You are on page 1of 7

Global heat flow maps: Their use for environmental processes a literature review

CHARLIE KENZIE
Department of Earth Sciences, University of Durham 2014

1. Introduction
The understanding of the composition and
environmental processes of the solid earth are
a major task for geoscientists. Earths surface
heat flux is a fundamental output of the
dynamic solid Earths heat engine (Davies
2013). The magnitude of heat loss measured
from the Earths surface is orders of
magnitude greater than the resultant energy
released by other energy fluxes such as
seismic strain release, or geomagnetic field
generation (Pollack et al 1993).

Thus, the determination and description of
regional variations in heat flow is an
important endeavor in global geophysics, and
provides a constraint on the internal state of
the mantle, and allows us to investigate the
tectonics of continents and ocean basins,
seismic wave velocities in the Earths crust
and upper mantle, seismicity, rheology,
crustal magnetism, hydrothermal circulation,
and the maturation of hydrocarbons, which all
depend strongly on the temperature
distribution in upper most few hundred
kilometers of the earth (Pollack et al 1993).

The following literature review aims to
highlight the importance of heat flux as an
insight into the environmental processes
described above. This short review will
follow the history of heat flow data
acquisition, common methodologies and
analysis for mapping, interpretation of heat
flow with reference to environmental
processes, and finally leading to a discussion
of the benefit and drawbacks of heat flow data
and the potential of heat flow mapping for
future studies.

2. Heat flow Data acquisition
Some of the first heat flow data was collected
by Benfield (1939) and Bullard (1939), who
assumed that the amount of heat flowing from
the earth in unit time per unit area (H), was
given by
!
H = kp where k and p are thermal
conductivity and temperature gradient in the
earths crust respectively. Rock conductivities
and temperature gradients were calculated by
means of placing individual rock specimens
under a number of thermojunctions inside a
glass thermometer. Benfield (1939) and, to a
further extent Bullard (1939), both accept that
errors, associated with the apparatus and
compounded by the removal of the rocks from
situ, can not be ignored when using this
technique. Although argued to be negligible
in their measurements, Stein (1995) highlights
the benefits of digital instruments, which have
superseded Benfields technique since the
1970s.

Firstly, digital instruments can determine
temperature much more accurately than the
Benfield apparatus. Furthermore, many
measurements can be made in a relatively
small area, increasing the resolution and
allowing local variations in heat flux to be
better identified (Stein 1995). Chapman and
Pollack (1975) underline the problems of too
few data in earlier heat flow maps, and
suggest that previous analyses of heat flow
show unreal distortions in the harmonic
representation of the heat flow field, caused
by lack of observations in several critical
areas. Although more advanced tools for
investigating heat flow in challenging settings
have allowed the global data set to increase,
the problem of under-representation is still a
compounding issue today (Fig.1), but we will
revisit this notion later on.

Stein (1995) describes the different
techniques used to acquire heat flow data.
Marine measurements are made by thrusting a
probe into ocean sediments to depths of about
5m. Since the sinking probe causes frictional
heating, it is allowed to rest for up to 30
minutes to allow this energy to dissipate.
Terrestrial measurements are made by drilling
holes and by descending instruments into the
hole by a cable. Similar problems arise from
the frictional energy caused by drilling. Stein
(1995) continues to discuss the associated
errors of such techniques, but again these
uncertainties will be revisited in detail later.

3. Data analysis
Almost all analysis of heat flux data, for the
purpose of creating heat flow maps, is
undertaken by spherical harmonic analysis
(Chapman & Pollack 1975; Davies 2013;
CHARLIE KENZIE
Hamza et al 2007; Pollack et al 1993; Stein
1995). Before this analysis takes place, a
number of variables need to be considered
before an analytical approach can proceed.
The reviewed literature argues over the
specific treatment of the data before and
during analysis in such length that it is
explored in its own right, more general
uncertainties are discussed later.
The first consideration is the effect of
hydrothermal activity on oceanic regions.
(Chapman & Pollack 1975; Pollack et al
1992; Stein 1993) suggest that heat flow data
from the oceanic crust shows an
underestimation of conductive heat flow.
Stein (1993) discusses how the discrepancy
between heat flow measured at the sea floor
and higher values predicted for the ages of the
oceanic crust, is attributed to hydrothermal
circulation. Several other studies (Anderson et
al 1977; Sclater & Francheteau 1970; Sclater
et al 1976; Stein 2007) reveal that
hydrothermal circulation of seawater through
the oceanic crust is a significant mode of heat
transfer in young oceanic regions. Therefore
areas affected hydrothermally will show
lower conductive heat flow values than would
be observed without such effects (Pollack et
al 1993).

