Weather
Holasek et al., 1996a]. At 19:40 PLT, the cloud reached a Forecasts (ECMWF) ERA-Interim reanalysis at 0.25ı reso-
stagnation point upwind but continued to grow downwind lution and 37 pressure levels (top at 1 hPa). Unfortunately,
[Koyaguchi, 1996; Holasek et al., 1996a]. Here we simulate reanalysis data were inconsistent with the stratospheric wind
the climatic phase of the eruption from 13:40 to 23:40 PLT intensity and direction suggested by cloud satellite images,
[Holasek et al., 1996a] using version 7.0 of the FALL3D giving a greater westward advection than observed in the
code (http://bsccase02.bsc.es/projects/fall3d/). For compu- satellite imagery [Holasek et al., 1996a; Koyaguchi, 1996].
tational reasons, we divided the climatic phase in three This discrepancy, already pointed out by Fero et al. [2009]
intervals named C1, C2, and C3, as follows: C1 from 13:40 also for the National Centers for Environmental Prediction
to 16:10 PLT with an average column height of about 37 reanalysis data set, can be attributed to the lack of regional
km above the vent level (a.v.l), C2 from 16:10 to 20:40 PLT meteorological observations during the pass of a typhoon
with an average column height of about 31.5 km a.v.l, and [Guo et al., 2004] and to the fact that no real-time atmo-
C3 from 20:40 to 23:40 PLT with an average column height spheric soundings were possible around Mount Pinatubo
of about 26 km a.v.l. [Holasek et al., 1996a]. during the eruption [Fero et al., 2009]. In order to match
[9] All the needed model parameters, given by Holasek satellite observations, we rotated the original ERA-Interim
et al. [1996a] and Suzuki and Koyaguchi [2009], are sum- reanalysis wind field 30ı anticlockwise around the vent and
marized in Table 1. For volcanological input parameters, halved the wind intensity.
FALL3D requires specification of the source term, i.e., the [10] Figure 1 compares observations of the time evolu-
vertical distribution of Mass Flow Rate (MFR) for each tion of the plume derived from Geostationary Meteorolog-
particle class, the column height, and the total grain size dis- ical Satellite (GMS) thermal infrared (IR) satellite images
tribution (TGSD). The source term was estimated by means [Holasek et al., 1996a] with the simulation results, with
of the eruption column model implemented in FALL3D and without considering the gravity current model. Clearly,
[Costa et al., 2006; Folch et al., 2009], based on the Buoy- the simulation considering the gravity current model at the
ant Plume Theory (BPT). The BPT plume model embedded NBL reproduces eruption cloud features much better than
in FALL3D 7.0 is similar to that of Bursik [2001], but that using atmospheric turbulent diffusion only, showing a
the radial entrainment coefficient is calculated similar to dominant effect of gravity current spreading both in cross
3
COSTA ET AL.: DENSITY-DRIVEN TRANSPORT OF TEPHRA

Figure 2. Dominant transport mechanism as function of


MFR and distance from the source. The colored area shows
the transition zone, delimited by Ri > 1 (density driven) and
Ri < 0.25 (passive), obtained using equations (1), (8), and
(9) assuming uw =15 m/s,  = 0.2, and N = 0.02. As ref-
erence, the upper x axis shows the corresponding column
height according to Mastin et al. [2009].

and downwind directions. Assuming that an integral column


load of 1 g/m2 represents the satellite detection threshold, the
coupled model is able to reproduce the volcanic cloud areas
with an average error of 15%, the crosswind axes with an
error of 5%, the downwind axes (that are more affected
by the wind field) with an error of 20%, and the cloud-
spreading velocity with an error of 5%. Using the values
estimated by Suzuki and Koyaguchi [2009] for the 1991
Pinatubo eruption (see Table 1) and considering an average
wind velocity of uw  15 m/s, from equations (8) and (9),
we obtain tb  5 h (corresponding to a radius of 450 km),
which is in good agreement with the observed time that the
cloud traveled upwind [Holasek et al., 1996a; Koyaguchi,
1996]. The passive dispersion approximation can be reason-
ably applied for times larger than tb , but it is fully valid only
for times larger than tp  40 h (corresponding to a radius
of 1800 km). This implies that for long-lasting eruptions
having an intensity similar to or larger than that of Pinatubo
1991, density-driven transport dominates for most of the
duration of the transport, whereas for smaller eruptions, the
Figure 1. Growth and movement of the plume on 15 June mechanism is relevant only in the very initial phase of the
1991 at the times reported (Philippine Local Time (PLT)). eruption (see Figure 2).
(a) Composite of GMS thermal IR satellite images (modi-
fied from Holasek et al. [1996a]), (b) simulated cloud PM10
4. Conclusions
column mass (vertical integration of concentration) coupling
FALL3D with the gravity current model. Contours are iso- [11] We analyzed the conditions under which the clas-
lines of 1 ton/km2 (i.e., 1 g/m2 ); (c) same simulation but sical passive transport assumption is valid for describing
without considering the gravity current model. tephra dispersal. An analytical model describing the spread-
ing of the cloud as a gravity current was coupled with an
advection-diffusion-sedimentation model. Conditions when
4
COSTA ET AL.: DENSITY-DRIVEN TRANSPORT OF TEPHRA

