A fracture-loop thermal balance model of black smoker circulation

A fracture-loop thermal balance model of black smoker circulation

Tectonophysics, 122 (1986) 307 -324 Elsevier Science Publishers 307 B.V., Amsterdam A FRACTURE-LOOP - Printed THEXMAL in The Netherlands BALANC...

1MB Sizes 0 Downloads 14 Views

Tectonophysics, 122 (1986) 307 -324 Elsevier Science Publishers

307

B.V., Amsterdam

A FRACTURE-LOOP

- Printed

THEXMAL

in The Netherlands

BALANCE

MODEL

OF BLACK

SMOKER

CIRCULATION

M.R. STRENS

and J.R. CANN

of Geology, The Unk.wrsity, Newcastle tcpon Tjme, NEI 7RU (Great Britain]

Department (Received

February

25, 1985; revised version

accepted

July, 1985)

ABSTRACT

Strens.

M.R. and Cann,

J.R., 1986. A fracture-loop

thermal

balance

model of black smoker

circulation.

Tectonoph_wirs, 122 : 307-324. This

time-dependent

circulation

during

model

suppIy to black smoker Flow is modelled mid-ocean systems

systems

location.

turbulent

flow, which

to flow is P~~na~tIy

digitally,

calculated throughout

is driven

conductive

is based

on seafloor

by the buoyancy heat

flux from

of a hot spring,

Results show that this system cannot

recently-solidified particularly

Jf the circulation

thus provides

magma

favourable.

chamber,

We conclude

penetrates

and ophiolite between

upflow

tbe rock is computed

enabling obtain

magma

zone. The fault is allowed

at the rock-water

interface

through

hydrot~~ma~ the major heat

chamber. and very close to the

evidence

of black

and downflow. to propagate

analyticalfy.

smoker

temperature

Hydraulic downwards.

with the water

at every position

the exit water

the resulting sufficient

and

whether

in the model by a series of pipes carrying

difference

in the discharge

for each time step. The model

smokers.

hot rocks

in a single fault, which is parallel

flow

in the loop can be and mass flow rate

ore mass to be estimated.

heat from a thin layer of rock (such as the 1-2

km thick layer of Iavas and dykes over@ing an axial magma black

between

in order to establish

The fault is represented

so that the heat exchange

the lifespan

balance

ore deposits,

comes from the rock or from an underlying

ridge axis. This configuration and their structural

The horizontal

the thermal

of volcanogenic

as an open thermosyphon

resistance treated

examines

the formation

chamber)

a greater

ore masses of a few million tonnes that systems producing

to sustain tong-lived, ore-forming

volume

of hot rock can be formed

to the base of a if conditions

larger ore masses must tap a magmatic

are heat

supply.

The model we present here is one of a number of thermal balance models, each designed to investigate the exchange of heat between rocks and circulating hydrothermal solutions during the formation of volcanogenic ore deposits. In this paper we examine one aspect of the thermat balance of ocean-floor high-temperature hydrothermal circulation, such as gives rise to the black smoker hot springs of the East Pacific Rise. In these black smoker systems, the overall heat

30X

flux is many segment

times greater

of mid-ocean

than

is available

ridge (Macdonald

either in hot rock or in magma

on a steady-state

until it is tapped

by the hydrothermal

Here we test the efficacy of hot rock as a heat store, and extract rock into water

circulating

in a single

basis from a short

et al., 1980). Heat must therefore

fault parallel

be stored circulation.

the heat from the

to the mid-ocean

ridge, The

water flow is confined to a fault because of geological constraints which are discussed below. Our target is to produce a high-temperature (350°C) water flow for long enough to generate a sulphide deposit of 3 million tonnes on the ocean floor. Other models are possible. We have developed one in which heat is extracted from a magma chamber, as an alternative to heat extraction from rock (Cann et al., in press), and have also examined a third model in which recharge is through a porous medium rather than in a fracture. This model may be more applicable to systems with a greater area of recharge

THE CHARACTERISTICS

than seems to apply to black smoker systems.

OF BLACK SMOKER

CIRCULATION

We define a black smoker system from the characteristics of the hot vents on the East Pacific Rise, where, at 21 “N, the exit temperature of the hydrothermal fluid has been estimated at about 35O”C, with an exit velocity of 1-5 m s-’ (RISE, 1980, Macdonald et al., 1980), and, subsequently in the range 275”-35O*C, with flow velocities of 0.7-2.4 m s-’ and a mass flow rate of about 150 kg s ’ (Converse et al., 1984). At 13”N the exit temperature of fluid partially diluted with seawater is up to 33O”C, so the maximum fluid temperature will be somewhat higher. The flow velocity at 13”N is estimated at 0.5-2 m s -’ (Hekinian et al., 1983, 1984). At both locations the iron concentration in the hydrothermal fluid is about 110 ppm (Edmond, 1981; Michard et al., 1984). We assume that such systems are broadly analogous to the systems responsible for the deposition of the volcanogenic sulphide ore bodies found in ophiolite suites, although the known black smokers are thought to be very young and are associated with very small sulphide

deposits.

