You are on page 1of 16

sjtg_373 328..

343

doi:10.1111/j.1467-9493.2009.00373.x

Are any granite landscapes distinctive of


the humid tropics? Reconsidering
multiconvex topographies
Piotr Migoń
Department of Geography and Regional Development, University of Wrocław, Wrocław, Poland

Correspondence: Piotr Migoń (email: migon@geogr.uni.wroc.pl)

Geomorphologists have long debated over whether certain geomorphic types of landscapes may
evolve only in certain climatic conditions, with granite terrains often chosen to illustrate such
climate–relief relationships. The view deeming inselberg landscapes in particular to be distinctively
tropical is difficult to sustain in the light of research demonstrating both the wide global distribution
of inselbergs and evident rock control on their origins and development. However, another
geomorphic type of granite landscapes has emerged as possibly specific to low latitude terrains,
characterized by efficient deep weathering. In the context of the ongoing debate, this paper
considers the case of multiconvex topography, consisting of low hills weathered throughout, which
is widely distributed in South America, Africa and Southeast Asia but not yet reported from middle
or high latitudes. Multiconvex landscapes require that weathering operates efficiently in all topo-
graphic settings and that dissection proceeds to maintain relative relief. Hence they tend to be
associated with areas subject to moderate uplift. There is little chance for this type of relief to
survive major environmental changes towards aridity or cooling because stripping of the weath-
ering mantle will not be compensated by saprolite renewal.

Keywords: geomorphology, granite, multiconvex relief, tropical environments

Introduction
That landforms and geomorphological landscapes reflect an interplay of various endoge-
nous and exogenous factors and forces is a truism. The challenge is to identify those
various controls, assess their cause and effect aspect, and establish the relative weights
of each controlling factor in a given environment and thus their hierarchy. The growing
appreciation of complexity and nonlinearity in geomorphic systems makes this more
difficult, despite the many conceptual advances. In particular, the question of whether
the appearance of a rock-cut erosional landscape is attributable to the role exerted by
the properties of the rock itself or to the external environment remains an unresolved
discussion in geomorphology, and one with profound implications. If a direct relation-
ship between environment and landform assemblages can be demonstrated, then these
assemblages acquire the status of indicators, or proxy records, of those environments
and are the evidence of major environmental change. Rock-cut landscape inheritance,
whether from a more recent periglacial period or more distant times, if proved beyond
doubt, bears on a number of issues in geomorphology including the age of the land-
scape, lifetimes of individual landforms and reasons for survival.
The issue of possible climatic control via different exogenous processes of weather-
ing, erosion and denudation has been raised in respect to many common rock types. An
evaluation of structural versus climatic control features in each summary of karst
geomorphology such as, most recently, by Ford and Williams (2007) and, some two
decades ago, by Jennings (1984). While the weight of current opinion seems to favour
lithology and structure as primary controls, the climatic factor cannot be dismissed too

Singapore Journal of Tropical Geography 30 (2009) 328–343


© 2009 The Author
Journal compilation © 2009 Department of Geography, National University of Singapore and Blackwell Publishing Asia Pty Ltd
Granite landscapes distinctive of the humid tropics 329

easily. Likewise, the rock versus environment issue is strongly present in the general
sandstone geomorphology by Young and Young (1992), who point out that our under-
standing of sandstone geology and sandstone properties is often insufficient to seriously
embark on the debate, and that many previous hypotheses about the primary role of
climate in shaping sandstone terrains failed the test of time after local geology began to
receive better recognition from geomorphologists.
However, nowhere perhaps has the discussion been so intense as in the field of
granite geomorphology. Several works have proposed elaborated morphoclimatic
schemes linking particular landform assemblages with specific environments (e.g.
Wilhelmy, 1958; Czudek et al., 1964; Godard, 1977), in particular ‘tropical’ assemblages
such as weathered plains or inselbergs. Other works address the subject with more
caution (e.g. Thomas, 1974; 1976; Gunnell, 2000) and preferred morphographic rather
than morphoclimatic classification of the variety of granite landscapes (Lidmar-
Bergström, 1995). Finally, there have been voices in which the role of climate was
evidently downgraded (Twidale, 1982). Since the peak of the debate in the 1960s and
1970s, the interest of geomorphologists has in more recent years shifted towards
mechanisms and rates of processes rather than landforms themselves. Valuable insights
into rates of weathering, particularly chemical denudation at a catchment scale have
been obtained (e.g. White & Blum, 1995; Oliva et al., 2003; Braun et al., 2005), while
modelling exercises increased understanding of the relationships between relief, physi-
cal erosion and chemical denudation (e.g. Riebe et al., 2004; Burke et al., 2007; Ferrier
& Kirchner, 2008). These advances in tropical and theoretical geomorphology provide a
context to return to the age-old question of rock versus climatic control on granite
geomorphology.

Diversity of granite landscapes


The diversity within granite geomorphology is enormous. In this respect granite terrains
do not differ from karst terrains which are also known to occur in a whole range of
variants (Jennings, 1984; Ford & Williams, 2007). Thomas (1974) offered an early
attempt to account for this variability and distinguished the following geomorphic types:
(a) multiconcave (basin form) landscapes, with individual basins being enclosed, par-
tially enclosed or dissected; (b) multiconvex (dome form) landscapes, which can be
further subdivided into those with dominant weathered convex compartments and
those with rock-cored compartments; (c) stepped or multistorey landscapes; (d) plains.
Another attempt worth recalling, devised by Lidmar-Bergström (1995) for the geomor-
phology of Sweden, was not specifically designed for granite terrains but, because of the
abundance of granite in the study area, provides a valuable contribution to the general
granite geomorphology. Her typology includes plains (peneplains), dissected plains,
plains with residual hills, undulating hilly lands, joint valley landscapes, large-scale joint
valley landscapes, deeply incised highlands and tectonic horst and graben relief.
Informed by such work and supplemented by my own observations, I have sug-
gested a comprehensive classification scheme (see Migoń, 2006). This classification of
regional samples includes nine principal variants: plains, plains with residual hills
(inselbergs), multiconcave relief, multiconvex relief, plateaux, dissected plateaux, joint
valley landscapes, stepped topography and all-slopes topography (Figure 1). It is delib-
erately descriptive, to avoid premature assertions about relationships between land-
forms and controlling factors. Obviously, it may be argued that certain categories are
difficult to distinguish from each other or are overlapping (e.g. plains and plains with
330 Piotr Migoń