cloud spreading due to gravity current effects are dominant, Fero, J., S. Carey, and J. Merril (2009), Simulating the dispersal of
or conversely are negligible, were identified on the basis of tephra from the 1991 Pinatubo eruption: Implications for the formation
of widespread ash layers, J. Volcanol. Geotherm. Res., 186, 120–131,
the cloud Richardson number. The coupled model satisfacto- doi:10.1016/j.jvolgeores.2009.03.011.
rily reproduced the observed evolution of the volcanic cloud Folch, A. (2012), A review of tephra transport and dispersal models: Evolu-
during the climatic phase of the 1991 Pinatubo eruption. tion, current status, and future perspectives, J. Volcanol. Geotherm. Res.,
235-236, 96–115, doi:10.1016/j.jvolgeores.2012.05.020.
[12] Acknowledgments. This work has benefited from funding pro- Folch, A., A. Costa, and G. Macedonio (2009), FALL3D: A computational
vided by the Italian Presidenza del Consiglio dei Ministri - Dipartimento model for transport and deposition of volcanic ash, Comput. Geosci.,
della Protezione Civile (DPC), agreement INGV-DPC 2012-2013. This 6(1334–1342), 35, doi:10.1016/j.cageo.2008.08.008.
paper does not necessarily represent DPC official opinion and policies. Folch, A., A. Costa, A. Durant, and G. Macedonio (2010), A model for wet
A.F. acknowledges funding by the Spanish project ATMOST (CGL2009- aggregation of ash particles in volcanic plumes and clouds: II. Model
10244). ERA-Interim reanalysis data were provided by European Centre for application, J. Geophys. Res., 115, B09202, doi:10.1029/2009JB007176.
Medium-Range Weather Forecasts (ECMWF). We are grateful to review- Folch, A., A. Costa, and S. Basart (2012), Validation of the FALL3D ash
ers of the paper, T. Koyaguchi and J. Telling, for useful comments that dispersion model using observations of the 2010 Eyjafjallajokull vol-
improved the paper. We also thank S. Self, C. Connor, and C. Bonadonna canic ash cloud, Atmos. Environ., 165–183, 48, doi:10.1016/j.atmosenv.
for helpful suggestions on an early version of the paper. 2011.06.072.
[13] The Editor thanks Takehiro Koyaguchi and an anonymous Galperin, B., S. Sukoriansky, and P. Anderson (2007), On the critical
reviewer for assistance in evaluating this manuscript. Richardson number in stably stratified turbulence, Atmos. Sci. Lett., 8(3),
65–69, doi:10.1002/asl.153.
Guo, S., W. Rose, J. Bluth, and M. Watson (2004), Particles in the
great Pinatubo volcanic cloud of June 1991: The role of ice, Geochem.
References Geophys. Geosyst., 5, Q05003, doi:10.1029/2003GC000655.
Armienti, P., G. Macedonio, and M. Pareschi (1988), A numerical model Holasek, R., S. Self, and A. Woods (1996a), Satellite observations and inter-
for the simulation of tephra transport and deposition: Applications to pretation of the 1991 Mount Pinatubo eruption plumes, J. Geophys. Res.,
May 18, 1980 Mount St. Helens eruption, J. Geophys. Res., 93 (B6), 101(B12), 27,635–27,655.
6463–6476. Holasek, R., A. Woods, and S. Self (1996b), Experiments on gas-ash sep-
Baines, P., and R. Sparks (2005), Dynamics of giant volcanic ash aration processes in volcanic umbrella plumes, J. Volcanol. Geotherm.
clouds from supervolcanic eruptions, Geophys. Res. Lett., 32, L24808, Res., 70(3), 169–181.
doi:10.1029/2005GL024597. Koyaguchi, T. (1996), Volume estimation of tephra-fall deposits from the