The efficiency of ore deposition is very variable between the observed vents. The black smokers themselves are extremely inefficient, since a high proportion of the sulphides is carried away by ocean currents. In contrast the warm vents on the Galapagos Spreading Centre are highly efficient, since sub-surface mixing with cofd water causes deposition of a high proportion of the dissolved metals as sulphides (Edmond,et al., 1979). Our goal in the model is the production of an average-sized Cyprus-type ore deposit of 3 million tonnes. We can estimate the time that this should take to form from the above data. With a mass flow rate of 150 kg s-‘, an exit water temperature of 35O”C, an Fe concentration of 110 ppm and 100% efficiency, the time required will be 3000 years, or proportionately longer for lower efficiency.

309 STRUCTURAL

CONSTRAINTS

ON BLACK SMOKER FLOW

All of the active black smokers discovered

so far are located within

a few hundred

metres of the axis of fast-spreading mid-ocean ridges, where the tectonic grain is very strongly aligned parallel to the ridge crest. The vents are invariably located on ridge-crest-parallel fissures or growth faults within the axial zone (RISE, 1980; Ballard et al., 1981; Hekinian et al., 1984), which is bounded by a series of major normal edge-crest-par~lel faults (Klitgord and Mudie, 1974; Searle et al., 1981). This close association of high-temperature hydrothermal substantiated by evidence from Cyprus, where the sulphide

upflow with faults is ore bodies and their

underlying feeder pipes are associated with faults running parallel to the ancient spreading centres (Adamides, 1980; Gass and Smewing, 1981), and, since they are mostly covered with a 100-500 m thickness of lavas, they also probably o~~nated within the narrow eruptive zone (Constantinou and Govett, 1972; Cann, 1980). There is therefore strong evidence that the hydrothermal upflow is channelled in faults. There is little direct evidence of the downflow zones, but consideration of the magnitude and structure of the permeability within the vicinity of the vents strongly suggests that the downflow is also predominantly in the faults. It can be shown quantitatively that a high proportion of the flow will be channelled by the faults since they form major discontinuities in the permeability, thus short-circuiting the convective flow which will preferentially follow the low resistance path. For example, a porous medium with a cross-sectional area of several square kilometres would require a permeabi~ty greater than lo-* m2 to carry the same mass flow as a 1 km length of a narrow, rough fault (see Appendix). There is little data on the permeability of the crust in these regions. The figure quoted by Anderson et al. (1982) from the sheeted dykes at the base of D.S.D.P. hole 504B is lo-t6 m2 in crust 6 m.y. old. In crust 7.2 m.y. old in D.S.D.P. hole 39514 near the Mid-Atlantic Ridge permeabilities

were

10-‘4-10-‘5

m2 (Hickm~

et al., 1984).

Even

if the pathways

have

become clogged with time it is unlikely that these rocks ever had a permeability as high as lo-* m*. It would therefore seem highly probable that recharge flow is also channelled by the faults. The major faults bounding the extrusive zone at a distance of about 1 km from the mid-ocean ridge axis impose a limit on the extent of the model perpendicular to the ridge axis, since these faults would have channelled previous circulation systems, thus putting a limit of about 1 km on the lateral extent of hot rock which has not already been cooled. The extent of the circulation system along the strike of the ridge also appears not to exceed 1 km from the observed spacing of black smoker vents, although this is highly variable (Ballard et al., 1981; Hekinian et al., 1984). The maximum area of hot rock from which heat might be gathered is thus about 2 km’.

310

DISCUSSION OF THE TYPE OF MODEL

The limited

structural

constraints

outlined

above,

namely

to a small area in which the fault zones, forming

permeability,

channel

the flow, mean that we cannot

as a valid representation In addition

that

smoker

flow is

major discontinuities

black

in the

use porous medium

flow theory

of this system in the model, but must turn to fracture

to the structural

constraints,

our choice of a fracture

flow.

flow system has

also been governed by the type of flow to be modelled. Porous medium flow theory is based on laminar flow, which is essentially slow and requires a well-developed permeability. These two conditions are not met for black smoker flow, which arises in rocks of low permeability (see above) and which has vent velocities of several metres per second. Several lines of evidence indicate that the flow is turbulent, as one would expect in conduits as rough as faults. Converse et al. (1984) described the type of flow in the 21’N vents as “plug” flow, which is a characteristic of turbulent flow in a circular pipe. In addition the Reynolds number of lo’-10h for the upflow (Converse et al., 1984) again implies that flow will be turbulent. We consider that the most realistic and practical means of representing flow in a very rough fault is by a series of small pipes carrying turbulent flow. We assume for simplicity that downflow occurs in the same fault as the upflow, but the model could easily be modified for downflow

in another

ridge-crest-parallel

fault.

PREVIOUS MODELS

Most existing models simulate large-scale, slow, low-temperature seafloor circulation, and none has been developed so far purely with the aim of modelling rapid, very high-temperature circulation. They have usually been designed to investigate the evolution and pattern of the convection cells over a large area in order to explain the distribution of heat flow values on the ocean floor (Ribando et al., 1976; Fehn and Cathles, 1979; Fehn et al., 1983). Since they cover large areas, the scale of the fault pattern is small relative to the total area and may be regarded as constituting a porous medium, so that porous medium theory may be used as a valid representation of the flow in this case. These models do not take into consideration any pre-existing fault pattern, although Fehn and Cathles incorporate zones of high permeability in the porous medium. The fracture-loop models of Bodvarsson and Lowell (1972) and Lowell (1975) predate the discovery of black smokers and hence do not attempt to simulate very high temperature flow. The only model so far that embraces black smoker flow is that of Lister (1974, 1982), which quantifies the evolution of the circulation system from a period of rapid flow, which causes the downward advance of a cracking front, thus creating the permeable matrix in which later, slow, closed circulation may persist for millions of years. Lister’s model predicts black smoker temperatures, but only when the permeability is as high as 10-7-10-9 m2, that is many orders of magnitude

greater

than the lO-“j

m2 measured

by Anderson

et al. (1982).