A further consideration arises from the
problem of incompleteness of the heat flow
data catalogue. Some of the first
representations of global heat flow (Lee and
Macdonald 1963; Lee and Uyeda 1965; Horai
and Simmons 1969) make use of analytical
methods in which the observed heat flow at
each geographic site is represented as a finite
series of spherical harmonics with unknown
coefficients. A system of over-determined
equations can then be solved for the
coefficients (Fig.2) (Lee & Uyeda 1965).
However, (Chapman & Pollack 1975; Hamza
et al 2007) highlight that this approach causes
areas where there is little data to be sensitive
to numerical instabilities (Hamza et al 2007).
Although this analysis operates only on real
point observations it yields coefficients that
are biased by the heterogeneity of the data
causing unrealistic distortions in the heat flow
map (Pollack et al 1993).

(Chapman & Pollack; Pollack et al 1993)
address this problem by approaching the
spherical harmonic analysis using a full set of
global heat flow data, equally spread over
5 ! 5 elements, thus determining a finite
number of coefficients by direct integration
over a sphere. However, this requires
estimates of heat flow in areas with no data.
The first studies to analyse heat flow data
using this technique (Chapman & Pollack
1975), and subsequently (Chapman et al
1993), argue that empirical predictors for
unsurveyed areas are reasonable estimators
and can be applied to supplement the
observed data set. The empirical predictors
are based on correlations between heat flow
and tectonic setting recognized by Polyak and
Smirnov (1968). Several other studies
(Chapman & Furlong 1977; Sclater &
Francheteau 1970; Vitorello & Pollack 1980)
also recognize the relationship between
continental heat flow and tectonic age (Fig.3).

Similar predictors were calculated for oceanic
regions, based on the decrease of heat flow
with age of the ocean floor (Anderson &
Hobart 1976; Polyak & Smirnov 1968; Stein
& Von Herzen 2007). For the Chapman &
Figure 1 (a) Geographic distribution of heat flow
measurement sites from Pollack et al (1993), which
includes 24,774 observations. (b) Geographic
distribution of heat flow measurements from Davies
(2013), which includes 38,374 observations. Notice
the very inhomogeneous distribution in both figures,
particularly in Fig.1a

a)
b)
CHARLIE KENZIE
Pollack (1975) and Pollack et al (1993)
analysis, this relationship provides an
empirical predictor for large areas of oceanic
crust where there are no data. The coupling of
the empirical heat flow predictors, for oceanic
and continental crust, and their combination
with the observed data produces a global heat
flow map shown above in Fig.2(c).

Despite the vast progress in data acquisition
since the 1970s (Stein 1995; Hamza et al
2007) the heat flow data analysis by Chapman
& Pollack (1975) (Fig.2c) has been largely
unchallenged until recently. A relatively
significant updated analysis by Pollack et al
(1993), using the same empirical predictor
technique, produced a similar global heat flow
distribution but with a 30% increase in mean
heat flow of the earth (Pollack et al 1993).
More recently, their treatment of the data and
analytical approach has been challenged by
Hamza et al, who argue that any analysis
using theoretical heat flow values as a
substitute for experimental data causes an
inherent ambiguity in the results.