Plain

Plain with residual hills

Multiconcave landscape

Multconvex landscape

Plateau

Dissected plateau

Joint-valley landscape

All-slopes topography

Bedrock Saprolitic cover

Figure 1. Geomorphological types of granite terrain (named after Migoń 2006).

residual hills) and that objective delimitation of the various morphographic types is
problematic. However at the scale of 101–102 km2 these landscapes are sufficiently
distinctive from one another to serve as a basis for further discussion. Before focussing
on multiconvex topography, the subject of particular interest in this paper, the other
eight variants are briefly characterized immediately below.

Eight principal variants of granite landscapes following Migoń (2006)


Plains. Plains are extensive monotonous surfaces of very low to nonexistent relief
located close to the regional base level, with drainage lines at the level of the plain.
Plains can be subdivided into two main variants: rock-cut and weathered. In the latter
case, the topography of the weathering front may be quite variable, and stripping of the
saprolite would reveal a landscape, possibly hilly, of the Grundhöcker relief type as
envisaged by Büdel (1957). Timescales of plain formation are unknown, but a few tens
of millions of years are probably required to arrive at the stage of a featureless landscape
cutting across structures. Plains occur widely across old granite shields, from the arid
interiors of Africa and Australia to formerly glaciated terrains in North America and
Fennoscandia (Twidale, 1982; Lidmar-Bergström, 1995).

Plains with residual hills. The monotony of many plains is interrupted by the occurrence
of residual hills. These may be very high and visible from a long distance, fulfilling the
Granite landscapes distinctive of the humid tropics 331

descriptive criteria for an inselberg, or low and subdued. Again, shield settings abound
in plains with residual hills, although such landscape facets of limited extent (<100 km2)
may be present within rejuvenated ancient mountain belts. Inselbergs within plains
may be of different origin and form due to long-term lowering of the surrounding
surface, scarp retreat, or alternating etching and stripping (Selby, 1977; Thomas, 1978).
Inselberg-studded plains are known from present-day deserts (e.g. Namibia) and dry
savannas (e.g. East Africa), as well as high latitude terrains (e.g. northern Finland).

Multiconcave topography. The variability within the characteristic relief in the multicon-
cave topography is considerable. Basins may be of different sizes, from much less than
1 km2 to 10–100 km2, and nested basins may occur; local relief is typically a few tens of
metres. Some basins stand in isolation, drained via narrow breaches in their rocky rims,
whereas others are interconnected and form spatially complex patterns. Their flat floors
are often marshy and poorly drained. Elongated channelless alluviated variants are
widespread in the tropics and are known under a variety of local names such as dambo
in central Africa or bolis in Sierra Leone (Mäckel, 1985). The knock and lochan land-
scape is a distinctive high latitude variant present in formerly glaciated terrains of
Scotland, Greenland and Canada. Similar landscapes are known from the eastern
highlands of Australia.

Plateaux. Plateaux differ from plains in that they occupy a different position in respect
to the regional base level: they are elevated surfaces of low relief, separated from their
surroundings by steep marginal slopes. Plateaux may be of tectonic origin and attain the
height of many hundreds of metres. In other cases, granite plateaux slope down via less
steep denudational escarpments, related to the apparently higher resistance of granite in
relation to the country rock. Plateaux usually have some relief and are hardly ever flat,
as are base-levelled plains. Some are gently rolling and the dominant landscape facet is
then a long planar slope connecting flattened summits and valley floors; others may
show considerable roughness, with abundant tors and boulders. A good example of
plateau relief is offered by the French Massif Central (Migoń, 2006).

Dissected plateaux. Dissected plateaux are a variant of the above category. Their promi-
nent geomorphic feature is a close juxtaposition of surfaces of low relief and deeply
incised valleys. There may occur a specific zonation within an upland, with a dissected
outer zone and a more compact inner zone not yet reached by headward erosion.
Dissection may have been accomplished by fluvial or glacial erosion, or by both.
Impressive dissection of plateaux can be admired in numerous European uplands
including the Cairngorms in Scotland (Sugden, 1968) and Serra da Estrela in Portugal
(Vieira, 2008).

Joint valley topography. This specific morphological type is distinctive owing to the close
relationship between fracture patterns and negative relief (Lidmar-Bergström, 1995;
Johansson et al., 2001). There is a recognizable lattice-like pattern of linear concave
landforms overprinted on major fractures zones. These landforms may be fluvial valleys,
narrow clefts and gorges, defiles and highly elongated basins. There are several differ-
ences between joint valley topography and dissected plateaux. First, dissection is fluvial
or glacial, whereas fracture exploitation in the joint valley landscape is predominantly
by selective weathering with fluvial erosion playing only a subordinate role. Second,
dissection proceeds from the outer parts of a plateau inward, resulting in a hierarchical
332 Piotr Migoń

pattern of drainage lines, whereas the pattern of fracture-guided valleys does not show
a hierarchy. Third, the adjustment to structure is less perfect in the dissected plateaux.