311

Our structure,

model

is therefore

the flow geometry, the constraints observed

different

since we are investigating

the above

models

balance

and then only of a high-temperature,

imposed

characteristics

THE STRUCTURE

from

only the thermal

by the geological of black smoker

structure

in both

rather

aims

and

than modelling

black smoker system within at mid-ocean

ridges

and

the

flow.

OF THE MODEL

The model examines the heat exchange between running parallel and close to the mid-ocean-ridge axis, The heat supply is maintained by horizontal conductive The model consists of an open thermosyphon within

convective flow in a fault, and the adjacent hot rocks. heat flow through the rock. a single constant-width fault,

which is treated as a series of pipes. Seawater at 0°C enters the fault and flows downward in a long zone extending along the fault, and eventually rises in a narrow upflow zone. The flow is driven by the buoyancy difference between the recharge and discharge zones, and the resistance to flow arises from the hydraulic properties of the system. Downflow and upflow velocities vary in proportion to the respective areas of recharge and discharge for continuity of mass. This model has been developed from

Fig. 1. The geometry of the model. The half-width of the fault only is shown.

312

a simpler model with the same geometry flow velocities were constant. Heat is exchanged conductive transient

at the water-rock

(Strens and Cant-r, 1982) interface

heat flow in the rock is calculated heat conduction

since it will be negligible

equation.

of the fault wall. The horizontal

analytically

from the one-dimensional

Vertical heat conduction

in the vicinity

but in which the

of a vent compared

in the rock is ignored, with the convective

heat

transfer. The convective water flow is treated digitally so that heat exchange is calculated at each level with the pencil of rock (referred to as a “block”) adjacent to each limb (Fig. 1). The two limbs of the system are assumed to be thermally independent.

Each block of rock has a different

but uniform

initial

temperature,

so

that any temperature profile may be set up. The model is run on a cyclical basis so that the contact temperature, which constrains the heat conduction, may be changed according to the new rock and water temperatures calculated at the end of each time increment. The fault is assumed to open suddenly to a given depth, and is then propagated downwards at a given rate unless the water temperature rises above a given temperature at which cracking is unlikely to continue. The fact that little is known about typical cracking temperatures, or the process of cracking at sub-seafloor pressures, necessitates this somewhat crude method. The advance of the fault is thus controlled by the fault propagation rate at lower water temperatures, but is arrested while the water temperature is high enough to inhibit cracking. Values given to these parameters

are discussed

below.

THE MODEL EQUATIONS

The solution to the one-dimensional transient heat conduction equation, constrained by a contact temperature, for a semi-infinite block of rock is given by Carslaw and Jaeger (1959): r,=T,+(T,-T,)erf

x i 2(at)“2

T,= T,

(Z=O,O
T, = T.

(x=0)

1

(1)

co)

T, is the rock temperature at distance x from the contact, Tc is the contact temperature between rock and water, T, is the initial rock temperature, x is the horizontal distance from the fault, (Y is the thermal diffusivity of the rock and t is the elapsed time. The contact temperature, q, in eqn. (1) is steady. To allow the contact temperature to be changed at the end of each cycle, while remaining steady during each cycle, eqn. (1) is modified in a way analogous to that given by Carslaw and Jaeger (1959, eqn. 3, p. 63) so that a factor is added for each change in contact

313

temperature: For n cycles: T,=

7;.,,,,+(T,-

X

7;.,,,) erf

erf i

If the heat transfer

with the water at x = 0 were perfect

the contact

temperature,

TC, would be the same as the water temperature T,. We have given some impedance to the heat flow by taking a weighted mean of the rock temperature (at x = 0.1 m) and the water temperature in inverse proportion to the thermal diffusivities of rock and water so that: TC= 0.2T,=, , + OJT,.. Other proportions were tested and the model was found to be relatively insensitive to this formulation, because the water and rock temperature are close except during a short initial period. The heat flow from the rock at x = 0 is derived as follows: Partial differentiation of eqn. (1) gives the temperature gradient at x = 0:

ar (T,-T,) Bx = ( Tat)‘/2 this means heat flow from rock at x = 0: -KA(q-c) q, =;” =

(2) ( nat)“2

where K is the thermal conductivity of the rock. The value of A, the heat exchange contact area, varies according to whether the equation is applied to the upflow zone or the downflow zone. The total heat flow, Q, up to time t is obtained by integrating eqn. (2): Q=2KA(7;-T,)

i

&

1

“I

(3)

Equation (3) is modified in the same way as eqn. (1). so that the contact temperature can be changed after each cycle. For n cycles, the conductive heat flux at the water-rock interface is: Q,,=2KA(q,,,-T,)

The heat flow from each block of rock during the nth cycle, Qcvc = Q,, - Q,,- ,, is equated to the heat removed by the water within and flowing through the fault adjacent to it, assuming the water to be a well-stirred fluid. Heat exchange equation:

Qcyc= c,,m4Ah

+ u,,Ar)(Ti

- T,.)

