Está en la página 1de 18

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, B11206, doi:10.

1029/2012JB009325, 2012

Insights into the mechanisms and timescales of pluton assembly


from deformation patterns of mafic enclaves
Luca Caricchi,1,2 Catherine Annen,1 Alison Rust,1 and Jon Blundy1
Received 20 March 2012; revised 25 September 2012; accepted 25 September 2012; published 15 November 2012.
[1] We propose a novel method to constrain both timescales and assembly styles of
intrusive bodies using the strain recorded by mafic enclaves, a common component of
granitic rocks. Petrology, thermal modeling, and magma rheology are combined to
investigate the evolution of strain recorded by enclaves during the piecemeal assembly of a
pluton growing at various rates of magma input. The different compositions (and hence
phase relations) of host magma and enclaves limits homogeneous deformation of these two
materials to restricted temperature ranges, which we term “windows of mutual
deformability.” Outside these windows only the less viscous host granite records any
appreciable deformation and enclaves are mainly transported as rigid objects. The temporal
and spatial development of the windows of mutual deformability reflects the emplacement
rate of magmas into the pluton. Consequently, the radial distribution of strain recorded
by enclaves can provide a picture of the thermal and rheological evolution of a magmatic
body during its construction. Our approach is applied to deformation patterns of mafic
enclaves in the Lago della Vacca Complex (LVC) of the Adamello Massif (Italy), to
estimate the timescales of pluton emplacement. Our calculations suggest total emplacement
timescales of the order 50 to 150 ky, in excellent agreement with recent high precision
radiometric dating of zircons from the LVC.
Citation: Caricchi, L., C. Annen, A. Rust, and J. Blundy (2012), Insights into the mechanisms and timescales of pluton assembly
from deformation patterns of mafic enclaves, J. Geophys. Res., 117, B11206, doi:10.1029/2012JB009325.

1. Introduction et al., 2009; de Saint Blanquat et al., 2011]. This further


implies that average, time-integrated magma injection rates,
[2] The rates of magma production and transfer from
determined by dividing the total volume of an intrusive body
depth control the dynamics of construction of the Earth’s by its range of ages, are generally lower than the actual
crust and potentially the frequency and magnitude of vol-
injection rates operating during individual pulses of mag-
canic eruptions at the surface. Detailed geological mapping,
matic activity [de Saint Blanquat et al., 2011]. The most
fabric and texture analysis give important insights into the
advanced techniques for radiometric dating give uncertain-
styles of magma intrusion in the crust and can be used to
ties of the order of 50 ka (this value is a function of the age
understand the processes that determine if magma is
of the rocks and their composition [e.g., Michel et al., 2008;
emplaced at depth, resulting in growth of a pluton, or if it is
Schaltegger et al., 2009; Leuthold et al., 2012]), which may
erupted at the surface. Detailed geological mapping and
not provide adequate resolution to estimate the injection rate
radiometric dating of intrusive bodies are often used to
of single magma pulses. In turn, this limits our ability to
constrain the rate of magma production in different tectonic
compare typical input rates for plutonic systems with those
settings [e.g., de Saint Blanquat et al., 2011]. Most intrusive currently obtained by, for example, inversion of geodetic
bodies display a range of radiometric ages that is signifi-
measurements in restless volcanic systems [e.g., Pritchard,
cantly larger than the time required for the cooling of a
2004; Biggs et al., 2010] and may jeopardize our chances
single pulse of magma equivalent in size to the entire pluton. of establishing a clear relationship between observed intru-
This observation strongly suggests a multistage, pulsed,
sion rates and the likelihood of an eruption. In fact, while
growth history for many plutons [Glazner et al., 2004; de
the long-term average injection rates control the thermal
Silva and Gosnold, 2007; Annen, 2009, 2011; DeCelles
evolution of magmatic systems [Annen, 2009, 2011], the
short-term injection rates exercise greater influence over
1
School of Earth Sciences, University of Bristol, Bristol, UK. the likelihood of magma storage versus magma eruption
2
Section des Sciences de la Terre et de l’Environnement, University of [Jellinek and DePaolo, 2003]. Because magma temperature
Geneva, Geneva, Switzerland. and rheology are strongly linked [Caricchi et al., 2007;
Corresponding author: L. Caricchi, School of Earth Sciences, Giordano et al., 2008], field mapping and detailed charac-
University of Bristol, Wills Memorial Building, Queen’s Road, terization of rock fabrics, in combination with thermal
Bristol BS2 1RJ, UK. (luca.caricchi@unige.ch) and rheological modeling, could potentially allow for a
©2012. American Geophysical Union. All Rights Reserved. more precise characterization of the typical rates of magma
0148-0227/12/2012JB009325

B11206 1 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 1. Field examples of enclaves from the Adamello massif, Italy. (a) Mingled enclaves with differ-
ent macroscopic textures. The picture was taken at Casinetto dei Dossi, Val Fredda Complex, Adamello
Massif. (b) Highly strained enclaves from the southern portion of the Lago della Vacca Complex
(LVC). This is the “highly deformed marginal unit” of John and Blundy [1993]. (c) Disaggregation of a
syn-plutonic dioritic dyke to produce a set of enclaves along its margins (Galliner Granodiorite unit of
LVC). (d) Foliation in tonalite deflected around an enclave. The picture was taken near Rifugio Tita Sec-
chi, Lago Della Vacca Complex, Adamello Massif.

injection in the crust and provide insights into the interpre- initial sphericity of the enclaves and the complex rheological
tation of geodetic and geophysical signals. history that enclaves experience during the accretion of intru-
sive bodies [Sparks and Marshall, 1986; Scaillet et al., 2000].
1.1. Mafic Enclaves Although there is considerable evidence against the notion of
[3] Fine-grained (or microgranular) mafic enclaves are a initial enclave sphericity [Paterson et al., 2004] (Figures 1a
ubiquitous feature of granitic (throughout this paper we use the and 1c), the different rheological evolution of host magma
terms granite and granitic in a general sense to describe any and enclaves during cooling [Sparks and Marshall, 1986;
felsic, plutonic rock, such as tonalite, syenite, granodiorite, or Scaillet et al., 2000] still affords an opportunity to use the
granite sensu stricto) plutons. Many such enclaves are frag- strain recorded by enclaves as a probe of the rheological, and
ments of less chemically evolved magma suspended in a more hence thermal, evolution of a pluton during its accretion.
evolved host [Vernon, 1984] (Figure 1). Their abundance in
the geological record testifies to the importance of magma 1.2. Conceptual Model
mingling in the assembly of plutons in the crust [Blundy and [4] Mafic enclaves are produced when hotter, more mafic
Sparks, 1992; Wiebe and Collins, 1998; Siegesmund and magma intrudes into cooler, more felsic magma and dis-
Becker, 2000; Turnbull et al., 2010] (Figure 1). Enclave aggregates [e.g., Turner and Campbell, 1986; Snyder and
shapes can be highly variable depending on the processes that Tait, 1995; Hodge et al., 2012]. Because mafic magmas
formed them and the contrast in viscosity with the host magma tend to have higher solidi temperatures, thermal equilibration
[Scaillet et al., 2000; Paterson et al., 2004] (Figure 1). The thermal equilibration initially results in significant crystalli-
contrast in composition between host and enclaves makes zation of the more mafic magma (Figure 2a), which, in turn,
them suitable candidates for studies of the dynamics of magma leads to a rheological inversion with the less chemically
emplacement [Holder, 1979; Ramsay, 1989; Molyneux and evolved magma becoming more viscous than its granitic host
Hutton, 2000; Siegesmund and Becker, 2000] and enclaves [Sparks and Marshall, 1986] (Figure 2b). The processes of
shapes have been used to constrain the amount of strain they genesis of mafic enclaves are not considered in detail here,
suffered during pluton emplacement [e.g., John and Blundy, and we assume that the new pulses of granitic magma enter
1993; Molyneux and Hutton, 2000]. This approach has been the magmatic reservoir with a suspended cargo of roughly
extensively debated and criticized [Siegesmund and Becker, spherical, thermally equilibrated enclaves. This last assump-
2000; Paterson et al., 2004], based on the assumption of tion is justified by considering that an approximate timescale