All-slopes topography. ‘All-slopes’ highlights the absence of flat terrain, neither along
divides nor valley floors, and the overall dominance of sloping surfaces of considerable
gradient. All-slopes topography is thus intimately associated with mountain regions and
the most efficient dissection by fluvial and/or glacial processes. All-slopes relief occurs in
many variants, depending on the relative intensity of each process. Mountain crests are
often serrated, with individual joint-bound compartments sculpted into pinnacles,
turrets and angular towers, separated by deep clefts. In certain locations, steep-sided
domes may rise from water divides, otherwise slopes tend to be straight. Mountainous
granite topography is widely present around the world, particularly within young
orogenic relief, irrespective of the present climatic zone (Twidale, 1982; Migoń, 2006).

Stepped topography. This type of topography includes a step-like rise in elevation from
the foothills toward the crest, accomplished by the alternating occurrence of steep scarps
at different altitudes and gently sloping surfaces between them. However, the problem
of scale is particularly disturbing here. Individual steps, if wide enough, may have their
own local relief that fulfils the criteria for multiconcave or perhaps multiconvex relief.
This special type of relief has been named by Wahrhaftig (1965) in the Sierra Nevada,
California, but few examples have been described from other parts of the world.

Multiconvex weathered landscape as a humid tropical granite relief


The above review of the occurrence of eight of the nine different types of granite
landscapes classified in Migoń (2006) reveals that many have a global distribution, from
equatorial latitudes to circumpolar environments, although plateaux and dissected
plateaux appear to be underrepresented in low latitudes. The remaining variant, the
hilly multiconvex topography, appears to be the exception that so far has not been
reported from middle or high latitudes.
The major components of multiconvex topographies comprise rather closely spaced
and irregularly distributed hilly compartments, typically a few hundred metres in
diameter, separated by narrow valleys and channelless elongated depressions, with little
intervening flat land. Individual hills tend to be oval in plan and convex in shape, hence
the descriptive name for this type of geomorphology (Thomas, 1974). The relative relief
within multiconvex relief is generally small, hardly exceeding 100 m, and typical slope
inclination is 20–25°. Although a certain proportion of hills may show an exposed rock
core, the majority are weathered to a great depth, or even throughout (Figure 2).
Generalizing about the depth of weathering is difficult as available exposures reveal only
a portion of the saprolite. Depths of up to 60 m have been reported in south China (Lan
et al., 2003), similar to those reported from a number of African localities (Thomas,
1966; 1994). Braun et al. (2005) reached bedrock at the depth of 40–50 m at the level
of the footslope surface in their Nsimi catchment study area in Cameroon. In southeast
Brazil a depth of weathering of up to 20–30 m is known (Coelho Netto, 1999). On the
other hand, isolated boulders and boulder piles may be visible on hillslopes, especially
in their lower parts, but very rarely on the summits.
It is hard to be entirely confident about the geographical distribution of multicon-
vex weathered granite relief, since the literature is dispersed and different descriptive
terms are used in different languages (Migoń, 2006). However, the Portuguese meias
Granite landscapes distinctive of the humid tropics 333

Figure 2. Weathered hilly compartment in the multiconvex landscape around Mylliem, Meghalaya plateau,
India, November 2006 (author’s photo).

laranjas (or ‘half-orange’) hills and the French demi-orange paysage are pertinent to this
type of landscape (Godard, 1977; Thomas, 1994). One region where the multiconvex
morphology has been characterized in much detail, including the processes at work, is
the coastal belt of southeast Brazil, in particular the intramontane trough of the
Paraiba do Sul (Coelho Netto, 1999). Deeply weathered multiconvex topography has
been reported from the northern Amazon Basin (Dubroeucq & Volkoff, 1998). Godard
(1977) cites further examples from the Ivory Coast and Guyana, and Thomas (1994)
mentions occurrences in equatorial Africa (Sierra Leone and Cameroon) and Hong
Kong. Scattered literature suggests that the multiconvex landscape is widespread in
southeast China beyond Hong Kong, covering large areas in the Guangdong province
(Xu, 1996). Another location is the Meghalaya plateau, northeast India, where granite
intrusions have been sculpted into an array of deeply weathered hills with occasional
boulders on their surfaces (Prokop, 2007). In the southern Deccan multiconvex land-
scapes are associated with the western, more humid part of the subcontinent and give
way to convexo-concave relief and eventually multiconcave pedimented landscapes
along the west–east transect (Gunnell & Bourgeon, 1998; Gunnell, 2000). The multi-
convex morphology is also observed in certain parts of the Sierra Madre del Sur in
Mexico (Figure 3) and in the humid part of Madagascar (Raunet, 1985). Figure 4
summarises the locations of multiconvex landscapes as reported in the literature and
from my own field observations.
Thorough weathering of hilly compartments in the multiconvex relief influences its
geomorphic development and the nature of surface processes. In both southeast Brazil
(Meis & Monteiro; 1979; Coelho Netto, 1999) and southeast China (Xu, 1996) it has
been observed that convex slopes are often incised by steep-sided hemispherical hollows
of various size, from a few tens to >100 m across, with a depth that may be in excess of
20 m (Figure 5). Locally the hollows occur on opposite sides of a hill, accounting for its
very scalloped appearance. Downslope, the hollows grade into low-angle planar sur-
faces or fans extending for up to 100–200 m, not uncommonly gullied and dissected.
This hollow and ramp association, recognized for the first time in southeast Brazil, has
334 Piotr Migoń

Figure 3. Multiconvex landscape in the southern Sierra Madre del Sur, near Xaltianguis north of Acapulco,
Mexico, December 2004 (author’s photo). Note the numerous boulders at the surface, which are exposed
corestones, and close spacing of the hills.