314

where cm. P,,, and u,,, are the specific heat, density and linear flow velocity of the water in the mth block. Temperatureand pressure-dependent values of density and specific heat are used. w is the half-width fault

and

increment,

takes

a different

i.e. thickness

temperature

value

of block,

of the fault, u is the dimension

for upflow

and

downflow,

At is the cycle length

and

along the

Ah is the vertical T: is the new water

after time At.

The mass flow through Continuit_); equation:

the system is defined

by the exit mass flow rate.

= awp,, u,, awp,,, u 111 where p_ and u,, are the exit density and flow velocity. This equation applies to both upflow and downflow provided the appropriate value for u is taken. The water temperature thus obtained at each position in the loop at the end of each time increment is then used to calculate the buoyancy force and hence the flow velocities for the next cycle. Flow equution:

where

u,. 5,. d, and f,

are the velocity,

mean

density,

friction factor for the upflow, p is the mean density the mean density of the downflow.

equivalent

diameter

and

for the whole system, and p1 is

Equation (4) is derived, as follows, by equating the total pressure drop in the system, assuming turbulent pipe flow, to the buoyancy force. The total pressure drop consists of the drops due to inertial forces at E (Fig. 2) the acceleration at A and friction in both limbs (Massey, 1979). Pressure drop = apuf + iP( U, - Up)’ + 4ph

$

? ?

Fig. 2. Diagrammatic limbs is indicative

representation of their thermal

E, and at the constriction,

:

fu’ + 2 y ’

i

of the flow system in the model. The space shown between independence.

Resistance

A, at the base of the upflow

the two

to flow arises at the entrance

to the system,

limb, as well as from the friction

in both limbs.

315

where uZ. d, and fi are the velocity,

equivalent

diameter

and friction

factor in the

downflow. Since

U, >> u2 and 2hf,/d,

B 4 for typical

values

of h - 1 km, f, - 0.01 and

d, - 0.03 m (see below) 2

Total pressure

drop

= 4$h 5

? 1

i.e. the pressure drop due to friction in the upflow pipe only. This pressure drop is equated to the buoyancy force kg(& - &) to give eqn. (4). The term 5d,/f, will be referred to as the friction constant (F). Its value is difficult to determine since no data are available on the roughness of the conduits, and reports of vent diameters range from 30 cm at the 21*N vents (Macdonald et al., 1980) to 3 cm at 13”N (Hekinian et al., 1984). Since the feeder zones seen beneath Cyprus ore deposits are composed of a network of narrow, anastomosing fissures, we have chosen to model the discharge zone as a bundle of narrow, rough pipes. If each pipe is taken to have a diameter of 3 cm and a friction factor of 0.01, corresponding to a roughness on the walls of the pipe of about a millimetre, which is consistent with geological observations, then, at a discharge water temperature of 350°C, and at a load pressure of 350 bar, it will pass fluid at a linear velocity of 2 m s-‘, which is the mean velocity observed for black smokers. A pipe of this diameter and roughness thus possesses the hydraulic resistance appropriate to black smoker conditions, and a bundle of 150 such pipes would pass 150 kg s-‘, the mass flow rate estimated for the 21°N black smoker field. The diameter, d,, and number, n, of pipes are related to the fault half-width, W, so that nndff4 = 4w2, assuming that the discharge zone is equidimensional. These values for d, and f, yield a value for F, the friction constant, of 15. CONSTRAfNTS

ON THE BOUNDARY

AND

INITIAL

CONDITIONS

The dimensions of the model are determined which the circulation occurs, which, for black mid-ocean

ridge axis. There is strong

evidence

by the structural smoker systems,

environment in is close to the

that very often, especially

on faster

spreading ridges, the eruptive zone is underlain by a magma-chamber at a depth of 1-2 km. Such evidence not only comes from seismic studies near spreading centres (Orcutt et al., 1976), but also from petrological studies on ocean-floor lavas (Bryan and Moore, 1977; Flower et al., 1977) and from geological mapping of ophiolite complexes, where thick gabbro units, clearly formed from crustal magma chambers, underlie about 1 km of lavas and 1 km of dykes (Moores and Vine, 1971; Gass and Smewing, 1981). Thus the hydrothermal circulation would be expected either to be confined to the l-2 km of hot rock overlying a magma chamber, or to extract heat from the magma itself, or to penetrate into the hot rock of a recently solidified magma chamber as far as the base of the oceanic crust. This provides clear

316

constraints circulation

on the depth of circulation in the model. We have therefore depths of 1 km and 6 km. For 1 km circulation the geothermal

taken as 1 K m-’ magma depth

chamber.

from 0°C at the seafloor

tested two gradient is

to 1000°C at 1 km depth, just above a

For 6 km deep circulation

of 1 km with the 5 km of rock below

the gradient at 1000°C

is again

1 K m‘-’

representing

to a

a recently

solidified magma chamber. Perpendicular to the ridge axis the limit of the extent of hot rock is about 1 km (see above). The extent of the mode1 in this direction is infinite, but the contribution to the heat flux from beyond a distance of 1 km is found to be negligible within the time span of several thousand

years.