2 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

for thermal equilibration of an enclave of 10 cm radius is less [5] The repeated injection of magmatic pulses into a pluton
than 3 h (t = R2/K, where t is time, R is the radius of the through a common feeder or injection point, produces expan-
enclave, and K is the thermal diffusivity, considered here sion, and tends to deform the host magma in which the mafic
106 m2/s; this timescale is a minimum estimate because it enclaves are suspended. Enclaves suffer significant viscous
does not account for the release of latent heat of crystalliza- deformation (e.g., flattening) only if the ambient temperature
tion). This is significantly shorter than the timescales of is such that the viscosity of host magma and enclaves is similar
assembly of an intrusion. (Figure 2b). If the enclaves are significantly more crystalline

Figure 2

3 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

(i.e., more viscous) than their host, they will behave as rigid 1.3. The Lago Della Vacca Complex
objects in a deformable medium. Considering two magmas (Adamello Pluton, Italy)
with variation of crystallinity and viscosity as a function of [7] The LVC outcrops as a subcircular intrusion roughly
temperature as depicted in Figure 2, enclaves suspended in, 4.5 by 4.7 km in diameter ranging in composition from gabbros
and thermally equilibrated with, the host magma, will deform and diorites of the Blumone unit at the margins, through rela-
appreciably only in temperature intervals corresponding to the tively homogeneous tonalite (Vacca Tonalite) to granodiorite
near-liquidus and near-solidus regions of the host magma (Galliner Granodiorite) at the core (Figure 3). The exposed
where the viscosities of the two magmas are similar relief of this intrusion is about 2 km, which provides a lower
(Figure 2b). We define the regions where the contrast in vis- limit on its vertical extension [John and Blundy, 1993]. The
cosity between host magma and enclaves is less than 1 order of LVC is truncated by younger intrusions to the north and west
magnitude as “Windows of Mutual Deformability.” There are but intrudes and deforms the older Val Fredda Complex to the
two such windows, one near the liquidus (WMDL) and south [John and Blundy, 1993]. High-precision zircon U-Pb
another near the solidus temperature (WMDS; Figure 2b). If dating indicates a maximum emplacement time for this suite of
the intrusion is assembled by expansion related to magma less than 300ky [Schaltegger et al., 2009; Schoene et al., 2012].
injection at its core, Figure 2c shows how the aspect ratio of an Field relationships suggest that the emplacement style of this
initially spherical enclave might change with increasing dis- complex changed from passive stopping in the early stages of
tance from the injection point (and hence decreasing temper- intrusion of more mafic magma to later forceful intrusion of
ature) and assuming pure shear deformation. The amount of more felsic magma [John and Blundy, 1993; Brack, 1981]. The
flattening suffered by an individual enclave as it becomes magmatic foliation throughout the LVC is defined by planar
displaced away from the injection point is a function of the alignment of plagioclase, amphibole, and biotite crystals and
variation of temperature of the host during intrusion growth, mafic enclaves. The foliation lies roughly parallel to the mar-
intrusion geometry, and host-enclave contrast in composition. gins of the intrusion and the intensity of deformation increases
All of these factors affect the variation of crystallinity as a generally toward the contacts with the country rocks, although
function of temperature of both the host magma and the the increase in deformation from core to margin is nonlinear
enclaves (Figure 2). These considerations highlight how the and marked by a number of reversals [John and Blundy, 1993].
analysis of the strain suffered by enclaves located in different Measurements of strain retrieved by the analysis of enclave
portions of a magmatic body, can be used to probe the varia- shapes also testify to a general increase in the degree of defor-
tions of temperature and rheology during the assembly and mation toward the rims of the intrusion, but again with striking
growth of an intrusion [Holder, 1979; Ramsay, 1989]. Our reversals [John and Blundy, 1993] (Figure 3). Traces perpen-
approach differs from previous studies of enclaves as strain dicular to the magmatic foliation trend toward a common area
markers [e.g., Holder, 1979; Ramsay, 1989] in the explicit located at the NW end of the intrusion within the Galliner
treatment of temperature variations during pluton assembly Granodiorite (Figure 3). In this region the magmatic foliations
and their rheological consequences. Specifically, in our lose their coherency and the patterns defined are less regular and
approach enclaves only record significant strain within the two margin-parallel than elsewhere in the LVC [John and Blundy,
WMD identified above, rather than throughout the history of 1993]. For these reasons we consider this as the likely magma
pluton emplacement. As we shall show, this has important feeder injection point (Figure 3).
implications for the radial variation in the enclaves strain [8] On the basis of structural and geochronological studies
across an individual pluton. [John and Blundy, 1993; Schaltegger et al., 2009; Schoene
[6] To illustrate our approach, we use as an example of our et al., 2012], we hypothesize that the LVC was emplaced by
approach the Lago della Vacca Complex (LVC) in the episodic injection of magma near the region we identify as the
southern Adamello Massif (northern Italy). The advantage of injection point. With each successive injection the pluton
the LVC is that the strain recorded by enclaves in the intru- expanded radially to produce the foliation patterns and the
sion has been previously measured [John and Blundy, 1993] deformation of the enclaves observed in the field. Because of
(Figure 3) and the maximum time required for its construc- the strongly nonlinear increase in the flattening of enclaves
tion has been determined by high-precision zircon geochro- from the core to the rims of the LVC, we propose that the
nology [Schaltegger et al., 2009; Schoene et al., 2012]. thermal evolution of the pluton during its construction

Figure 2. Conceptual diagrams showing the influence of compositional contrast between host magma and enclaves on the
evolution of the strain ellipsoid recorded by mafic enclaves. (a) Variation of crystallinity as functions of temperature for a fel-
sic (host) and a mafic (enclave) magma. (b) Variation of viscosity (logarithmic scale) for the same magmas as functions of
temperature. The largest effect on viscosity is related to the increase of crystallinity with cooling. (c) Variation of the aspect
ratio (X/Z) of an enclave. The square symbol at the end of the dashed line marks the final position of the enclave in the intru-
sion. The temperature of the host magma decreases with increasing distance from the injection point. As enclaves and host are
assumed to be in thermal equilibrium this variation in temperature controls their relative crystallinities and viscosities as in
Figures 2a and 2b. The enclave is assumed to be initially spherical it is considered that the injection of magma results in
the expansion of the intrusion and deformation. At relatively high temperatures, near the injection point, both enclave and host
deform at near liquidus conditions. With increasing distance from the point of injection and decreasing temperature, the
enclaves stop deforming because of their significantly higher viscosity with respect to the host magma. Enclaves start deform-
ing again at near solidus conditions where their viscosity is similar to the viscosity of the host (Figure 2b). Between these two
extremes there is little change in enclave strain with distance.

4 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 3. Geological map of the intrusive units of the Lago della Vacca Complex (LVC) and the surround-
ing formations redrawn from John and Blundy [1993]. Also shown are foliation trajectories (dashed) and the
shapes of the strain ellipsoids measured by John and Blundy [1993]. The highly strained enclaves in the
“highly deformed marginal unit” may be attributed to its less chemically evolved nature (see text and
Figure 11) [John and Blundy, 1993]. The white circle indicates the tentative location of the point of magma
injection. This was identified using the point of convergence of lines perpendicular to the magmatic foliation.