Meghalaya
Southeast China
Sierra Madre del Sur
Karnataka
Guyana Sierra Leone Cameroon
Amazon

Malawi

Serra do Mar Madagascar

0 2500 5000 km

Figure 4. Global distribution of multiconvex landscapes (based on Shroder, 1976; Godard, 1977; Meis &
Monteiro, 1979; Raunet, 1985; Thomas, 1994; Xu, 1996; Dubroeucq & Volkoff, 1998; Coelho Netto, 1999;
Gunnell, 2000; Migoń, 2006; and my own observations).

been named the rampa complex (Meis & Monteiro, 1979), while the Chinese-named
benggang, meaning ‘collapsing hill’ (Xu, 1996) appears to be an Asian equivalent. Slope
hollows and ramps or fans below testify to the important role of landslides in shaping
multiconvex terrain. Since granite is considerably weathered, the significance of relict
discontinuities as potential instability planes diminishes and curved slip surfaces develop
within the saprolitic material. Consequently slides are typically rotational and, if local
relief allows, transform into earthflows that contribute to building up footslope surfaces
and filling the valleys.
Other important surface processes in the multiconvex topography are gullying and
soil creep. The former is closely linked with landsliding and the relationships are
Granite landscapes distinctive of the humid tropics 335

Figure 5. Multiconvex topography of Hong Kong shows the evidence for the role of landslides and subsequent
gullying in the geomorphic development, January 2008 (photo courtesy of Young CY Ng). Note two persons on
the footpath (lower centre) for scale.

two-directional. On one hand, gully development typically follows landsliding, being


focused on newly created steep slopes in the landslide head where unprotected
weathered bedrock is exposed (Figure 5). Similarly, extensive gullying is recorded
within failed masses, and footslope rampas are commonly dissected. On the other hand,
new gullies alter subsurface hydrology, promote flow towards gully heads, focus piping
and thereby lead to further slope instability (Coelho Netto et al., 1988; Oliveira, 1990).
An important aspect here is the age and causes of gullying. Although they may develop
spontaneously within head scarps of a landslide (Figure 5), an acceleration of gully
erosion following land clearance is a well known phenomenon in many parts of the
world (Valentin et al., 2005). The role of soil creep is evidenced by thick colluvial
materials which cover the saprolite by a mantle 2–3 m thick. Stonelines may be abun-
dant, indicating a very complex history of creep, shallow slides, surface wash and
bioturbation. It may be hypothesized that the efficacy of soil creep increases downslope
due to overall slope convexity, and that a sort of equilibrium is maintained.

Long-term evolution and survival of multiconvex landscapes

The geomorphic evolution of multiconvex weathered terrain is accomplished by two


sets of antagonistic processes. Deep weathering breaks down the bedrock and increases
the thickness of the saprolitic mantle, while surface processes remove the weathered
material thereby reducing the thickness of the weathering mantle. Unfortunately, there
is very little data on the rates with which the different processes operate in the
multiconvex landscapes over the long-term. An early global survey of granite weath-
ering rates by White & Blum (1995) includes only one site from the humid low latitudes
– in the mountainous setting of Puerto Rico. A more recent and much more compre-
hensive dataset by Oliva et al. (2003) is rather enigmatic about the topography and
landforms within the investigated catchments; furthermore, the reported rates are those
of solute exports from the entire catchments and not of weathering front descent, which
336 Piotr Migoń

is more important in the context of this study. The rates of weathering front advance,
seen in the light of the rather sparse extant database, appear to vary by one order of
magnitude, from anything up to about 50 m in one million years (Thomas, 2006). A
recent study by Braun et al. (2005) from Cameroon, in a setting similar to the multi-
convex one dealt with in this paper, yielded a figure of 4.7 mm/ka (i.e. 47 m/Ma)
obtained from short-term measurements. Their comparative survey indicates that this
numerical value is one of the lowest recorded figures, but higher rates obtained from
other areas of steeper relief are accompanied by much higher rates of physical erosion,
hence weathering profiles do not thicken over time. To receive meaningful figures for
landsliding and gullying is even more challenging, given their highly episodic nature.
Attempts to quantify slope lowering by landslides have concerned highly dissected
young mountain ranges and usually a bedrock other than granite (Hovius et al., 2000),
which limit their relevance to this study.
Landsliding in multiconvex topography may be extensive and individual rainstorm
events are known to trigger hundreds or even thousands of slope failures over a limited
territory (Xu, 1996). Severe human interference with natural conditions, mainly wide-
spread vegetation clearance for agriculture or housing, exacerbates the problem and
increases vulnerability to mass movement. However, slope failures are the natural
component of a geomorphic system in the multiconvex terrain and occur without a
human contribution. In southeast Brazil extensive dating of colluvial sediments has
been performed and it is now clear that the history of landsliding, colluvial deposition,
and subsequent gullying may be traced back at least to the Late Pleistocene (Meis &
Moura, 1984; Coelho Netto, 1999). In another deeply weathered tropical terrain, in
Malawi, Shroder (1976) mapped the occurrence of more than 250 individual landslide
features of apparently very different ages. Landslides are thus both a component of
short-term landform development, possibly influenced by human pressure to some
extent, and a significant contributor to the long-term relief evolution.
The key issue is therefore whether a multiconvex weathered terrain can be main-
tained and how. Since hilly compartments apparently lose mass by landsliding and other
denudational processes, the survival of the regional landscape requires that weathering
operates efficiently in all topographic settings and that dissection proceeds to sustain
relative relief. Otherwise the multiconvex landscape will be reduced to a rolling plain,
with occasional ‘shield inselbergs’ – remnants of the unweathered cores of the long gone
hills, and widely distributed sedimentary veneer from past landslides. This requirement
dictates a specific relationship between deep weathering and physical erosion, with low
to moderate rates of the latter. Recent modelling exercises by Gabet (2007) and Ferrier
& Kirchner (2008) suggest that these processes will not balance each other infinitely and
that with rates of physical erosion increasing, the rates of chemical weathering will
diminish. Since landslides are potent agents of denudation and can easily remove the
whole thickness of weathering mantle, deep weathering – the other side of the equation
– has to be particularly efficient. Weathering systems operate efficiently in most equa-
torial, humid and subhumid tropical, or even subtropical, regions (see Oliva et al., 2003),
which might explain why multiconvex landscapes are so widely reported from these
areas. Referring to general concepts in geomorphology of low latitudes (Thomas, 1994),
it seems that multiconvex landscapes may survive under the dynamic/episodic etchpla-
nation regime, whereas their occurrence indicates long-term landscape lowering.
The current status of dissection and degradation of multiconvex relief varies from
one area to another. In southeast Brazil valley floors are extensively alluviated, with a
fill up to 20 m thick, and in certain places a ‘drowned landscape’ appears to have
Granite landscapes distinctive of the humid tropics 337