ORE MASS CALCULATION

The model calculates the exit water temperature and mass flow rate for each time cycle during the life span of a hot spring. The resultant ore mass can then be calculated if one assumes that all of the dissolved iron is precipitated as Fe!&. The accuracy of this calculation depends on the accuracy of the iron solubility-temperature data. We have used line B in Fig. 3, which is based on experimental data of Mottl et al. (1979) with two points added: the measured concentrations in black smokers at 21“N and 13”N (Edmond, 1981; Michard et al., 1984), and Seyfried and Bischoff’s (1977) experimental point. Mottl et al.‘s points indicate an exponential increase in iron solubility with increasing temperature (line A). Since the system being modelled is an open system with a much higher overall water/rock ratio than Mottl et al.‘s experimental systems, we have constructed line B, which is heavily weighted towards the 21’N observational data and also Seyfried and Bischoff’s point which has a water/rock ratio of 50 : 1.

Fig. 3. Fe-solubility (Edmond,

1981),

data. Data are: e-from

q-Seyfried

Moitl et at. (1979),

@-21”N

black smoker

analysis

and Bischoff (1977). Line A is drawn through the points of Mottl et al. Line

B is the weighted line (see text) used for ore mass calculations.

317

The exponential relationship between Fe solubility and temperature shown in Fig. 3 means that the water temperature is the do~nating

factor in determining

the

resultant ore mass, and only when the water temperature is greater than 300°C is the iron solubility sufficiently high for it to contribute a significant amount of ore. The following results are based on an efficiency of deposition of 100%. RESULTS FROM THE M0DEL

Our aim, as stated above, is to test whether sufficient heat is available from the rocks to sustain the circulation system in the model, so that the vented water temperature is at least 35PC with a mass flow rate of about 150 kg s-l for several thousand years (or ~roportion~ly longer if the flow rate is less) in order to produce an iron sulphide deposit of 3 million tonnes. The values taken for the parameters required by the model are given in Table 1. In general these parameters are not well constrained, so the values taken represent the best estimate available to us for the ridge axis environment. The recharge zone length is taken as 1 km. This is based on observed spacings of black smoker fields, and on the fact that convection cells tend to be ~~dimension~. The fault propagation rate of IO m yr- 1 is based on Lister’s (1982) estimate. Other lower values were tested, including zero, but these gave lower water temperatures, thus reducing the potential for ore deposition and therefore of little interest in this analysis. The value taken for the crack propagation rate has little significance at high water temperatures which will inhibit cracking. Since we are unable to define the relative water and rock temperatures at which cracking will occur, we have tried to

TABLE 1 Values for the variable parameters in the mode8 Faranleter

Standard wfue

Recharge zone length

I km

Fault propagation

IOmyr-’

rate

Range of values

laitiai fault depth

5OQm

-

Width of fault

a.4 m

0.2-0.8

Friction constant

1.5

l-la0

m

Initial temp. field in rock circ. to 1 km

1

Km-’

10 K’m-’

to 100 m depth

10oO°C beIow zoo m ck

to 6 km

lKrnwl

tolkmdepth

and 1000°C below 1 km

-

318

make the model as realistic as possible by only allowing the crack to advance while the water temperature at the base of the circulation is under 370°C. By doing this the maximum observed

exit water

black smoker

temperatures

temperatures.

in the model

was chosen as the level at which the rock temperature estimate ashighasl

of the cracking

temperature

are similar

to the maximum

The value of 500 m for the initial

fault depth

will be below Lister’s (1982)

of 527’C, even when the geothermal

gradient

is

Km-‘.

The parameters that have highly variable, and therefore unknown values, namely the fault width and the friction constant, were tested through a range of realistic values. The fault width determines the water/rock ratio of the volumes exchanging heat, and the friction constant determines the linear flow velocity for a given water temperature. These parameters determine the mass flow rate and therefore the rate of heat removal. The standard values were chosen so that together they give a mass flow of around 150-200 kg s-’ at black smoker temperatures, which is consistent with estimates for observed black smoker fields (Converse et al., 1984). In addition to the standard initial temperature profile in the rock a “high” rock temperature profile was also tested for the 1 km deep circulation, since it is possible that black smoker circulation may arise shortly after volcanic activity. In this case the rock temperature was taken as 1000°C to within 100 m of the seafloor with a linear gradient above.

Circulation to I km depth The model was first run with the standard parameters (Fig. 4, a). After an initial burst of heat during the first year the exit water temperature never reached 3OO”C, so that the resultant ore mass would be negligible.

I

OO

500.

i 000

1500

1

Years

Fig. 4. Exit water temperatures for 1 km deep circulation. Geothermal gradient in rock = 1 K m- ‘. All parameters take standard values (Table 1) unless specified otherwise. a -standard friction (F = 1). c-narrow

fault width (0.2 m).

parameters, b-high

319

I

OO

500

Fig. 5. Exit water temperatures

1000

I

1500

for 1 km deep circulation.

Years Geothermal

100 m depth with rock at 1000°C below 100 m, to represent conditions Standard parameter values unless specified otherwise. c-narrow

fault width (0.2 m), d-high

a -standard

friction (F=l)

gradient in rock = 10 K m-’

to

shortly after volcanic activity.

parameters,

b-high

friction (F = l),

and narrow fault width (0.2 m). Potential ore

mass given in million tonnes.