Hamza et al (2007) start their criticism of the
Pollack analysis by re-appraising the
assumption that heat flow is underestimated
in oceanic regions due to hydrothermal
activity. Several studies (Garg & Kassoy
1981; Holst & Aziz 1972; Ribando &
Torrance 1976) have suggested that an
underestimation of heat loss only occurs in
areas where hydrothermal activity is
unconfined i.e. heat energy can escape.
Several other studies (Becker & Fisher 2000;
Bryant et al 1981; Embley et al 1991;
German et al 1994; Lowell et al 1993; Meyer
& Hemley 1967) suggest that fluid discharges
are rare in areas of ocean crust away from
spreading centers, that thick sedimentation
rates associated with young oceanic crust act
as an impermeable barrier against widespread
hydrothermal circulation and that, due to the
fall in the permeability of fracture systems
Figure 2 (a) Orthogonal function representation (to the third-order spherical harmonics) of 987 heat flow values.
Contour lines are in cal/cm
2
and are dashed over where no data exist from Lee & Uyeda (1965). (b) Global map of
Earth surface heat flow in mW m
-2
. It uses the ocean heat flux estimates from geological data. Heat flow data and
geology correlation components use the median as opposed to the mean in deriving the estimate in unioned polygons
(Davies 2013). (c) Degree 12 spherical harmonic representation of global heat flow from observations supplemented by
predictor, units are in mW m
-2
(Chapman & Pollack 1975). (d) Global heat map derived from spherical harmonic
expansion to degree 36 of conductive heat flow data, units are mW m
-2
(Hamza et al 2007). (e) Map of differences in
harmonc representations of global heat flow by Pollack (1993) and Hamza et al (2007), units are mW m
-2
(Hamza 2007).
a) b)
c) d) e)
CHARLIE KENZIE
away from oceanic ridges, hydrothermal
activity is inhibited in older oceanic regions.
Additionally, because of the random nature of
heat flow data points, the probability of
measurement sites being preferentially
situated close to recharge zones of
hydrothermal systems is relatively small,
which is likely to provide biased estimates of
background heat flow rather than an
underestimate (Hamza et al 2007).
Hamza et al (2007) re-evaluate global heat
flow using spherical analysis that has more
emphasis on experimental data. The same
technique as the Pollack analysis is used, so
not to create biased distortions in the heat
flow map, but this still requires a significant
proportion of the data set to be accompanied
by estimated values from the empirical
predictor. However, the predictor algorithm is
set up based on experimental relationships
rather than theoretical values used in the
predictor by the Pollack analysis.

Chapman & Pollack (1975) admit that the
discrepancies between tectonic province
average heat flow for different continents and
the Polyak-Smirnov values suggest that the
heat flow estimator analysis requires some
updating. However, they maintain that the
data remains reliable for most continents. In
response to the Hamza et al (2007) re-
annalysis, Pollack & Chapman (2007) argue
that use of measured conductive heat flow
values in young seafloor areas to compute
spherical harmonics is incorrect, and maintain
that heat flow values are underestimated
because of systematic bias in measurement
sites. They argue that the Pollack analysis
from 1993 remains the best estimate of global
heat flow. Figure 2 shows the heat flow maps
from the Pollack and Hamza analyses and the
difference between them (Fig.2e).

4. Uses of heat flow maps for investigating
environmental processes
Links or relationships to environmental
processes are affected by the reliability of the
data and the analysis. Thus, the previous
section allows the reader to consider the
challenges of mapping heat flow even before
the data are associated with any geological
process.

One of the primary uses of heat flow data and
maps is to investigate the relationship
between heat flux and the age of the
lithosphere (Stein 1995; Anderson & Hobart
1976; Polyak & Smirnov 1968) (Fig.3). The
extent to which the ages of both terrestrial and
oceanic lithosphere are related to heat flux is
discussed above. These relationships are often
investigated in a local setting and then applied
to estimate heat flow across areas where there
Figure 3 (a) Secular decrease of continental heat flow
with age and its three principle components: I
Radiogenic heat, II heat from transient thermal
perturbation associated with tecto-genesis, III
background heat from deeper sources. (b) Reduced heat
flow versus mean heat flow in 17 heat flow provinces
across the globe. Bars show estimated uncertainties in
the determinations of the reduced heat flow. Both
graphs indicate relationships used to estimate the
empirical predictors for spherical analysis in Chapman
& Pollack (1975) and Pollack et al (1993). Note the
large uncertainties in both, the significance of which is
discussed in Hamza et al (2007). Graphs taken from
Vitorello & Pollack (1980).
a)
b)
CHARLIE KENZIE
are no data (Stein 1995). The heat flux-age
relationship has also been found to relate to
lithosphere thickness, since the age of the
lithosphere effects its rheology (Jaupart &
Mareschal 2005).