91O 50’ 00.00”E 91O 55’ 00.00”E

25O 30’ 00.00”N

25O 30’ 00.00”N


A

0 2 4 km

91O 50’ 00.00”E 91O 55’ 00.00”E

Figure 6. Diversity of relief in the Mylliem granite batholith, Meghalaya, India, (image courtesy of Paweł
Prokop) showing A – multiconcave landscape and B – multiconvex landscape; the thick white line indicates the
boundary of the granite massif.

recently evolved from a former multiconvex relief (Coelho Netto, 1999). This indicates
inability of the local fluvial system to export products of hillslope denudation, at least in
the short-term, which may suggest the role of human-accelerated denudation consis-
tent with widespread land clearance and gullying along terrain tracks. However, in
Karnataka, southern India, the flatness of the terrain between the hills is not so much
because of massive sedimentation, which is only 1–2 m thick, but rather due to weath-
ering and regolith redistribution (Gunnell & Bourgeon, 1998). The situation in the
Mylliem batholith in the Meghalaya plateau is different. As can be discerned in Figure 6,
the southwest part of the granite terrain has been reached by headward erosion initiated
at the southern escarpment of the Meghalaya, resulting in high relative relief and steep
slopes (Prokop, 2007). Major rivers have cut down to bedrock and given the steep
gradient of the main Umiew river (55 ‰) will continue to incise at a high rate. In the
area under the contemporary influence of heavy erosion by the Umiew, the multi-
convex landscape is best developed.
Hence, the survival of the multiconvex landscapes appears to be dependent on both
environmental and geotectonic factors. Warm and humid conditions favour deep
weathering, which is able to keep pace with landsliding, is the decisive surface process
for this peculiar type of morphology. However, long-term regional base-level lowering
is necessary to maintain the relief, hence a long-term relative surface uplift is indis-
pensable. Broad topographic swells and plateaux in nonorogenic settings provide the
most appropriate locations for the multiconvex landscape to develop. If the rate of uplift
is too high, balance between weathering and landsliding is lost, weathering becomes
nonsaprolitic and an all-slopes terrain evolves. Thomas (1997) suggests the uplift rate of
0.5 mm yr-1 as the threshold between saprolitic and nonsaprolitic weathering. The
above requirements match the distribution of the majority of multiconvex weathered
terrains present near topographic margins of elevated plateaux (Meghalaya), rift shoul-
ders (Nyika, Malawi), broad swells related to intraplate uplift (Sierra Leone), or great
escarpments along passive margins (southwest Deccan and southeast Brazil).
338 Piotr Migoń

Figure 7. Degradation of a multiconvex weathered landscape may leave residual boulders and low rock hills
such as those widely present in central Portugal, including this example from near Évora, Alentejo, January
2008 (author’s photo).

The necessary combination of efficient weathering and moderate uplift rate provides
at least a partial answer as to why the multiconvex weathered terrain should be
specifically tropical. It is hypothesized here that the rates of deep weathering outside the
humid and semihumid tropics, if compared with the rates of uplift and of other surface
processes, are too low and insufficient to maintain thick saprolites. Therefore, dissected
plateaux rather than a complicated multiconvex topography evolve in response to
regional uplift. Even if a former multiconvex landscape happens to occur outside the
tropical belt as the consequence of a major environmental change or continental drift,
then continuous stripping of the weathering mantle may not be compensated by
saprolite renewal, as suggested Ferrier & Kirchner (2008). In the long-term, an etched
topography of the weathering front will be exposed, with boulder piles, tors and low
domes as the only remnants of deeply weathered hills. Such a topography is typical for
many granite terrains in both arid and semiarid zones such as in the Joshua Tree
National Park, California (see Oberlander, 1972) or the Iberian Peninsula (Figure 7) and
in temperate central Europe (Migoń, 1996; 2006). Unfortunately, none of these offer
many clues to their antecedent geomorphology. The very nature of the multiconvex
terrain, in granite or otherwise, dictates that it is highly vulnerable to environmental
change and may simply not survive. Active multiconvex topographies in the humid
tropics would hardly have their inherited counterparts.