The variable parameters were then tested systematically. The friction in the upflow pipe was increased to its maximum feasible value (Fig. 4, b) and the fault width was halved to 0.2 m (Fig. 4, c), but temperatures above 300°C were still obtainable only for about 30 years. These runs were repeated with the “high” initial rock temperature profile (Fig. 5, a, b, c) but temperatures above 3OO’C were still achieved for a maximum of only 130 years (Fig. 5, c). The extreme values for friction and fault width were then combined (Fig. 5, d) yielding an exit water temperature in excess of 300°C for 1000 years, but, because the prolonged high temperature was obtained by reducing the mass flow rate to about 20 kg s-l, the resulting ore mass was only 0.2 million tonnes, and very far short of our goal of 3 million tonnes.

Circulation to 6 km depth

Water temperatures of 300°C were again achieved only during the first year when the model was run with the standard set of parameters (Fig. 6, a). Once again it was necessary to reduce the mass flow rate to prolong high-temperature flow (Fig. 6, b, c) so that, with a fault width of 0.2 m for example, exit temperatures of over 300°C could be maintained for 3500 years with a mass flow rate of 50 kg s-l, potentially yielding nearly 2$ million tonnes of ore. Our goal of 3 million tonnes could clearly be reached by changing the parameters to reduce the mass flow rate further, but the system would then be rather different from the natural systems we are trying to model, since at low mass flow rates either the water temperature must be higher than 350°C or the flow must continue for many thousands of years.

320

0

0

2000

1000

Fig. 6. Exit water temperatures km depth with rock at 1000°C parameter

3000

for 6 km deep circulation. below 1 km, to represent

values unless specified

row fault width (0.2 m). Potential

otherwise.

a -standard

Years

Geothermal a recently

gradient

solidified

parameters,

b -high

in rock = 1 K m

magma

chamber.

friction

’ to 1

Standard

(F = l), c -nar-

ore masses given in million tonnes.

CONCLUSIONS

The foregoing results show that the system modelled can produce sizeable ore deposits (> 1 million tonnes) if the circulation penetrates to the bottom of the crust, but that no significant ore deposits will result from 1 km deep circulation. These conclusions are based on the results obtained using those values for the parameters most favourable to ore deposition. Whatever reasonable parameters are taken the 1 km deep circulation will not produce large ore deposits, and the ore masses resulting from the 6 km deep circulation will be maxima unless local conditions permit a longer recharge zone. The ore masses have been calculated on the assumption that precipitation of dissolved iron as iron sulphide is 100% efficient. If the efficiency were low, as it clearly

is at the East Pacific

Rise vents,

then the total

mass flow, and hence

the

amount of heat required, would need to be several times greater to produce the same ore deposit. In this case the flow would need to continue for much longer, since the model has shown that a higher mass flow rate would result in lower exit temperatures. Alternatively large ore deposits may only be possible where local conditions cause highly efficient deposition, such as in the Galapagos warm springs field. It is clearly possible then that ore deposits of a few million tonnes could result from convective flow in a fault zone which penetrates through a newly-solidified magma chamber to the base of the crust, provided that conditions favour highly efficient deposition, but such circulation is unlikely to account for the largest ore masses, which can be lo-15 million tonnes in Cyprus. If the circulation is shallow (penetrating to only 1 km depth), the model shows that black smoker flow is very short-lived, implying that longer-lived ore-forming black smokers require a greater heat flux than can be obtained by the system in the

321

model from a 1 km thick layer of hot rock. There are three possible the system may acquire that

the flow penetrates

shallow circulation capable

through

may be sustained

to be an underlying systems

a greater heat flux. One we have already a greater

crustal magma chamber. of depositing

thickness

by an additional

of hot rock.

namely

Secondly

the

heat source which would have

The third alternative

very large ore masses

ways in which

considered,

may only

is that circulation arise in locations

where there are no major faults to channel the flow, so that heat can be extracted from an extensive area of rock by porous medium circulation, either with flow entirely in a porous medium as in Lister’s (1982) model, or as in our third model with porous recharge and fault discharge. This form of circulation would be more efficient at heat extraction than the system modelled (which relies on conduction, a slow method of heat transfer), but only if the permeability is high enough. This alternative is unlikely to be a valid representation of the system generating the Cyprus ore deposits, which are clearly associated with major faults (Adamides, 1980) close to a spreading centre. Of these three alternatives we favour the possibility that a magma chamber is the principal source of heat for the formation of large volcanogenic sulphide bodies, since the heat supply from rock alone is always limited by the volume of rock that has not been cooled by previous circulation (Cann and Strens, 1982). The heat supply from the magma chamber will mainly be in the form of latent heat, produced at a rate of lo9 J m-3 of magma crystallised. Our model of magma-driven hydrothermal circulation (Cann et al., in press) shows that a 3 million tonne ore deposit could be formed, for example, in 4000 years by a black smoker system with 70% efficiency with the heat from 30 km3 of magma, the heat being conducted through a solid boundary layer about 10 m thick. So much heat is needed to form the very largest ore deposits that they may require the total freezing of a magma chamber and its subsequent penetration by the hydrothermal circulation system. We have shown, however, with this model that it is possible for fault-dominated circulation, located close to the mid-ocean-ridge axis, to be maintained by the heat supply from the rock for long enough to produce ore deposits of several million tonnes but only when conditions are particularly favourable. Although we have used this model to examine black smoker circulation at spreading centres, it could also be applied to other environments, and appears particularly suitable for modelling flow in fault zones in an expanding sedimentary basin, and thereby providing the means of predicting the size of the resultant lead-zinc sulphide deposits.