exercised a fundamental control on the amount of deformation and lateral push by the magma itself [de Saint Blanquat et al.,
suffered by the enclaves in different portions of this complex. 2011]. The results of the present study are applicable to situa-
To quantitatively investigate these observations, we model tions in which magma is forcefully emplaced and creates space
cylinder-shaped intrusions growing radially by successive for its intrusion by displacing the crust [e.g., Molyneux and
injection of cylindrical pulses of magma into their core Hutton, 2000]. This scenario seems appropriate for relatively
(Figure 4a). This geometry approximates the structure and the small plutons assembled over a few hundreds of thousands of
modalities of magma injection of the LVC, and the emplace- years where postemplacement cooling is rapid, and tectonic
ment of laterally extended bodies of magma (i.e., width ≫ processes do not significantly overprint the magmatic fabrics
height), which are common elsewhere in the crust [de Saint [de Saint Blanquat et al., 2011].
Blanquat et al., 2011]. Successive pulses displace and stretch [10] The characterization of the evolution of deformation
earlier pulses as the pluton expands. Our calculations illustrate recorded by enclaves during forceful emplacement of an
how an accurate determination of the strain recorded by intrusive body requires the determination of the variations of
enclaves in different portions of a pluton, combined with crystallinity and residual melt composition as functions of
petrology, rheology, and thermal modeling, can elucidate the temperature for both host magma and enclaves, calculation
construction history of intrusive bodies such as the LVC. of the associated variation of viscosity for both magmas (i.e.,
crystal plus melt), and thermal modeling to trace the evolution
2. Methods of temperature of the magmatic pulses in time and space.
[9] Structural studies of several intrusions show that in many 2.1. Petrology
cases the space for the emplacement is created by a combination [11] We selected three chemical compositions for our
of roof uplift, downward translation of the base of the intrusion, study, a granodiorite and a tonalite as two possible host

5 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 4. (a) Schematic illustration of the model setup. Each successive pulse of magma occurs at the
core of the previous and displaces the previous injections at increasing distances from the core of the intru-
sion. The figure illustrates the evolution in time for selected pulses that are used throughout the paper
(I, V, X, XV, XX). (b) Sketches showing how expansion was tracked in the model, in order to calculate
the enclave strain using equations (6)–(9).

magmas and a diorite to represent the composition of enclaves residual melt composition during crystallization of granodi-
[e.g., Blundy and Sparks, 1992; Wagner et al., 2006; Johnson, orite and diorite exerts a second-order effect on the rheolog-
2003]. The variation of crystal content with temperature for ical properties of the host and enclave magmas [Giordano
these compositions under H2O-saturated conditions has been et al., 2008]. We make the simplifying assumption that the
constrained from literature data at typical pluton emplacement residual melt is always haplogranitic (containing 5 wt.%
depths of 5–10 km [Piwinskii and Wyllie, 1968, 1970; Martel water) for both compositions and its viscosity changes only
et al., 1999]. Notable features of these data are (1) the liquidus as function of temperature [Giordano et al., 2008]. The
temperature for the dioritic magma is about 100 C higher than evolution of magma viscosity as function of temperature and
for granodiorite (Figure 5a); (2) the crystallinity of the dioritic crystallinity were calculated using the model of Giordano
magma increases significantly between 900 and 850 C, at et al. [2008] for the residual melt and the algorithm of
which temperature the granodiorite lies close to its liquidus Caricchi et al. [2007] to account for the presence of crystals
(Figure 5a); and (3) the variation of crystallinity with temper- (Figure 5b).
ature for the tonalite is intermediate between the diorite and the
granodiorite (Figure 5a). These compositions were selected 2.3. Thermal Modeling
based on their similarity with the rocks present in the LVC; [13] Two sets of models were performed. In both cases
however, the variation of melt fraction as function of temper- successive pulses of magma are emplaced at the center of the
ature is controlled by several parameters, including chemical intrusion. In the first set, the pulses have infinite vertical
composition and volatile content, depth of magma emplace- extent and heat is released only from the lateral contacts with
ment, and oxygen fugacity [e.g., Martel et al., 1999; Piwinskii the country rock; in the second set the height of the magmatic
and Wyllie, 1968, 1970]. The important point is that the dif- body and the thickness of injections (H) is kept constant at
ferent crystal contents of host magma and enclaves as a func- 0.5, 1, and 2 km (Figure 4), centered at 4 km depth, and heat
tion of temperature control their respective rheologies and is lost through the wall rocks both along the vertical and
consequently the extent of deformation between them. horizontal coordinates of the numerical space. In this con-
figuration the roof of the intrusion is located at a depth of
2.2. Rheology 4 km  H/2 and the floor at 4 km + H/2 (Figure 6). The
[12] We first present discuss model results for the case of temperature evolution of the radially growing magmatic body
an exclusively granodioritic host with suspended dioritic (Figure 4) has been modeled using finite differences and
enclaves. We then show how a tonalitic host, chemically cylindrical coordinates. Except for simulations of infinite
closer to the enclave composition, reduces the rheological vertical extent, which were performed on a 1-D horizontal
contrast with the enclaves thereby changing their strain evo- cell array, temperatures and melt fractions were calculated on
lution during the expansion of the pluton. The variation of the a 2-D grid representing a vertical section through the system.

6 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 5. (a) Variation of crystal fraction and (b) logarithmic viscosity ratio for a granodiorite [Piwinskii
and Wyllie, 1968], a tonalite [Piwinskii and Wyllie, 1970] and a diorite/andesite [Martel et al., 1999], all
water-saturated, as a function of temperature, at constant pressure of 200 MPa. The numbers next to the
fitting lines give the sum of the squared residuals. WMDL and WMDS denote the two windows of mutual
deformability (for the granodiorite-diorite case only) that occur close to the granodiorite liquidus and sol-
idus, respectively. Between these two windows the diorite is substantially more viscous that its host
granodiorite.

The left boundary of the domain is the symmetry axis of the [14] The magma body grows in pulses with each new
intrusion (Figure 6). The horizontal dimension of the cylindrical pulse of radius (r) emplaced at the core of the
numerical domain is 10 km and, for 2-D simulations, the body; its volume is accommodated by radially displacing
vertical dimension of the numerical domain is from 0 (sur- previous magma pulses and the country rocks (Figure 4a)
face) to 8 km depth with magma cylinders injected with such that:
centers at 4 km depth. The boundary conditions of the qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
numerical domain are no heat flux through the left and right xnþ1 ¼ x2n þ r2 ð1Þ
boundaries. For 2-D simulations, the temperature of the
upper boundary (Earth’s surface) is fixed at 20 C. For the where xn+1 is the new position of a point on the numerical
lower boundary we calculate a heat flux qm assuming that the grid, xn is its former position, and r is the radius of the new
temperature corresponding to the liquidus TL of the host pulse causing this displacement. This type of displacement
magma (900 C for granodiorite) is located at a depth of conserves the volume of the successive pulses. We do not
45 km, so that qm = k(TL  Tn)/37,000 (37,000 is the differ- allow for contraction of magma upon solidification, which is
ence in meters between 45 km and the vertical extend of the of the order of 20% for typical and continuously degassing
numerical domain, which is 8 km), where k is the thermal hydrous magmas [John and Blundy, 1993].
conductivity and Tn the temperature of cells on the lower [15] Simulations consist of either 400 pulses of initial
boundary of the numerical domain. The simulations were radius of 200 m or 25 pulses of initial radius of 800 m
carried out with an initial geothermal gradient of 20 C/km. generating, in both cases a pluton with a final radius of 4 km

7 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 6. Setup of the thermal model, where q is heat flux and qm is calculated assuming a magma tem-
perature at depth (see text). T is temperature. For 1-D model, there is no vertical heat flux and the calcula-
tions are done on a horizontal array of cells. The symmetry is cylindrical. The cells and numerical domain
are not drawn to scale. The area within the dashed line shows the final volume of the pluton. The gray area
represents the last pulse of height (H) equal to that of the pluton. The enclave we consider in our calcula-
tion is located at half of the total high of a pulse (H/2) and at half of the pulse radius (r/2).