Inselbergs and other possibly specific tropical landscapes


Inselbergs and domes
Granite inselbergs are probably the tropical landforms in the eyes of many. Not only
were they described scientifically for the first time from the equatorial Africa (Höver-
mann, 1978, after Bornhardt, 1900), but the majority of subsequent geomorphological
studies of inselbergs have come from Africa or South America (Kesel, 1973; Thomas,
1978). The popularity of the two-stage model of the origin of inselbergs emphasizing
differential deep weathering followed by stripping of the saprolite and exposure of the
unweathered compartments (Twidale, 1964; Thomas, 1965) has led, perhaps uninten-
tionally, to even an stronger association of inselbergs with the humid tropics. This
thinking has had its implications for the interpretation of pre-Quaternary landscapes in
central and northern Europe. Occasional inselbergs, built of granite or otherwise, have
Granite landscapes distinctive of the humid tropics 339

been typically seen as inherited tropical landforms and used as solid benchmarks to
establish regional denudation chronologies (e.g. Gellert, 1970; Büdel, 1978).
However, this view is no longer tenable. The distribution of inselbergs far exceeds
the boundaries of the intratropical belt (see Twidale, 1982; Migoń, 2006), and inherit-
ance from past tropical conditions, claimed in a number of studies from mid to high
latitudes, is often an assumption rather than a finding from scrupulous research.
Inselbergs are as common within tropical savannas as they are in many of the world’s
deserts. For instance, they occur widely in the Namib Desert, where no evidence exists
for their inheritance from past humid conditions (Selby, 1977). By contrast, they show
marked lithological and structural control on both their location and form and can be
perfectly explained in terms of ongoing differential denudation, without the necessity to
recourse to past deep weathering (Selby, 1982; Migoń & Goudie, 2003). Likewise,
domed granite residuals are by no means specific to the tropics and may occur wherever
massive rock compartments exist. Examples of massive domes currently emerging from
mountain ranges may be forwarded from the Sierra Nevada, California (Huber, 1989) or
Sierra da Guadarrana in central Spain (Pedraza et al., 1989).

Joint valley landscape


Joint valley landscapes are not widely distributed and only a few examples have been
discussed in more detail. Low latitude representatives include certain areas in South
Africa (Mabbutt, 1952) and Sri Lanka (Bremer, 1981). Pye et al. (1984) mention joint
and fault guidance of valleys in the Matopos batholith in southern Zimbabwe and
provide an aerial photograph showing an excellent example of joint valley landscape.
There is a distinct rhomboidal pattern of concave relief features, adjusted to
northwest–southeast, west northwest–east southeast and west–east fractures. Massive
domes and enclosed shallow basins typify higher ground between the valleys and
defiles.
However, the most elaborate account of this type of relief comes not from the tropics
but from southern and western Sweden, where it occurs widely along the present-day
coast, in granite and gneiss alike (Lidmar-Bergström, 1995; Johansson et al., 2001;
Migoń & Johansson, 2004). In terms of environmental conditions of the origin, the
occurrence of the joint valley landscape in Sweden is deceptive. There are good reasons
to believe that this peculiar landform association is exhumed from beneath the Creta-
ceous cover and that the main relief features are of Mesozoic age, as testified by relict
Cretaceous deposits within the valleys (Lidmar-Bergström, 1995). In this context one
needs to note that the Mesozoic in Europe was a period of warmth and efficient deep
weathering, often likened to the tropics today.
The probable reason why joint valley landscapes appear a rarity is that they require
a specific combination of efficient, but highly selective deep weathering acting over a
geologically short timescale. Lidmar-Bergström (1995) suggested that the joint valleys
were the first stage in the transformation of the peneplain to an undulating hilly land
and that landform development was arrested by rising sea level and deposition of
Cretaceous marine deposits, conserving the etched valleys. A longer period of etching
would not only deepen fracture-aligned concavities, but also result in their widening,
gradual reduction and fragmentation of intervening upland surfaces. Consequently, the
joint valley topography would grade into other relief types, particularly into multicon-
cave relief. The Idanre Hills in southern Nigeria exemplify an advanced stage of etching
structural discontinuities, and their present-day landscape is dominated by domes and
dissected slopes (Jeje, 1974).
340 Piotr Migoń

Summarizing, there is no unequivocal evidence that the joint valley landscape may
only form in the tropical environments, although this is not unreasonable. Too little is
known about this distinctive type of relief and far too few examples have been reported
in the literature with sufficient detail to be conclusive about its origin.

Conclusions

The diversity of granite landscapes has long called for explanation. Whereas the primacy
of rock control has been advocated by one group of researchers, others have explored
associations between characteristic landforms and their assemblages, and environmen-
tal factors. Inselbergs, tors, plains, have all been hypothesized to be specific low latitude
phenomena. Consequently, their occurrence outside the tropical belt has often been
interpreted in terms of landform inheritance and tropical legacy.
Although neither of these landforms seems distinctively tropical according to the
accumulating evidence and knowledge, one specific type of relief, the multiconvex
landscape, has not yet been reported from outside the low latitudes. Consisting of
closely spaced deeply weathered hills which may or may not have a rock core, it appears
a repetitive landscape across the equatorial and humid tropical belt, from South America
through Africa to Southeast Asia. Efficient deep weathering as the means to compensate
mass loss due to landslides and gully erosion is a prerequisite for this type of geomor-
phology to form and survive. One of the challenges is to further quantify the relation-
ships between weathering and surface denudation in multiconvex terrain, as those
established so far concern dissimilar geomorphic settings. It would also be useful if
attempts to seek such relationships are accompanied by proper consideration of geo-
morphological context.
If there is an imbalance of rates of weathering and surface denudation, whether
short- or long-term, with the latter being higher, weathered multiconvex landscape
will likely transform into a rolling plain with scattered tors and boulders, or into a
multiconcave relief. This may explain the apparent absence of multiconvex landscapes
in arid or temperate lands. Even if global environmental changes ‘moved’ the mul-
ticonvex terrains into an arid or cool climate domain, uncompensated removal of
the saprolite has resulted in its transformation beyond recognition, and its eventual
disappearance.