Bonadonna, C., and J. Phillips (2003), Sedimentation from strong volcanic June 15, 1991, eruption of Mount Pinatubo by theoretical and geological
plumes, J. Geophys. Res., 108(B7), 2340, doi:10.1029/2002JB002034. methods, in Fire and Mud, Eruptions and Lahars of Mount Pinatubo,
Britter, R., and J. McQuaid, (1988), Workbook on the dispersion of dense Philippines, edited by C. Newhall, and R. Punongbayan, pp. 583–600,
gases, Tech. rep., HSE Contract Research Report No. 17/1988, Trinity University of Washington Press.
Road Bootle, Merseyside L20 3QY, U. K. Lawrence, J., M. Ashley, A. Tokovinin, and T. Travouillon (2004), Excep-
Bursik, M. (2001), Effect of wind on the rise height of volcanic plumes, tional astronomical seeing conditions above Dome C in Antarctica,
Geophys. Res. Lett., 18(28), 3621–3624. Nature, 431, 278–281.
Carazzo, G., E. Kaminski, and S. Tait (2006), The route to self-similarity in Mastin, L. G., et al. (2009), A multidisciplinary effort to assign realis-
turbulent jets and plumes, J. Fluid Mech., 547, 137–148. tic source parameters to models of volcanic ash-cloud transport and
Carey, S., and R. Sparks (1986), Quantitative models of the fallout and dispersion during eruptions, J. Volcanol. Geotherm. Res., 186, 10–21,
dispersal of tephra from volcanic eruption columns, Bull. Volcanol., 48, doi:10.1016/j.jvolgeores.2009.01.008.
109–125. Morton, B., G. Taylor, and J. Turner (1956), Turbulent gravitational convec-
Cornell, W., S. Carey, and H. Sigurdsson (1983), Computer simulation tion from maintained and instantaneous sources, Proc. Roy. Soc. London,
of transport and deposition of the Campanian Y-5 ash, J. Volcanol. Ser. A, 234, 1–23.
Geotherm. Res., 17, 89–109. Sparks, R., M. Bursik, S. Carey, J. Gilbert, L. Glaze, H. Sigurdsson, and
Cortis, A., and C. Oldenburg (2009), Short-range atmospheric dispersion of A. Woods (1997), Volcanic Plumes, pp. 574, John Wiley & Sons Ltd.,
carbon dioxide, Boundary Layer Meteorol., 133(1), 17–34, doi:10.1007/ Chichester, U. K.
s10546-009-9418-y. Suzuki, Y., and T. Koyaguchi (2009), A three-dimensional numerical sim-
Costa, A., G. Macedonio, and A. Folch (2006), A three-dimensional Eule- ulation of spreading umbrella clouds, J. Geophys. Res., 114, B03209,
rian model for transport and deposition of volcanic ashes, Earth Planet. doi:10.1029/2007JB005369.
Sci. Lett., 241, 634–647. Woods, A., and J. Kienle (1994), The dynamics and thermodynamics of
Costa, A., A. Folch, and G. Macedonio (2010), A model for wet aggregation volcanic clouds: Theory and observations from the April 15 and April
of ash particles in volcanic plumes and clouds: I. Theoretical formulation, 21, 1990 eruptions of Redoubt Volcano, Alaska, J. Volcanol. Geotherm.
J. Geophys. Res., 115, B09201, doi:10.1029/2009JB007175. Res., 62(1-4), 273–299.
Costa, A., A. Folch, G. Macedonio, B. Giaccio, R. Isaia, and V. Smith Zilitinkevich, S., T. Elperin, N. Kleeorin, I. Rogachevskii, I. Esau, T.
(2012), Quantifying volcanic ash dispersal and impact of the Mauritsen, and M. Miles (2008), Turbulence energetics in stably stratified
Campanian Ignimbrite super-eruption, Geophys. Res. Lett., 39, L10310, geophysical flows: Strong and weak mixing regimes, Q. J. R. Meteorol.
doi:10.1029/2012GL051605. Soc., 134(633), 793–799, doi:10.1002/qj.264.

También podría gustarte