(see equation (1)). The injections of magma are assumed to be Xd at the different temperatures (the sum of the squared resi-
instantaneous (i.e., pulse injection is rapid relative to the duals are reported in Figure 5).
timescales of cooling). Calculations were carried out for dif- [16] The composition and properties of the country rocks
ferent time intervals between pulses (i.e., 0.6 to 60 ky for and the magma are identical: density 2500 kg/m3, heat
800 m pulse radius; 0.0375 to 3.75 ky for 200 m pulse radius) capacity of 1000 J/kg K, thermal conductivity of 2.5 W/m K,
resulting in different total emplacement durations (between and latent heat of 3.5  105 J/kg. For simplicity we kept these
15 and 1500 ky). Each new pulse is injected at the granodiorite parameters constant although we are aware that, in nature,
liquidus temperature of 900 C. Heat transfer is by conduction density, thermal conductivity and heat capacity vary with
and latent heat is produced by crystallization so that: temperature [Whittington et al., 2009]. We further neglected
convection and radioactive heat production within the
dT dX
rc  rL ¼ kr2 T ð2Þ growing pluton, although we implicitly require some local
dt dt convection at the point of injection in order to disperse the
where T is temperature, t is time, X is melt fraction, r is den- enclaves through the host granite magma at this location. The
sity, c is heat capacity, and k is thermal conductivity. Between results presented here track the evolution of the deformation
the solidus temperature (TS = 710 C, X = 0) and the liquidus of an enclave that begins at a radial position of r/2 from the
temperature (TL = 900 C, X = 1) the melt fraction-temperature center of the injected cylinder of magma (Figure 6). The
relationship of the host granite is based on experiments calculations are carried out for one enclave in each pulse.
[Piwinskii and Wyllie, 1968]. We used the following empirical
2.4. Strain During Pluton Assembly
fits to describe the variation of melt fraction in the granodiorite
(Xg) and diorite (Xd) as function of temperature (T; Figure 5a): [17] All calculations are performed by tracing the tem-
perature of the pulses at the half height of the intrusion (H/2;
1 Figure 6). The thermal modeling simulates the instantaneous
Xg ¼ 800T ð3Þ
1þe 23 injection of each magma pulse followed by different repose
times during which the temperature of the pulses decreases.
 B  B  Changing the size of the pulses (200 and 800 m radius) and
10K maintaining the total time required for the assembly of the
C  C
10K
Xd ¼ 0:5   B  B  ð4Þ pluton does not affect the long-term variation of temperature
2 10C
K
þ 10CK in the middle portion of each pulse with increasing distance
from the injection point (Figure 7). However, for the same
total time of assembly the time interval between two pulses
where K = T(in  C)/1000, B is 12.69, and C is 7.55. is longer for larger pulses so that:
These fitting parameters become 5.58 and 7.13, respectively,  2
for the tonalitic host magma that is considered in section 5. r2
t2 ¼ t1 ð5Þ
The equations were obtained by minimizing the residuals Xg or r1

8 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 7. Temperature distribution calculated as a function of distance from the injection point. Contin-
uous and dashed lines show the temperature evolution for five representative pulses of 800 and 200 m
radius, respectively. In order to facilitate direct comparison between these two scenarios the pulses were
chosen such that they correspond to identical distances from the injection point. For the smaller
(200 m) pulses this equates to greater number of injections than for larger (800 m) pulses. The numbers
next to the lines refer to the injection sequence; the first number refers to pulses of 800 m and second num-
ber to pulses of 200 m. Thus the temperature profile for the lowest trace in this figure is generated after one
pulse of 800 m or 16 pulses of 200 m.

with t2 and t1 being the time interval between pulses of initial where de(n) and de(n  1) are the distance of the enclave from
radius r2 (800 m in our simulations) and r1 (200 m in our the injection point at time step n and n  1, respectively. This
simulations), respectively. If pulses are large (and few), the calculation assumes that there is no shear localization or slip
system evolves in stepwise manner, while the injection of (e.g., faults) within or around the host magma. The calculated
more numerous smaller pulses simulates more closely a sce- strain of the host magma at a given position and time depends
nario where magma is continuously injected into the system only on its location (radial position) at the time in question.
(Figure 7). For larger pulses the time interval between intru- The simplified geometry of our model implies that deforma-
sions is longer, such that host magma and enclaves reach lower tion occurs in pure shear conditions.
temperatures before the following injection occurs (Figure 7). [19] The strain recorded by the enclaves (ɛe) differs from
This produces, in turn, differences in the strain recorded by the that of the host due to the difference in composition and
enclaves as a function of the pulse size. For example, for the consequently rheology between host magma and enclaves as
simulation shown in Figure 7, the continuous and dashed lines a function of temperature. Considering that the background
give the variation of temperature with distance for a pluton of stress during pluton growth is uniformly applied to both
2 km thickness, total time of pluton assembly of 150 ky, and enclave and host magmas for the same time (i.e., the ratio of
pulses with initial radii of 800 and 200 m, respectively. The the strain rates is proportional to the viscosity ratio of the two
data show that for a pulse of 800 m radius (continuous line) the magmas) ɛe was calculated as:
temperature at 400 m is lower than for the smaller pulse
(dashed line) at the moment of injection of the successive hg
ɛe ¼ ɛg ð7Þ
pulse (horizontal step in distance in Figure 7). This tends to hd
increases the amount of time that an enclave spends in WMDS
compared to pulses of 200 m radius (dashed black line in where hg and hd are the viscosities of granodiorite and diorite,
Figure 7) and consequently increases the amount of flattening respectively.
deformation that an enclave experiences. However, geometri- [20] Assuming constant volume for the enclaves during
cal aspects of our model may counteract this effect as deformation (i.e., no expulsion of any interstitial melt or
explained in section 3.3.1. volatiles) and an initial radius of the strain ellipsoid of 1, the
[18] The strain suffered by the host granodiorite (ɛg) was major (X) and the minor (Z) axes of oblate enclaves (Y is the
calculated considering progressive stretching (parallel to the intermediate axis) are horizontal in our model with X par-
circumference of the pluton) during pluton expansion allel to the rim of the pluton and Z along the radius of the
(Figure 4b): pluton. During deformation they vary as
de ðnÞ
ɛg ¼ 1 ð6Þ 1
de ðn  1Þ X ¼ 1 þ ɛe ; Z¼ ðX ¼ YÞ ð8Þ
X2

9 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 8

10 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

For prolate enclaves: from the injection point by subsequent magma pulses
rffiffiffiffi (Figure 4). The square symbols at the end of the X/Z trajec-
1 tories give the aspect ratio of the enclave at its final location
Z¼ ðZ ¼ Y Þ ð9Þ
X in the intrusion. It is important to notice that enclaves located
away from the middle of each pulse will follow different
In both cases the ratio X/Z gives the maximum aspect ratio thermal paths and attain aspect ratios similar but not identical
of the strain ellipsoid. Since in the LVC the majority of the to those of the enclaves considered here. This aspect will not
enclaves are oblate and also because our geometrical model be discussed in this contribution but could provide additional
results in pure flattening deformation, prolate enclaves will constraints on the thermal evolution of a pluton during
not be discussed further. The thermal modeling provides the emplacement.
evolution of temperature of the intrusion during expansion [23] An important feature of the X/Z versus distance dia-
and this allows us to calculate the evolution of crystallinity grams is that if the injection rate of magma is such that there
of both host and enclaves (equations (3) and (4)). We then is no significant temporal variation of temperature in the
calculate the respective viscosities and use equations (6)–(9) system, the X/Z patterns recorded by the enclaves in suc-
to calculate the evolution of the strain ellipsoid recorded in cessive injections follow exactly the same trajectories. This
the enclaves during pluton expansion. situation occurs in simulations performed with an infinite
vertical extent of the pulses and injections occurring every
3. Results 60 ky (Figures 9a and 9d). In this scenario each pulse cools
3.1. Thermal Modeling below solidus before the successive injection of magma
occurs and all the pulses enter the WMDS at similar distance
[21] The simulations were performed varying the initial and follow the same X/Z trajectory (Figures 9a and 9d). On
radius and the height of the pulses and the pluton. Fixing the the contrary, if the repeated injection of magma causes
average rate of magma injection (Qav) to be constant, an heating to a temperature above WMDS where no enclave
increase of r or H leads to a decrease in the frequency of the deformation occurs, then enclaves in successive pulses
injections ( f = 1/t, see also equation (5); Qav = pr2Hf ). In record less accumulated strain than previous ones at the
general, an increase of injection frequency or height of the same distance from the injection point (Figures 9b and 9e). If
pulses leads to higher temperature and to the presence of the temperature of the intrusion continues to increase, in the
higher melt fractions throughout the intrusion (Figure 8). If long term, enclave deformation may eventually occur in
injections (800 m radius) of granodioritic magma occur WMDL. If these conditions are met, the strains recorded by
every 60 ky, the amount of heat advected to the system is enclaves in the inner portions of the intrusion become
insufficient for the long-term generation of a partially molten greater even than those suffered by enclaves in peripheral
body of magma and all the pulses of magma cool to subsolidus portions of the intrusion, leading to a strain reversal close to
temperature prior to the following injection (Figure 9a). This is the pluton core (pulse XX in Figure 9f). In all cases the final
true even for simulations with a pluton of infinite vertical trajectories of X/Z are controlled by the thermal evolution
extension with no heat loss from roof and floor. Increasing the and the deformation geometry. Enclaves that describe gentle
frequency of injection (i.e., the net flux of magma) generates a increasing strain trajectories from the core to the rim of an
reservoir of partially molten material (Figure 8). If an injection intrusion are consistent with magma emplacement over a
of magma occurs every 6 ky (with a thickness of H = 2 km), a timespan in which the thermal state of the intrusion did not
reservoir of low crystallinity magma with a radius of about change significantly. Conversely, steeper gradients of
1 km is established and its dimensions remain almost constant increase of X/Z of enclaves with distance suggest tempera-
throughout the 150 ky of pluton construction (Figure 8). ture variations during assembly of the pluton related to
3.2. Deformation Associated With Pluton Growth progressive heating or cooling or changes in the rate or size
of magma pulses.
[22] The presence of two WMD implies that the evolution
of the strain ellipsoids recorded by the enclaves during the 3.3. Enclave Deformation Trajectories in Intrusions
growth of the intrusion is not solely related to subsolidus With Finite Vertical Extension
deformation [Holder, 1979; Ramsay, 1989]. Enclaves deform [24] Figure 10 shows the results from models in which
substantially only within the two WMD while at intermediate heat is also dissipated in the vertical direction. The addi-
temperatures the contrast in viscosity between host magma tional heat loss, with respect to simulations in which the
and enclaves is too large for significant deformation to occur vertical extension of the intrusion is infinite, results in faster
(Figure 2). Our results are presented using diagrams showing cooling, more time spent within the WMDS, and conse-
the X/Z aspect ratios for enclaves as a function of distance quently higher values of strain suffered by the enclaves
from the core of the intrusion, where injection occurs (e.g., (Figures 9 and 10). The deformation trajectories recorded by
Figure 9). These are the trajectories followed by the strain the enclaves, for a total injection time of 1500 ky (average
ellipsoid of enclaves as they are displaced radially outwards injection rate 1.7105–6.7105 km3/yr) are identical to