Heat flux is often linked to the temperature
profile with depth through a section of
lithosphere. Based on this assumption, many
physical properties of the Earths crust and
upper mantle can be linked, since they are
also temperature dependent (Pollack et al
1993). Several studies have investigated
lithospheric geotherms (Blackwell 1971;
Clark & Ringwood; Sclater & Francheteau
1970), but the use of heat flux to investigate
the geotherm assumes that heat transfer
occurs only by conduction and not by
convection (Pollack et al 1993; Stein 1995).
Thus heat flow maps can only give
indications of the temperature profile until a
depth at which conduction loses out to
convection. Pollack et al (1993) presume this
occurs at the boundary of the lithosphere, and
hence interpret heat flow data related to
processes only within the lithosphere.

Stein (1993) discusses the thermal models of
oceanic lithosphere to a greater degree. The
decrease in heat flow and increase in seafloor
depth with age prompted two classes of
models. The half-space model, where depth
and heat flow vary as the square root of age
and the reciprocal of the square root of age,
respectively (Davis and Lister 1974).
Secondly, the preferred plate model (Langseth
et al 1966; Louden 1989), which suggests that
the lithosphere behaves as a cooling boundary
layer until it reaches ages at which the effects
of the lower boundary cause the depth and
heat flow curves to flatten and vary more
slowly with age (Stein 1993). The asymptopic
plate thickness to which the lithosphere
evoloves corresponds to the depth at which
the additional heat is supplied from below.
The temperature variations cause changes in
ocean bathymetry.

As such theories progressed, heat flow maps
have allowed studies of lithosphere thickness
to be linked with heat flow over the entire
globe. On the back of their heat flow analysis
in 1975, Pollack & Chapman (1976) suggest
that the lithosphere is less than 100 km thick
over most of the globe, but thickens
appreciably and becomes more viscous
beneath the Pre-Cambrian shields and
platforms, characterized by regions of low
heat flow. The initial Chapman & Pollack
analysis (Fig.2c) revealed that oceanic ridge
systems, marginal basins of the West Pacific,
Alpine Europe, and the American Cordillera
are all dominated by heat flow highs (Fig.2)
and low heat flow regions include all the
major shields and platforms and the oldest
oceanic regions (Fig.2). These observations
are implicit in the theory that high heat flow is
associated with thin, young, newly generated
oceanic crust at mid ocean ridges (Davies
1999) and that old, archaic continental
lithosphere is associated with a cold and thick
lithosphere. More recent heat flow maps have
high enough resolution to determine heat flow
patterns of local tectonic units (Fig.2b) and
clearly show high areas of heat flow
associated with oceanic regions and divergent
plate boundaries (Davies 2013).

In contrast, some of the first theories of global
heat flow (Bullard 1939, 1954; Revelle &
Maxwell 1952) presumed that heat flux would
be broadly similar across the continents and
oceans, with some suggesting that heat flow
under continents would actually be larger due
to radiogenic elements within the crust
(McDonald 1959). Recent global heat flow
data and maps (Chapman & Pollack 1975;
Hamza et al 2007; Davies 2013; Davies 1999)
show that heat flow through the ocean regions
is roughly twice that of continental regions
(Fig.2). The resolution of this puzzle has been
that in the oceans, heat is transported from the
deep interior by the mass motions involved
with seafloor spreading, not just by
conduction (Davies 1999).

Another important recognition from early heat
flow studies, show that the heat flow
distribution across subduction zones are
dominated by patterns of low flux values from
the trench axis to the volcanic arc, high values
over the volcanic zone and consistent flux
values in the back arc region similar to those
for major ocean basins of a similar age
(Anderson 1980). The heat flow of the
incoming plate is therefore lower than
expected, and is attributed to increased
CHARLIE KENZIE
hydrothermal circulation from flexure (Stein
1993). However, more recent studies (Stein
2003) suggest that earlier studies were bias
because many of the heat flow data were
collected across western Pacific subduction
zones, which involves old, and therefore cold,
crust. A re-appraisal of heat flow data by
Stein (2003) suggests that there are no
significant differences between heat flow near
the trench for subducting crust and the global
averages for the same age crust.