ACKNOWLEDGEMENTS

This work was supported typed by Elizabeth Walton.

by N.E.R.C.

grant

GR3/5073.

The manuscript

was

322 Appendix: COMPARISON OF FLOW IN A FAULT AND A POROUS MEDIUM

Compare medium

the downflow

in a 1 km length

of fault

40 cm wide and a porous

of 2 km’ extent.

(1) Fracture flow Flow will be turbulent since the extreme roughness of the fault would prevent laminar flow. In addition the Reynolds numbers calculated for the 21 “N system with a mass flow of 150 kg s-’ are greater than the critical Reynolds number in all but the coolest part. The fault is considered as many small pipes. Darcy’s relation for steady, turbulent pipe flow (Massey, 1979) defines the pressure gradient due to friction in a pipe. AP

2f

2-

P

z=--p

(1)

where f is the friction jj the mean density.

factor, d is the pipe diameter,

u is the linear flow velocity and

u = J/A&i where J is the mass flow rate and A, is the total cross-sectional area of the fault. Taking d =0.03 m, f = 0.05, p = 923 kg m-’ (at 150°C, the mean downflow temperature), flow:

A, = 0.4 x lo3 m and

J = 150 kg SK’ (cf. 21”N

system).

For fault

-AP = 5 x 10m4 kg m-2 se2 AZ (2) Porous medium frow The porous medium flow equation, gradient due to viscous laminar flow. AP _=_ AZ where

also named

after Darcy, defines

the pressure

Jv

(3)

A,,pk n is the dynamic

viscosity,

A,,

is the cross-sectional

area

and

k

the

permeability of the porous medium. Taking n = 1.85 X 10m4 kg m-’ ss’ (at 15O”Q A,, = 2 X lo6 m2, ,E= 923 kg mm3 (at 150°C) and substituting from (2) the permeability necessary to carry the same mass flow as the fault can be calculated, giving k = lo-’ m2. REFERENCES Adamides, N.G., 1980. The form and environment of formation of the Kalavasos ore deposits, Cyprus. In: A. Panayiotou (Editor), Ophiolites. Proc. Int. Ophiolite Symp., Cyprus, 1979, pp. 117-127.

323

Anderson.

R.N., Honnorez,

Kinoshita. reference Ballard,

J., Becker, K., Adamson,

H., Laverne.

C.. Mottl.

M.J. and

section over I km through

R.D.. Francheteau,

tectonic Bodvarsson,

1982.

R., Kempton,

DSDP

crust. Nature,

Hole

504B.

W., 1981. East Pacific Rise at 21”N:

at the central

axis. Earth Planet.

processes

P.D..

the first

300: 589-594.

C., Normark,

G. and Lowell, R.P.. 1972. Ocean-floor

Geophys.

R.L.,

Layer 2 of the oceanic

J., Rangan,

and hydrothermal

A.C.. Ah. J.C., Emmermann,

Newmark,

the volcanic,

Sci. Lett., 55: l-10.

heat flow and the circulation

of interstitial

waters.

J.

Res., 77: 4472-4475.

Bryan, W.B. and Moore, J.G., 1977. Compositional

variations

of young basalts

in the Mid-Atlantic

Ridge

rift valley near lat. 36O49’N. Geol. Sot. Am. Bull., 88: 556-570. Cann,

J.R., 1980. Availability

Cann,

J.R. and Strens,

Cann,

J.R., Strens,

formation Carslaw,

of sulphide

ores in the ocean crust. J. Geol. Sot. London,

M.R., 1982. Black smokers

fuelled by freezing

M.R. and Rice, A., 1986. A simple,

of volcanogenic

massive sulphides.

magma.

magma-driven

Nature,

thermal

137: 381-384.

298: 147-149.

balance

model

for the

Earth Planet. Sci. Lett.. in press.

H.S. and Jaeger, J.C.. 1959. Conduction

of Heat in Solids. Oxford

University

Press, London,

2nd

ed. Constantinou,

G. and Govett.

G.J.S.,

Inst. Min. Metall. Trans.. Converse,

D.R., Holland,

Pacific Rise (21”N):

1972. Genesis

of sulphide

deposits,

ochre and umber

of Cyprus,

81: B34-46.

H.D. and Edmond, implications

J.M., 1984. Flow rates in the axial hot springs

for the heat budget

and the formation

of the East

of massive sulphide

deposits.

Earth Planet. Sci. Lett., 69: 159-175. Edmond,

J.M.. 1981. Hydrothermal

Edmond.

J.M., Measures,

activity

C.. McDuff,

J.B., 1979. Ridge crest hydrothermal ocean: Fehn.

the Galapagos

U. and

Cathles,

Tectonophysics, Fehn, U., Green, M.F.J.,

systems Gass,

1979.

Hydrothermal

Robinson,

Center.

P.T.. Schmincke,

the Mid-Atlantic

H.-U.

J.D.. 1981. Ophiolites:

Lithosphere.

in the

mid-ocean

ridges,

and

models for the hydrother-

Res., 58: 1033-1048. Ohnmacht,

Contrib.

Obducted

W.. 1977.

Mineral

oceanic

Magma

fractionation

Petrol.. 64: 167-195.

lithosphere.

In: C. Emiliani

(Editor).

Vol. 7. The Sea. Wiley, New York. pp. 339-363. J., Renard.