Acknowledgements

This paper was first delivered as a lecture to the 2006 Regional Conference of the International
Association of Geomorphologists (in association with the biannual meeting of Brazilian Geomor-
phological Union) in Goiania, Goias, Brazil. I thank the main organizers of that meeting, Edgardo
Latrubesse and Selma Simões de Castro, as well as various individuals whom I have, at different
times, had the privilege to be in the field with and discuss the origin and evolution of multiconvex
landscapes, in particular Irasema Alcántara-Ayala, Nelson Fernandes, Claudio Limeira Mello, Maria
Naise de Oliveira Peixoto, Leszek Starkel and Paweł Prokop. I also gratefully acknowledged the
helpful comments received from the SJTG reviewers in revising the paper.

References

Braun JJ, Ndam Ngoupayou JR, Viers J et al. (2005) Present weathering rates in a humid tropical
watershed: Nsimi, South Cameroon. Geochimica et Cosmochimica Acta 69, 357–87.
Bremer H (1981) Reliefformen and reliefbildende Prozesse in Sri Lanka [Landforms and geomor-
phic processes in Sri Lanka]. Relief, Boden and Paläoklima 1, 7–183.
Granite landscapes distinctive of the humid tropics 341

Büdel J (1957) Die ‘Doppelten Einebnungsflächen’ in den feuchten Tropen [‘Double planation
surfaces’ in the humid tropics]. Zeitschrift für Geomorphologie NF 1, 201–28.
Büdel J (1978) Das Inselberg – Rumpfflächenrelief der heutigen Tropen und das Schicksal seiner
fossilen Altformen in anderen Klimazonen [Inselbergs and denudation surfaces in the con-
temporary tropics and the history of relict inherited landforms in other climatic zones].
Zeitschrift für Geomorphologie NF, Supplementband 31, 201–28.
Burke BC, Heimsath AM, White AF (2007) Coupling chemical weathering with soil production
across soil-mantled landscapes. Earth Surface Processes and Landforms 32, 853–73.
Coelho Netto AL (1999) Catastrophic landscape evolution in a humid region (SE Brazil): inherit-
ances from tectonic, climatic and land use induced changes. Supplementi di Geografia Fisica e
Dinamica Quaternaria 3 (3), 21–48.
Coelho Netto AL, Fernandes NF, Deus CE (1988) Gullying in the southeastern Brazilian Plateau
Bananal. In Bordas MR, Walling DE (eds) Sediment Budgets – Proceedings of the Porto Alegre
Symposium, 35–42. International Association of Hydrological Sciences Publication No. 174,
Wallingford.
Czudek T, Demek J, Marvan P, Panoš V, Raušer J (1964) Verwitterungs- und Abtragungsformen
des Granits in der Böhmischen Masse [Granite weathering and denudation landforms in the
Bohemian Massif]. Petermanns Geographische Mitteilungen 108, 182–92.
Dubroeucq D, Volkoff B (1998) From Oxisols to Spodosols and Histosols: evolution of the soil
mantles in the Rio Negro basin (Amazonia). Catena 32, 245–80.
Ferrier KL, Kirchner JW (2008) Effects of physical erosion on chemical denudation rates: a numerical
modeling study of soil-mantled hillslopes. Earth and Planetary Science Letters 272, 591–9.
Ford DC, Williams PW (2007) Karst Hydrogeology and Geomorphology. Wiley, Chichester.
Gabet EJ (2007) A theoretical model coupling chemical weathering and physical erosion in
landslide-dominated landscapes. Earth and Planetary Science Letters 64, 259–65.
Gellert JF (1970) Climatomorphology and palaeoclimates of the Central European Tertiary. In Pecsi
M (ed) Problems of Relief Planation, 107–12. Akadémiai Kiado, Budapest.
Godard A (1977) Pays et paysages du granite [Granite Terrains and Landscapes]. Presses Universitaires
de France, Paris.
Gunnell Y (2000) The characterization of steady state in Earth surface systems: findings from the
gradient modelling of an Indian climosequence. Geomorphology 35, 11–20.
Gunnell Y, Bourgeon G (1998) Soils and climatic geomorphology on the Karnataka Plateau,
Peninsular India. Catena 29, 239–62.
Hovius N, Stark CP, Chu HT, Lin JC (2000) Supply and removal of sediment in a landslide-
dominated mountain belt: Central Range, Taiwan. Journal of Geology 108, 73–89.
Hövermann J (1978) Untersuchungen und Darlegungen zum Inselbergproblem in der deutschen
Literatur der 1. Hälfte des 20. Jahrhunderts [Studies and views concerning the problem of
inselbergs in the German literature of the first half of the 20th century]. Zeitschrift für Geomor-
phologie NF, Supplementband 31, 64–78.
Huber NK (1989) The Geologic Story of Yosemite National Park. Yosemite Association, Yosemite.
Jeje LK (1974) Effects of rock composition and structure on landform development: the example
of the Idanre Hills of western Nigeria. Singapore Journal of Tropical Geography 39, 43–53.
Jennings JN (1984) Karst Geomorphology. Blackwell, Oxford.
Johansson M, Migon P, Olvmo M (2001) Development of joint-controlled rock basins in Bohus
granite, SW Sweden. Geomorphology 40, 145–61.
Kesel RH (1973) Inselberg landform elements: definition and synthesis. Revue Géomorphologie
Dynamique 22, 97–108.
Lan HX, Hu RL, Yue ZQ, Lee CF, Wang SJ (2003) Engineering and geological characteristics of
granite weathering profiles in South China. Journal of Asian Earth Sciences 21, 353–64.
Lidmar-Bergström K (1995) Relief and saprolites through time on the Baltic Shield. Geomorphology
16, 33–59.
Mabbutt JA (1952) A study of granite relief from South West Africa. Geological Magazine 89,
87–96.
342 Piotr Migoń