A further observation of global heat flow
mapping is the inferred heatflow anomalies
associated with hotspot swells. Von Herzen
et al (1982) observed that heat flow on the
Hawaiian swell was higher than predicted for
the surrounding sea floor, and attributed this
to the elevated heat flow expected for
hotspot reheating model. Several other
studies in other hotspot localities related
inferred heat flow anomalies in a similar
way. However, more recent and rigorous
analyses (DeLaughter & Stein 2005; Stein &
Stein 1992; Stein 1993) claim these
anomalies were overestimated because
reference thermal models predicted greater
depths and lower heatflow than was typical of
lithosphere older than 70 Ma. Models derived
by joint fitting of heatflow and bathymetry
data, show that at most the heatflow observed
at Hawaaii is only marginally higher than
expected and that a similar situation is true of
the inferred anomalies at Bermuda, Cape
Verde and Crozet.

5. Uncertainties of heat flow maps
The potential application of heat flow maps in
understanding environmental processes is
clearly varied. However, many question the
reliability of any such application. Firstly,
Stein (1995) explains the full extent of
potential measurement error. As already
mentioned, drilling of the holes causes
frictional heating. Additionally, most heat
flow sites, especially from older parts of the
record, ignore local factors such as
topography, sedimentation rates, and surface
temperature changes, which may disturb the
heat flux. These uncertainties exist for real
observations and are almost certainly
compounded once empirically estimated data
is substituted for large unsurveyed areas.
These uncertainties are particularly prevalent
in oceanic regions (Stein 1993) because
processes such as hydrothermal circulation
are not well studied. Even in continental
areas, where processes such as radiogenic
heat and water circulation have been well
studied, they redistribute heat so extensively
that it is difficult to calculate corrections for
total heat flow, especially across a large area
(Stein 1993).

Jaupart & Mareschal (2005) discuss the
reliability of using heat flow as a constraint
for studies of the lithosphere. Studies into the
thickness of lithosphere are "#$$%&'&()&*
+, *-)- ./0' 0)1&/ 2&0$1,"34-% 40(")/-3()"
"#41 -" "&3"'34 0/ 2&03* *-)-5 However,
different geophysical methods provide
constraints on different parts of the boundary
layers and hence cannot be compared without
care. Additionally, heat flow is not sensitive
to the same parts of the boundary layer in
transient and steady-state conditions and
furthermore, different dynamics of mantle
convection and secular cooling are likely to
affect heat flow in different ways. Thus,
Jaupart & Mareschal (2005) argue that a
purely empirical approach to heat flow data is
doomed to fail, and that at best, when used in
conjunction with other methods, heat flow
provides an insight into mantle dynamics.

As discussed above (section 3), any
interpretations of heat flow maps are
ultimately dependent on how reliable the
analyses of the data are. Even in recent
analyses (Hamza et al 2008; Davies 2013) the
results are still affected by uncertainties in
estimating values for unsurveyed areas
(Fig.3). Similarly, Harris & McNutt (2007)
suggest that our ability to resolve anomalous
basal heat flow will depend on a better
understanding of environmental conditions
and non-conductive processes. This
understanding will likely require more high-
resolution heat flow surveys coupled with
seismic reflection, observations of fluid flow
and more sophisticated analysis. Until then,
using solely heat flow data to distinguish
between thermal and non-thermal origins may
be premature (Harris & McNutt 2007).