Figure 8. Distribution of temperature calculated as a function of depth and distance from injection point at the center of the
modeled space. Isotherms are shown as white lines labeled with temperature in  C. All the figures are symmetric with respect
to their left-hand side. Shown are (left) temperature distribution at incremental time steps (increasing from top to bottom) of
3.75 ky for a total emplacement time of 15 ky and (right) the temperature distribution for a total emplacement time of 150 ky
at intervals of 37.5 ky. The color coding represents the melt fraction distribution in the granodiorite in the modeled space
following the color key at the bottom.

11 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 9

12 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Figure 10. Evolution of the X/Z aspect ratio of the strain ellipsoids (logarithmic scale) for representative
simulations. The conventions followed are the same as used in Figure 9. The four panels on the left give
the results for 200 m radius pulses; the four panels on the right give the results for 800 m radius pulses.
The results for pulses of (top) 500 m and (bottom) 2000 m thickness, respectively, are also shown. The
time reported on top of each column represents the total time of intrusion. The average magma injection
rate (Qav) and the time interval between single injections (t) are also reported. Also shown are strain tra-
jectories followed by pulses I, V, X, XV, and XX for r = 800 m and corresponding 16th, 80th, 160th,
240th, and 320th pulse for r = 200 m.

those calculated in a model with no vertical heat loss A larger distance of the enclaves from the injection point
(Figures 9a and 9c) and are consequently not shown in reduces the total potential strain that enclaves can suffer
Figure 10. The variation of pulse radius and height, and (equations (6) and (7)). Conversely, the variation of injection
compositional contrast between host magma and enclaves, frequency associated with changing pulse radius can result in
all affect the final distribution of strain recorded by the different temporal evolution of temperature. Larger pulses
enclaves at different distances from the rim of the intrusion. (lower injection frequency) cool to lower temperature before
The effects of these parameters are discussed in below. the following magmatic pulse is emplaced (Figure 7), which
3.3.1. Effect of Initial Pulse Radius increases the potential for deformation within the WMDS. As
(r = 200 and r = 800 m) before, smaller and more frequent injection may increase the
[25] Variations of pulse radius produce two competing temperature in the proximity of the injection point to values
effects: (1) the distance between the injection point and the high enough for deformation in the WMDL to occur. Com-
initial location of the enclaves (in the middle of the pulses; paring the four panels on the left hand side of Figure 10 (pulse
Figure 6) increases with the radius of the pulse (100 and radius = 200 m) with those on the right-hand side (pulse
400 m for pulses of 200 and 800 m radius, respectively), radius = 800 m), it can be noticed that, except for the case
(2) maintaining constant the long-term rate of magma injection illustrated in Figures 10c and 9f, the main difference associ-
(Qav), the frequency of injection decreases with increasing ated with pulses of different radii is related to a larger amount
pulse radius (see equation (5); for a total assembly time of of flattening occurring in the WMDL when smaller pulses are
150 ky the time interval between injection increases from 375 injected frequently (Figure 10a). This indicates than in the case
to 6000 years as the pulse radius increases from 200 to 800 m). of small frequent injections the intrusion becomes hotter near

Figure 9. (a–c) Variation of temperature in the middle of the granodioritic pulses and (d–f) associated variation of the X/Z
aspect ratio (logarithmic scale) of the strain ellipsoids of the enclaves as a function of distance from the injection point for
three different time intervals between injections: 60 ky (Figures 9a and 9d), 12 ky (Figures 9b and 9e), and 6 ky (Figures 9c
and 9f). The models are run in one dimension and the intrusions are infinite in the vertical direction. All panels are symmetric
about their left-hand axis. Temperature and associated X/Z aspect ratios are illustrated for five successive different pulses,
labeled sequentially I, V, X, XV, and XX. The shaded areas in Figures 9a–9c represent the temperature regions within which
deformation occurs (WMDL and WMDS are the near-liquidus and near-solidus windows of mutual deformability, respec-
tively). The dashed lines represent the trajectories followed by an enclave during continuous expansion of the intrusion;
the filled square symbols give the final value of the aspect ratio X/Z. The colored ellipsoids at the top represent the calculated
strain ellipsoids of the enclaves, in plan view, relative to their final position, to show how radial distribution of enclave strain
is sensitive to emplacement frequency and duration.