J.C., Boulegue,

V., Ballard.

R.D., Choukroune.

J. and Charlou. submersible

P., Cheminee,

J.L., 1983. Intense witnesses

J.L.. Albarede.

hydrothermal

the growth

of sulphide

F.,

activity

at the

chimney.

Mar.

Res., 6: 1-14.

R., Renard,

13”N: geological Laubier

at slow-spreading

L.M., 1983. Numerical

J. Geophys.

Ridge at 36-37ON.

axis of the East Pacific Rise near 13’N: Hekinian,

convection

R.P. and Cathles,

Spreading

J.F., Marty,

Geophys.

L.I. and Corliss,

of the major and minor elements

L.M.,

R., Francheteau.

Minster,

and the balances

Earth Planet. Sci. Lett., 46: 1-18.

K.E., Von Herzen,

I.G. and Smewing,

Hekinian,

290: 87-88. B., Gordon,

55: 239-260.

beneath

The Oceanic

activity

ridge axes. Nature,

L.H., Collier, R., Grant.

data.

mal field at the Galapagos Flower,

at mid-ocean

R.E.. Chan.

V. and Cheminee, setting

J.L., 1984. Hydrothermal

and distribution

and K.L. Smith (Editors),

deposits

of active sulfide chimneys.

Hydrothermal

Processes

on the East Pacific Rise near In: P.A. Rona.

at Seafloor

Spreading

K. Bostrom. Centers.

L.

Plenum,

New York. pp. 571-594. Hickman.

S.H., Langseth.

M.G. and Svitek, J.F., 1984. In situ permeability

ments near the mid-Atlantic U.S. Government Klitgord.

K.D.

and

survey. Geophys. Lister, C.R.B., Lister,

C.R.B.,

physical Ocean

ridge, DSDP Hole 395A. Initial Reports

Printing Mudie,

Office, Washington, J.D.,

J.R. Astron.

conditions.

Project.

D.C., Vol. 78B, pp. 699-708.

1974. The Galapagos

and

Spreading

Centre:

of water into hot rock. Geophys. “passive”

In: F.T. Manheim

Floor. D.C. Heath,

measure-

a near-bottom

geophysical

Sot., 38: 563-586.

1974. On the penetration 1982. “Active”

and pore pressure

of the Deep Sea Drilling

Lexington,

hydrothermal

and K. Fanning Mass., pp. 44-470.

systems (Editors),

J.R. Astron.

Sot.. 39: 465-509.

in the oceanic The Dynamic

crust:

predicted

Environment

of the

324

Lowell. R.P.. 1975. Circulation crests. Geophys. Macdonald.

G., Albarede.

solutions

heat transport

R.D., 1980. Hydrothermal

of Fluids. Van Nostrand-Reinhold,

F.. Michard.

A., Minster.

J.-F.,

on mid-ocean

ridge

heat flux of the “black

E.M. and Vine, J.F.,

1971. The Troodos

and implications.

M.J.. Holland.

Philos. Trans.

Massif.

II. Experimental

J.-L..

Tan,

N.. 1984. Chemistry

of

site. Earth Planet. Sci. Lett.. 67: 297-307. Cyprus

R. Sot. London.

H.D. and Corr. R.F.. 1979. Chemical

basalt by seawater.

New York, 4th ed.

Charlou,

from the 13”N East Pacific Rise hydrothermal

evaluation Mottl,

and convective

vents on the East Pacific Rise. Earth Planet. Sci. Lett.. 48: 1-7.

B.S., 1979. Mechanics

Michard. Moores.

hot springs

Sot., 40: 351-365.

K.C.. Becker, K., Spiess. F.N. and Ballard,

smoker” Massey,

in fractures.

J.R. Astron.

and other

ophiolites

as ocean

crust:

Ser. A, 268: 433-466.

exchange

during

hydrothermal

results for Fe. Mn and S species. Geochim.

alteration

Cosmochim.

of

Acta. 43:

8699884. Orcutt.

J.A.. Kennett.

bottom

B.L.N. and Dorman.

seismometer

Ribando,

R.J.. Torrance,

in the oceanic R.C..

Barrow, Geophys. Seyfried.

Francis,

L.M., 1976. Structure J.R. Astron.

K.E. and Turcotte.

crust. J. Geophys.

RISE. 1980. Hot springs Searle,

survey. Geophys.

B.J. and Bicknell.

Hilde.

D.L.. 1976. Numerical

models for hydrothermal

circulation

Res., 81: 3007-3012.

and geophysical T.J.G.,

of the East Pacific Rise from an ocean

Sot.. 45: 305-320.

experiments

T.W.C..

on the East Pacific Rise. Science. 207: 1421-1433.

Somers,

S.V.. 1981. “Gloria”

M.C..

side-scar

Revie. J.. Jacobs. sonar

C.L.,

in the East Pacific,

Saunders. Eos. Trans.

M.R., Am.

Union. 62: 121-122.

W.E. and Bischoff,

seawater/basalt

J.L.. 1977. Hydrothermal

transport

Strens, M.R. and Cann. J.R.. 1982. A model of hydrothermal crests. Geophys.

of heavy metals by seawater:

the role of

ratio. Earth Planet. Sci. Lett.. 34: 71-77. J.R. Astron.

Sot.. 71: 225-240.

circulation

in fault zones at mid-ocean

ridge