CHARLIE KENZIE
References:
ANDERSON, R. N. "Update of heat flow in the east
and southeast asain seas." The tectonic eveolution
of southeast asian seas and islands, Geophysical
Monograph Series 23 (1980): 319-326.
ANDERSON, R. N., and M. A. Hobart. The relation
between heat flow, sediment thickness, and age in
the eastern pacific. Journal of Geophysical
Research (American Geophysical Union) 81, no.
17 (1976): 2968-2989.
ANDERSON, R. N., M. G. Lansgeth, and J. G. Sclater.
The mechanisms of heat transfer through the
floor of the Indian Ocean. Journal of Geophysics
82 (1977): 3391-3409.
BENFIELD, A. E. Terrestial heat flow in Great
Britain. Proceedings of the royal society of
mathematical physical and engineering sciences
(royal society publishing) 123, no. 1098 (1939):
428-449.
BLACKWELL, D. D. "The thermal structure of the
continental crust, in The Structure and Physical
Properties of the Earth's Crust." Geophysics
Monograph Series (American Geophysical
Union) 14 (1971): 169-184.
BULLARD, E. C. Heat flow in South Africa.
Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences
(Royal Society Publishing) 173, no. 955 (1939):
474-502.
BULLARD, E. C. "The flow of heat through the floor
of the Atlantic ocean." Proceedings of the Royal
Society 222 (1954): 408-422.
CHAPMAN, D. S., and H. N. Pollack. Global Heat
Flow, A New Look. Earth and Planetary
Science Letters (Elsevier Scientific Publishing
Company) 28 (1975): 23-32.
CHAPMAN, D. S., and K. Furlong. Continental heat
flow-age relationships. Eos Transactions,
American Geophysical Union (John Wiley &
Sons), 1977.
CLARK, S. P., Jr., and A. E. Ringwood. "Density
distribution and constitution of the mantle." Rev.
Geophysics 2 (1964): 35-88.
DAVIES, G. F. "Heat Flow." In Dynamic Earth:
Plates, plumes and mantle convection, by G. F.
Davies, 82-85. Cambridge: Cambridge University
Press, 1999.
DAVIES, J. H. Global map of solid Earth surface heat
flow. Geochemistry Geophysics Geosystems
(AGU and the Geochemical Society) 14, no. 10
(2013): 4608-4622.
DAVIS, E. E., and C. R. B. Lister. "Fundamentals of
ridge crest topography." Earth Planetary Sciences
21 (1974): 405-413.
JAUPART, C., and J. -C. Mareschal. "Heat flow and
thermal structure of the lithosphere."
http://www.earth.ox.ac.uk/. December 2, 2005.
ftp://ftp.earth.ox.ac.uk/pub/tony/TOG/MS105_Jau
part_Text.pdf (accessed January 28, 2014).
LANGSETH, M. G., X. Le Pichon, and M. Ewing.
"Crustal structure of the mid-ocean ridges, 5, Heat
flow through the Atlantic Ocean floor and
convection currents." Journal of Geophysics 71
(1966): 5321-5355.
LOUDEN, K. E. "Marine heat flow data listing,
Handbook of seafloor heat flow." CRC Press,
1989: 325-485.
MCDONALD, G. J. F. "Calculations on the thermal
history of the earth ." Journal of Geophysics 64
(1959): 1967-2000.
POLLACK, H. N., and D. S. Chapman. "Comment on
"Spherical harmonic analysis of earth's
conductive heat flow" by V.M. Hamza, R. R.
Cardoso and C.F. Ponte Neto." International
Journal of Earth Science (Springer-Verlang) 97
(2007): 227-231.
POLLACK, H. N., S. J. Hurter, and J. R. Johnson.
Heat flow from the earth's interior: analysis of
the global data set. Reviews of Geophysics
(American Geophysical Union) 31, no. 3 (1993):
267-280.
POLYAK, B. G., and Ya. B. Smirnov. Relationship
between terrestial heat flow and the tectonics of
continents. Geotectonics, 1968: 205-213.
REVELLE, R., and A. E. Maxwell. "Heat flow through
the floor of the eastern north pacific ocean."
Nature (Nature) 170 (1952): 199-200.
SCLATER, J. G., and J. FRANCHETEAU. The
implication of terrestiral heat flow observations
on current tectonic and geochemical models of
the crust and upper mantle of the Earth. Journal
of Geophysics (Astronomy Society) 20 (1970):
509-542.
SCLATER, J. G., J. Crowe, and R. N. Anderson. On
the reliability of oceanic heat flow averages.
Journal of Geophysics 81 (1976).
STEIN, C. A. "Heat flow and flexure at subduction
zones." Geophysical Research Letters 30, no. 23
(2003): 2197-2200.
STEIN, C. A., and R. P. Von Herzen. Potential effects
of hydrothermal circulation and magmatism on
heatflow at hotspot swells. The Geological
Society of America (The Geological Society of
America), 2007: 261-274.
VITORELLO, I., and H. N. Pollack. On the variation
of continental heat flow with age and the thermal
evolution of continents. Journal of Geophysical
Research 85, no. B2 (1980): 983-995.

You might also like