13 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

(and pluton) thickness reduces the rate of cooling, leading to


an increase of temperature in the core of the intrusion. This
has a twofold effect: the deformation of the enclaves located
near the contacts with the wall rocks is lower if the pulses are
thicker and more voluminous (compare top and bottom dia-
grams in Figure 10) and the X/Z ratios of the enclaves are
larger in the core if the pulses are thicker. This last effect is
related to higher temperature in the inner parts of the pluton
and the higher amount of near-liquidus (WMDL) deforma-
tion (Figure 10).
3.3.3. Effect of Compositional Contrast
[28] The contrast in composition between host magma and
enclaves controls the variation of crystal content (and rhe-
ology) of the two magmas as a function of temperature
(Figure 5). Figure 11 compares the deformation trajectories
that the enclaves would follow if the host magma is grano-
dioritic versus tonalitic. No additional thermal modeling was
performed for the tonalitic host and the temperature-crystal
fraction relationship for granodiorite was substituted for that
of a tonalite (Figure 5). This implies that in both simulations
magma is injected at the same temperature of 900 C and that
the difference in latent heat released during cooling and
crystallization between granodiorite and tonalite is not con-
sidered. The tonalitic magma is characterized by a variation
of crystallinity as function of temperature close to that of the
diorite, while the difference is larger between granodiorite
and diorite (Figure 5). The dioritic enclaves suffer a signif-
icantly higher degree of deformation if they are suspended in
tonalitic magma because the rheological contrast between
these two magmas is much less than between a granodiorite
host and diorite enclaves (Figures 5 and 11). Variations in
composition of the enclave will also produce similar effects;
the closer enclaves are in composition to their host the
Figure 11. Variation of X/Z for diorite enclaves as a func- greater the degree of deformation that they will record for a
tion of distance from the injection point for a host magma of given set of emplacement parameters. Thus it is clear that an
(a) granodioritic and (b) tonalitic composition, respectively, accurate chemical characterization of host and enclaves is
where H = 2 km, r = 800 m, Qav = 2103 km3/yr, important for the quantitative application of our method.
t = 2000 years, and the total time of emplacement 50 ky.
4. Discussion
the injection point than for large, infrequent pulses and this
[29] Thermal and rheological modeling show that the fre-
accounts for the main difference in the X/Z patterns observed
in these two sets of models (Figure 10). To exclude the geo- quency of magma injection and intrusion thickness are
responsible for the production of substantially different X/Z
metrical consequences of different initial emplacement place
ratios of the enclave strain ellipsoid with increasing radial
of the enclaves (100 and 400 m for 200 and 800 m radius
pulses, respectively), the amount of flattening acquired by the distance from the injection point (Figures 9 and 10). These
variations are related primarily to the competition between
enclaves between 100 and 400 m from the injection point can
heat (magma) advected into the system and conductive
be subtracted. If deformation in the WMDL for the 200 m
cooling. Increasing the input of magma results in a general
radius pulses is not considered, the X/Z patterns for the two
increase of temperature of the magmatic system, which
sets of models are quite similar. This suggests that the tem-
enhances the rheological contrast between host magma and
perature variations in time associated to the variation of pulses
enclaves and may preclude significant enclave deformation
size are dominant with respect to the geometrical effects
in the WMDS (Figures 5, 9, and 10). It is important to note
associated with different initial distances of the enclaves from
that if the injections occur in a magmatic reservoir in which
the injection point.
the temperature remains approximately constant over time,
[26] The strain ellipsoids produced for different pulse radii
can also be used to visualize the range of X/Z associated with such that WMDL and WMDS are located at the same dis-
tance from the injection point, the strain ellipsoids recorded
an uncertainty or variations of the location of the injection
by the enclaves in different regions of the intrusion would
point of 300 m.
follow a single X/Z-position curve. Conversely, sudden
3.3.2. Effect of Pulse Height
drops of X/Z with increasing distance from the rim of an
[27] The thickness of the pulses is associated both with the
intrusion testify to variation in the position of the WMD and
volume of magma and heat advected into the system by sin-
consequently a variation in the thermal evolution of the
gle pulses and to the rate of cooling of the pluton from the
intrusion. This is illustrated by the 2-D models simulating
interior through its walls, floor, and roof. An increase of pulse

14 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

indicators of subsolidus deformation. It is important to note


that while the results presented here can only be applied to
intrusions with geometrical characteristics similar to those
modeled, more qualitative results on the thermal evolution of
an intrusion can be obtained applying equations (6)–(9) to
different pluton or injection geometries [e.g., Ramsay, 1989].
The only requirements are knowledge of the variation of
composition of the host magma and enclaves during cooling
and a basic structural study identifying the magma injection
region [e.g., Ramsay, 1989; John and Blundy, 1993]. Mod-
eling the position of the windows of mutual deformability
can be used to fit the X/Z ratio of enclaves and reconstruct the
size of individual pulses and the thermal history of an intru-
sion. Thermal modeling can then be performed to place time
constraints on pluton construction.
Figure 12. Variation of X/Z (logarithmic scale) of enclaves [32] In the following we show how, even within the geo-
as function of distance from the injection point (Figure 3) for the metrical limitations of our approach, and without taking into
enclaves measured by John and Blundy [1993] in the LVC (dia- account the potential effects of convection on cooling rates,
monds) and three different models with different total emplace- useful constraints can be placed on the timescales of pluton
ment times (indicated next to the lines). For all simulations construction from studies of mafic enclave deformation
H = 2 km and r = 800 m. Qav is 6.7105 km3/yr for the patterns in the LVC.
1500 ky model, 6.7104 km3/yr for an emplacement time
of 150 ky, and 2.01103 km3/yr for a duration of 50 ky.
The time interval between injections is 60, 6, and 2ky for a 5. Application: The LVC (Adamello Pluton, Italy)
total emplacement time of 1500, 150, and 50 ky, respectively. [33] Our models were performed using cylindrical geom-
The trend of the X/Z aspect ratio of the enclaves in the LVC etries whereas the LVC seems to have developed mainly as a
are better described by the models in which the total emplace- half cylinder extending eastward from the injection point
ment is 150 and 50 ky but not 1500 ky. These constraints on (Figure 3). This implies that the estimates we propose for
intrusion duration are consistent with recent zircon U-Pb data emplacement time are upper limits since they do not account
from the LVC. for additional heat loss to the northwest portion of the plu-
ton. To compare the modeling results with the field data we
pluton construction over 1500 and 300 ky. Longer assembly use the injection point, obtained from the structural investi-
times are related to lower rates of magma injection and in the gation of John and Blundy [1993] (Figure 3), to calculate the
1500 ky case injections occur only once the previous injec- strain at each location where strain ellipsoids of the enclaves
tion has cooled below its solidus temperature (Figures 9a were measured. John and Blundy [1993] performed between
and 9d). As a consequence, all magma pulses enter 16 and 90 measurements of the longest and shortest axes of
WMDS at about the same time (distance) from the injection the enclaves at each location using two or three orthogonal
point (Figure 9a). In the 300 ky case the overall increase of planes. From these data they calculated the principal axes of
temperature with time causes the pulses to enter the WMDS the strain ellipsoid representative of the enclaves population
at progressively shorter times (distances) after intrusion measured at each location. They found that the enclave were
(Figure 9b). As a consequence, the amount of deformation “weakly prolate to highly oblate.” For our comparison of
suffered by the enclaves at increasing distance from the these data with model results we use the near-horizontal
contact with the wall rock is lower than that calculated for a principal components of the strain ellipsoids determined
total assembly time of 1500 ky (Figures 9d and 9e). from the enclave shapes. We present calculations only for
[30] The clear differences in the X/Z ratios recorded by the oblate strain ellipsoids and consequently we do not include
enclaves for different injection rates suggest that there is prolate enclave data of John and Blundy [1993]. The method
considerable potential to invert data on enclave strain in can also be applied to prolate enclaves using equation (9)
different portions of a pluton to retrieve information on the instead of equation (8). Further details of the enclave mea-
thermal history of an intrusion, the evolution of magma surements and associated uncertainties are given in John and
rheology during pluton assembly and hence the rates of Blundy [1993]. The uncertainties related to the simplified
magma injections. The analysis performed in this study is geometry of our models are larger than those reported for the
concerned with enclaves of a single composition; we notice field measurements.
that simultaneous strain analysis of enclaves of different [34] The strain patterns defined by the enclaves in differ-
bulk compositions, including perhaps melt-free country rock ent portions of the LVC differ significantly from the patterns
xenoliths, would increase the sensitivity of our proposed expected if deformation had occurred exclusively at sub-
approach. In fact, country rock xenoliths should only deform solidus conditions in a total emplacement time of 1500 ky
at temperatures near or below the solidus of the host magma, (Figure 12). Specifically, enclaves are much less deformed
which would permit one to determine whether deformation than would be expected for subsolidus deformation. All the
of nearby enclaves occurred exclusively subsolidus or within model simulations we performed for a total emplacement
both the WMDL and WMDS. time of 1500 ky and different thicknesses of the intruding
[31] Additional analysis of deformation textures in miner- pulses fail to reproduce the X/Z ratios of the enclaves in the
als contained in host magma and enclaves can also be used as LVC (Figures 9a and 9d; Figure 12). This strongly suggests