Mäckel R (1985) Dambos and related landforms in Africa – an example for the ecological approach
to tropical geomorphology. Zeitschrift für Geomorphologie NF, Supplementband 52, 1–23.
Meis MRM, Monteiro AMF (1979) Upper Quaternary ‘rampas’: Doce river valley, Southeastern
Brazilian plateau. Zeitschrift für Geomorphologie NF 23, 132–51.
Meis MRM, Moura JRS (1984) Upper Quaternary sedimentation and hillslope evolution – south-
east Brazilian plateau. American Journal of Science 281, 241–54.
Migoń P (1996) Granite landscapes of the Sudetes Mountains – some problems of interpretation:
a review. Proceedings of the Geologists’ Association 107, 25–38.
Migoń P (2006) Granite Landscapes of the World. Oxford University Press, Oxford.
Migoń P, Goudie A (2003) Granite landforms of the Central Namib. Acta Universitatis Carolinae,
Geographica 35, Supplement, 17–38.
Migoń P, Johansson M (2004) Lithological and structural influence on the development of basin-
and-hill landscape within a basement complex in SW Sweden. Zeitschrift für Geomorphologie NF
48, 305–22.
Oberlander T (1972) Morphogenesis of granitic boulder slopes in the Mojave Desert, California.
Journal of Geology 80, 1–20.
Oliva P, Viers J, Dupre B (2003) Chemical weathering in granitic environments. Chemical Geology
202, 225–56.
Oliveira de MAT (1990) Slope geometry and gully erosion development: Bananal, Sao Paulo,
Brazil. Zeitschrift für Geomorphologie NF 34, 423–34.
Pedraza J, Angel Sanz M, Martín A (1989) Formas graniticas de la Pedriza [Granite Landforms of
Pedriza]. Agencia de Medio Ambiente, Madrid.
Prokop P (2007) Degradacja środowiska przyrodniczego południowego skłonu wyżyny Meghalaya, Indie
[Land degradation of the southern slope of the Meghalaya plateau, India]. Prace Geograficzne
No. 210, Institute of Geography and Spatial Sciences, Warsaw.
Pye K, Goudie AS, Thomas DSG (1984) A test of petrological control in the development of
bornhardts and koppies on the Matopos Batholith, Zimbabwe. Earth Surface Processes and
Landforms 9, 455–67.
Raunet M (1985) Les bas-fonds en Afrique et à Madagascar. Géomorphologie – géochimie –
pédologie – hydrologie [Topographic basins in Africa and Madagascar. Geomorphology –
geochemistry – pedology – hydrology]. Zeitschrift für Geomorphologie NF, Supplementband 52,
25–62.
Riebe CS, Kirchner JW, Finkel RC (2004) Erosional and climatic effects on long-term chemical
weathering rates in granitic landscapes spanning diverse climate regimes. Earth and Planetary
Science Letters 224, 547–62.
Selby MJ (1977) Bornhardts of the Namib Desert. Zeitschrift für Geomorphologie NF 21, 1–13.
Selby MJ (1982) Form and origin of some bornhardts of the Namib Desert. Zeitschrift für Geomor-
phologie NF 26, 1–15.
Shroder JF Jr. (1976) Mass movements on the Nyika Plateau, Malawi. Zeitschrift für Geomorphologie
NF 20, 56–77.
Sugden DE (1968) The selectivity of glacial erosion in the Cairngorm mountains. Transactions of the
Institute of British Geographers 45, 79–92.
Thomas MF (1965) Some aspects of the geomorphology of domes and tors in Nigeria. Zeitschrift für
Geomorphologie NF 9, 63–81.
Thomas MF (1966) Some geomorphological implications of deep weathering patterns in crystalline
rocks in Nigeria. Transactions of the Institute of British Geographers 40, 173–93.
Thomas MF (1974) Granite landforms: a review of some recurrent problems of interpretation. In
Brown EH, Waters RS Progress in Geomorphology, 13–37. Institute of British Geographers Special
Publication 7, London.
Thomas MF (1976) Criteria for the recognition of climatically induced variations in granite land-
forms. In Derbyshire E (ed) Geomorphology and Climate, 411–45. Wiley, London.
Thomas MF (1978) The study of inselbergs. Zeitschrift für Geomorphologie NF, Supplementband 31,
1–41.
Granite landscapes distinctive of the humid tropics 343

Thomas MF (1994) Geomorphology in the Tropics. Wiley, Chichester.


Thomas MF (1997) Weathering and landslides in the humid tropics: a geomorphological perspec-
tive. Journal of the Geological Society of China 40, 1–16.
Thomas MF (2006) Lessons from the tropics for a global geomorphology. Singapore Journal of
Tropical Geography 27, 111–27.
Twidale CR (1964) A contribution to the general theory of domed inselbergs. Transactions of the
Institute of British Geographers 34, 91–113.
Twidale CR (1982) Granite Landforms. Elsevier, Amsterdam.
Valentin C, Poesen J, Li Y (2005) Gully erosion: impacts, factors and controls. Catena 63, 132–53.
Vieira GT (2008) Combined numerical and geomorphological reconstruction of the Serra da Estrela
plateau icefield, Portugal. Geomorphology 97, 190–207.
Wahrhaftig C (1965) Stepped topography of the southern Sierra Nevada, California. Geological
Society of America, Bulletin 76, 1165–90.
White AF, Blum AE (1995) Effects of climate on chemical weatering in watersheds. Geochimica et
Cosmochimica Acta 59, 1729–47.
Wilhelmy H (1958) Klimamorphologie der Massengesteine [Climatic Geomorphology of Massive
Rocks]. Westermann, Braunschweig.
Xu J (1996) Benggang erosion: the influencing factors. Catena 27, 249–63.
Young RW, Young AM (1992) Sandstone Landscapes. Springer, Heidelberg.

You might also like