15 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

(Figure 12). We performed additional simulations for a total


emplacement time of 50 ky and a thickness of 2 km (as
estimated by John and Blundy [1993]) to further constrain
the total emplacement time (Figure 12). This simulation
produce the flat X/Z trends observed in the inner portion of
the LVC even thought the X/Z in this portion of the pluton
are now higher than those measured in the field (Figures 12).
This is related to higher deformation at near-liquidus (WMDL)
conditions with respect to the simulation for a total injection
time of 150 ky. These results would consequently indicate that
the total emplacement time for the LVC is between 150 and
50 ky. Assuming a cylindrical shape for the LVC, with a diameter
of 4.5 km and a thickness of 2 km, the average magma
injection rate would lie between 2104 and 6104 km3/yr,
which is comparable to other intrusive bodies [de Saint
Blanquat et al., 2011]. These results were obtained consid-
ering the host magma to be granodioritic; however, the LVC
is characterized by a progressive evolution in chemistry from
more tonalitic to more granodioritic toward the core. Taking
into account this compositional variation would improve the
quality of the match between model and the measured X/Z
(Figure 11). Additionally, the chemical composition of the
“highly deformed marginal unit” is closer to that of the
enclaves than to the composition of the “marginal unit” [John
and Blundy, 1993] (Figure 3). This compositional similarity
provides a simple explanation for the reversals in X/Z that are
observed between these units (Figures 3, 5, and 11) [John and
Blundy, 1993]. However, to fully account for the effects of
compositional contrast between enclaves and host, a detailed
investigation of the chemistries of host and enclaves at dif-
ferent distances from the outer margin of the intrusion are
required.
[36] The nearly constant X/Z ratios measured from the
inferred injection point to about 2 km distance, seem to be
indicative of temperatures between 900 and 750 C (tem-
perature at which the enclaves do not deform) in this portion
Figure 13. Variation of X/Z (logarithmic scale) of enclaves of the intrusion (Figures 2, 5, and 12). Flat X/Z trends in the
as function of distance from the injection point for a set of internal portions of the intrusion could also be generated by
models performed with a constant r of 800 m and different progressive cooling with the WMDS moving progressively
ranges of Qav (reported in the figure). Qav was varied closer to the injection point [e.g., Ramsay, 1989]. However,
changing the thickness of the pulses and the total duration both the microstructures, which are suggestive of super-
of emplacement. Increasing Qav produces, first, an increase solidus deformation, and the occurrence of several examples
of the gradient of the X/Z aspect ratios of the enclaves with of foliation wrapping around enclaves in the core of the
increasing distance from the injection point (Figure 13b) and intrusion (e.g., Figure 1d) suggest that the pluton was still
finally leads to X/Z reversals and flat patterns of the X/Z heating up before the final episode of magma emplacement
ratio (Figure 13c). in the crust. This suggests that toward the end of the
assembly of the LVC, a magmatic body (containing poten-
tially eruptible magma) of about 2 km diameter with tem-
that the time for assembly of the LVC is less than 1500 ky, peratures between 900 and 750 C, and a thickness of about
in agreement with the geochronological data [Schaltegger 2 km (more than 6 km3) was present in the crust. However, it
et al., 2009; Schoene et al., 2012]. The calculations sug- is not know if the LVC, or any other units of the Adamello
gest also that if the total time of emplacement is of the order Massif, ever produced any volcanic eruptions.
of 300 ky, the X/Z trajectories produced by the calculations [37] Our results can be used in more qualitative terms to
do not match field measurements, especially in the inner determine the thermal and rheological evolution of a pluton
portions of the intrusion where the calculated flattening is during its assembly (Figure 13). Higher rate of magma
significantly smaller than that measured by John and Blundy injection tend to produce steeper variations of the X/Z aspect
[1993] (Figure 9e). These simulations also do not reproduce ratio with distance from the injection point (Figure 13b).
the flat trend of X/Z with distance that is distinctive in the However, very high average rates of magma injection may
inner portion of the LVC (Figures 3 and 12). generate X/Z reversals and flat X/Z trends if there is signif-
[35] The strains calculated from the 3-D simulations for a icant deformation in the WMDL (Figure 13c). Such flat
total intrusion time of 150 ky and a pulse thickness of 2 km trends have been related to a decreasing rate of magma input
give results that are comparable to the field measurements (WMDS progressively closer to the injection point) with

16 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

time [Holder, 1979; Ramsay, 1989] but are more likely Blundy, J., and R. Sparks (1992), Petrogenesis of mafic inclusions in gran-
produced by increasing temperatures during pluton con- itoids of the Adamello Massif, Italy, J. Petrol., 33(5), 1039–1104.
Brack, P. (1981), Structures in the southwestern border of the Adamello
struction. In fact, the first hypothesis implies that the rate at intrusion (Alpi Bresciane, Italy), Schweiz. Mineral. Petrogr. Mitt., 61,
which the WMDS move closer to the point of magma input 37–50.
is constant, while progressive heating results in flat X/Z ratio Caricchi, L., L. Burlini, P. Ulmer, T. Gerya, M. Vassalli, and P. Papale
(2007), Non-Newtonian rheology of crystal-bearing magmas and implica-
of enclaves if the temperature reaches a broad interval tions for magma ascent dynamics, Earth Planet. Sci. Lett., 264, 402–419,
between the two WMD (Figure 2). doi:10.1016/j.epsl.2007.09.032.
DeCelles, P. G., M. N. Ducea, P. Kapp, and G. Zandt (2009), Cyclicity in
6. Conclusions Cordilleran orogenic systems, Nat. Geosci., 2(4), 251–257, doi:10.1038/
ngeo469.
[38] We have developed a thermomechanical model for de Saint Blanquat, M., E. Horsman, G. Habert, S. Morgan, O. Vanderhaeghe,
R. Law, and B. Tikoff (2011), Multiscale magmatic cyclicity, duration of
the development of magmatic strain in a growing granitic pluton construction, and the paradoxical relationship between tectonism
pluton as recorded by deformed mafic enclaves. In contrast and plutonism in continental arcs, Tectonophysics, 500, 20–33,
to previous treatments we take explicit account of the rheo- doi:10.1016/j.tecto.2009.12.009.
de Silva, S. L., and W. D. Gosnold (2007), Episodic construction of batho-
logical differences between mafic enclaves and their host liths: Insights from the spatiotemporal development of an ignimbrite
granitic magma as a function of temperature. We have flare-up, J. Volcanol. Geotherm. Res., 167, 320–335, doi:10.1016/j.
identified two windows of mutual deformability (WMD) jvolgeores.2007.07.015.
within which enclaves and the host deform homogeneously. Giordano, D., J. K. Russell, and D. B. Dingwell (2008), Viscosity of
magmatic liquids: A model, Earth Planet. Sci. Lett., 271, 123–134,
By comparing our modeled strain trajectories as a function doi:10.1016/j.epsl.2008.03.038.
of time and distance from the magma injection point, we Glazner, A., J. Bartley, D. Coleman, W. Gray, and R. Taylor (2004), Are
show that it is possible to constrain the emplacement time of plutons assembled over millions of years by amalgamation from small
magma chambers?, GSA Today, 14, 4–11, doi:10.1130/1052-5173(2004)
granitic plutons. 014<0004:APAOMO>2.0.CO;2.
[39] We have applied our results to data on strained mafic Hodge, K. F., G. Carazzo, and A. M. Jellinek (2012), Experimental con-
enclaves in the Lago della Vacca Complex in the southern straints on the deformation and breakup of injected magma, Earth Planet.
Adamello Massif to show that this pluton was emplaced over Sci. Lett., 325–326, 52–62, doi:10.1016/j.epsl.2012.01.031.
Holder, M. (1979), An emplacement mechanism for post-tectonic granites
a 50–150 ky interval, in remarkable agreement with recent and its implications for their geochemical features, in Origin of Granite
geochronological estimates (about 300 ky) [Schoene et al., Batholiths, edited by M. Atherton and J. Tarney, pp. 116–128, Birkhau-
2012]. Depending on the 3-D geometry of the pluton this ser, Boston, Mass., doi:10.1007/978-1-4684-0570-5_10.
Jellinek, A., and D. DePaolo (2003), A model for the origin of large silicic
equates to average magma injection rates between 2104 magma chambers: Precursors of caldera-forming eruptions, Bull. Volcanol.,
and 6104 km3/yr. 65(5), 363–381, doi:10.1007/s00445-003-0277-y.
[40] The simultaneous strain analysis of enclaves of dif- John, B., and J. Blundy (1993), Emplacement-related deformation of granitoid
magmas, southern Adamello Massif, Italy, Geol. Soc. Am. Bull., 105(12),
ferent compositions (i.e., different WMD) in different por- 1517–1541, doi:10.1130/0016-7606(1993)105<1517:ERDOGM>2.3.CO;2.
tions of a pluton, and a more accurate characterization of the Johnson, S. (2003), Structure, emplacement and lateral expansion of the San
strain (to avoid the biases related to the assumption of initial José tonalite pluton, Peninsular Ranges batholith, Baja California, México,
sphericity of the enclaves) would increase the accuracy of J. Struct. Geol., 25(11), 1933–1957, doi:10.1016/S0191-8141(03)00015-4.
Leuthold, J., O. Müntener, L. P. Baumgartner, B. Putlitz, M. Ovtcharova,
our approach. Additionally, more accurate data on the evo- and U. Schaltegger (2012), Time resolved construction of a bimodal lacco-
lution of melt fraction as function of temperature for differ- lith (Torres del Paine, Patagonia), Earth Planet. Sci. Lett., 325–326, 1–8,
ent magma compositions are required to define accurately doi:10.1016/j.epsl.2012.01.032.
Martel, C., M. Pichavant, F. Holtz, B. Scaillet, J. Bourdier, and H. Traineau
the windows of mutual deformability. While the analysis we (1999), Effects of f(O2) and H2O on andesite phase relations between
propose can be improved, this approach, in combination 2 and 4 kbar, J. Geophys. Res., 104, 29,453–29,470, doi:10.1029/
with other techniques, can be applied to other plutons and 1999JB900191.
Michel, J., L. Baumgartner, B. Putlitz, U. Schaltegger, and M. Ovtcharova
provide constraints on the mechanisms and timescales of (2008), Incremental growth of the Patagonian Torres del Paine laccolith
emplacement of magma in the crust. over 90 k.y., Geology, 36(6), 459–462, doi:10.1130/G24546A.1.
Molyneux, S., and D. Hutton (2000), Evidence for significant granite space
[41] Acknowledgments. We would like to thank Simon Wallis for creation by the ballooning mechanism: The example of the Ardara pluton,
fruitful discussions in the field. L.C. is grateful for the support provided Ireland, Geol. Soc. Am. Bull., 112(10), 1543–1558, doi:10.1130/0016-
by the NERC grant (NE/G012946/1). AR acknowledges the financial sup- 7606(2000)112<1543:EFSGSC>2.0.CO;2.
port given by the Royal Society (URF fellowship). C.A. was funded by Paterson, S., G. Pignotta, and R. Vernon (2004), The significance of
the ‘VOLDIES’ ERC Advanced Grant. J.B. was supported through the microgranitoid enclave shapes and orientations, J. Struct. Geol., 26(8),
‘CRITMAG’ ERC Advanced Grant. The manuscript benefited from thor- 1465–1481, doi:10.1016/j.jsg.2003.08.013.
ough reviews by Mark Jellinek, Othmar Müntener, and the Associate Editor Piwinskii, A., and P. Wyllie (1968), Experimental studies of igneous rock
Mark D. Behn. series—A zoned pluton in Wallowa Batholith Oregon, J. Geol., 76(2),
205–234, doi:10.1086/627323.
Piwinskii, A., and P. Wyllie (1970), Experimental studies of igneous rock
References series. Felsic body suite from Needle Point Pluton, Wallowa-Batholith,
Annen, C. (2009), From plutons to magma chambers: Thermal constraints Oregon, J. Geol., 78(1), 52–76, doi:10.1086/627488.
on the accumulation of eruptible silicic magma in the upper crust, Earth Pritchard, M. E. (2004), An InSAR-based survey of volcanic deformation in
Planet. Sci. Lett., 284, 409–416, doi:10.1016/j.epsl.2009.05.006. the southern Andes, Geophys. Res. Lett., 31, L15610, doi:10.1029/
Annen, C. (2011), Implications of incremental emplacement of magma bod- 2004GL020545.
ies for magma differentiation, thermal aureole dimensions and plutonism- Ramsay, J. (1989), Emplacement kinematics of a granite diapir—The Chin-
volcanism relationships, Tectonophysics, 500, 3–10, doi:10.1016/j. damora Batholith, Zimbabwe, J. Struct. Geol., 11, 191–209, doi:10.1016/
tecto.2009.04.010. 0191-8141(89)90043-6.
Biggs, J., P. Mothes, M. Ruiz, F. Amelung, T. H. Dixon, S. Baker, and S.-H. Scaillet, B., A. Whittington, C. Martel, H. Pichavant, and F. Holtz (2000),
Hong (2010), Stratovolcano growth by co-eruptive intrusion: The 2008 Phase equilibrium constraints on the viscosity of silicic magmas II: Impli-
eruption of Tungurahua Ecuador, Geophys. Res. Lett., 37, L21302, cations for mafic-silicic mixing processes, Trans. R. Soc. Edinburgh
doi:10.1029/2010GL044942. Earth, 91, 61–72.

17 of 18
B11206 CARICCHI ET AL.: ENCLAVES B11206

Schaltegger, U., P. Brack, M. Ovtcharova, I. Peytcheva, B. Schoene, Turnbull, R., S. Weaver, A. Tulloch, J. Cole, M. Handler, and T. Ireland
A. Stracke, M. Marocchi, and G. M. Bargossi (2009), Zircon and titanite (2010), Field and geochemical constraints on mafic-felsic interactions, and
recording 1.5 million years of magma accretion, crystallization and initial processes in high-level arc magma chambers: An example from the Half-
cooling in a composite pluton (southern Adamello batholith, northern moon Pluton, New Zealand, J. Petrol., 51(7), 1477–1505, doi:10.1093/
Italy), Earth Planet. Sci. Lett., 286(1–2), 208–218, doi:10.1016/j. petrology/egq026.
epsl.2009.06.028. Turner, J., and I. Campbell (1986), Convection and mixing in magma cham-
Schoene, B., U. Schaltegger, P. Brack, C. Latkoczy, A. Stracke, and bers, Earth Sci. Rev., 4, 255–352, doi:10.1016/0012-8252(86)90015-2.
D. Gunther (2012), Rates of magma differentiation and emplacement in Vernon, R. (1984), Microgranitoid enclaves in granites - Globules of hybrid
a ballooning pluton recorded by U-Pb TIMS-TEA, Adamello batholith, magma quenched in a plutonic environment, Nature, 309(5967), 438–439,
Italy, Earth Planet. Sci. Lett., 355-356, 162–173, doi:10.1016/j.epsl. doi:10.1038/309438a0.
2012.08.019. Wagner, R., C. Rosenberg, M. Handy, C. Mobus, and M. Albertz (2006),
Siegesmund, S., and J. Becker (2000), Emplacement of the Ardara pluton Fracture-driven intrusion and upwelling of a mid-crustal pluton fed from
(Ireland): New constraints from magnetic fabrics, rock fabrics and age a transpressive shear zone—The Rieserferner Pluton (Eastern Alps),
dating, Int. J. Earth Sci., 89(2), 307–327, doi:10.1007/s005310000088. Geol. Soc. Am. Bull., 118, 219–237, doi:10.1130/B25842.1.
Snyder, D., and S. Tait (1995), Replenishment of magma chambers: Whittington, A. G., A. M. Hofmeister, and P. I. Nabelek (2009), Tempera-
Comparison of fluid-mechanic experiments with field relations, Contrib. ture-dependent thermal diffusivity of the Earth’s crust and implications
Mineral. Petrol., 122(3), 230–240, doi:10.1007/s004100050123. for magmatism, Nature, 458(7236), 319–321, doi:10.1038/nature07818.
Sparks, R., and L. Marshall (1986), Thermal and mechanical constraints on Wiebe, R., and W. Collins (1998), Depositional features and stratigraphic
mixing between mafic and silicic magmas, J. Volcanol. Geotherm. Res., sections in granitic plutons: Implications for the emplacement and crystal-
29, 99–124, doi:10.1016/0377-0273(86)90041-7. lization of granitic magma, J. Struct. Geol., 20, 1273–1289, doi:10.1016/
S0191-8141(98)00059-5.

18 of 18

También podría gustarte