You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/256695730

Petrology of the Namib Sand Sea: Long-distance transport and compositional


variability in the wind-displaced Orange Delta

Article in Earth-Science Reviews · May 2012


DOI: 10.1016/j.earscirev.2012.02.008

CITATIONS READS

161 2,774

6 authors, including:

Eduardo Garzanti Sergio Andò


Università degli Studi di Milano-Bicocca Università degli Studi di Milano-Bicocca
471 PUBLICATIONS 25,735 CITATIONS 193 PUBLICATIONS 8,467 CITATIONS

SEE PROFILE SEE PROFILE

Giovanni Vezzoli Michele Lustrino


Università degli Studi di Milano-Bicocca Sapienza University of Rome
143 PUBLICATIONS 7,256 CITATIONS 160 PUBLICATIONS 5,369 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Maria Boni on 04 April 2018.

The user has requested enhancement of the downloaded file.


Author's personal copy

Earth-Science Reviews 112 (2012) 173–189

Contents lists available at SciVerse ScienceDirect

Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev

Petrology of the Namib Sand Sea: Long-distance transport and compositional


variability in the wind-displaced Orange Delta
Eduardo Garzanti a,⁎, Sergio Andò a, 1, Giovanni Vezzoli a, 1, Michele Lustrino b, 2,
Maria Boni c, 3, Pieter Vermeesch d, 4
a
Laboratorio di Petrografia del Sedimentario, Dipartimento di Scienze Geologiche e Geotecnologie, Università Milano-Bicocca, 20126 Milano, Italy
b
Dipartimento di Scienze della Terra and Istituto di Geologia Ambientale e Geoingegneria (IGAG, CNR), Università di Roma La Sapienza, 00185 Roma, Italy
c
Dipartimento di Scienze della Terra, Università di Napoli, 80134 Napoli, Italy
d
School of Earth Sciences, Birkbeck, University of London, London WC1E 7HX, UK

a r t i c l e i n f o a b s t r a c t

Article history: Sourced as the Nile in distant basaltic rift highlands, the Orange River is the predominant ultimate source of
Received 8 April 2011 sand for the Namib Desert dunes, as proved independently by bulk-petrography, heavy-mineral, pyroxene-
Accepted 20 February 2012 chemistry, and U/Pb zircon-age datasets. Additional local entry points of sand do exist at the edges of the
Available online 3 March 2012
desert, and were quantified by comparison with detrital modes and heavy-mineral suites of hinterland-
river sediments.
Keywords:
Provenance analysis
After long-distance fluvial transport, Orange sand is washed by ocean waves and dragged northwards by vig-
Mechanical abrasion orous longshore currents. Under the incessant action of southerly winds, sand is blown inland and carried far-
Grain roundness ther north to accumulate in the Namib erg, a peculiar wind-dominated sediment sink displaced hundreds of
Aeolian sorting kilometres away from the river mouth. And yet changes in sand mineralogy along the way are minor. After a
Pyroxene chemistry multistep journey of cumulative 3000 km from their source in Lesotho, volcanic rock fragments and pyroxene
U–Pb zircon ages are found in unchanged abundance as far as the northern edge of the desert. Only locally is volcanic detritus
slightly depleted and minor but regular enrichment in quartz and garnet is observed, the sole potential effect
of prolonged transport or recycling of Tertiary aeolianites. Selective comminution of fragile minerals is thus
proved unable to substantially modify sand composition in fluvial, coastal, or aeolian settings. Mechanical
processes have a much greater effect on the morphology of detrital grains, which in Namib dunes appear
commonly shaped into nearly perfect spheres. Aeolian sorting concentrates denser minerals locally in placer
lags, but such effects can be identified and compensated for. This study demonstrates that mechanical break-
down is unable to markedly affect provenance signatures even during long-distance and prolonged multistep
transport in high-energy settings. In arid climates, where chemical processes are negligible, high-resolution
bulk-petrography and heavy-mineral analyses are thus powerful techniques to quantitatively reconstruct
provenance, and to trace sediment sources and dispersal paths over distances up to thousands of kilometres.
© 2012 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
2. The Orange River and the Namib Sand Sea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
2.1. The Orange River . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
2.2. Northward sediment dispersal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
2.3. The Namib Sand Sea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
2.4. Geology of southwestern Namibia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

⁎ Corresponding author. Tel.: + 39 02 64482088; fax: +39 02 64482073.


E-mail addresses: eduardo.garzanti@unimib.it (E. Garzanti), sergio.ando@unimib.it (S. Andò), giovanni.vezzoli@unimib.it (G. Vezzoli), michele.lustrino@uniroma1.it
(M. Lustrino), boni@unina.it (M. Boni), p.vermeesch@ucl.ac.uk (P. Vermeesch).
1
Tel.: + 39 02 64482088; fax: +39 02 64482073.
2
Tel.: + 39 06 49914158; fax: +39 06 4454729.
3
Tel.: + 39 081 253 5068; fax: +39 081 2535070.
4
Tel.: + 44 020 76792418; fax: +44 020 76792867.

0012-8252/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.earscirev.2012.02.008
Author's personal copy

174 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

3. Sand petrology and mineralogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177


3.1. Sampling and analytical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
3.2. Orange River . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
3.3. Hinterland rivers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
3.4. Namib dunes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
4. Provenance of Namib dune sand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
4.1. Contributions from hinterland rivers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
4.2. Recycling from Tertiary aeolianites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
5. Physical processes controlling compositional variability of dune sand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
5.1. Aeolian sorting and placer formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
5.2. Mechanical durability of diverse detrital components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
5.3. Rounding by aeolian abrasion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
6. Sediment budget . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
6.1. Quantitative provenance assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Appendix A. Supplementary material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

“The river”, he reminded her. “Tell me about the river”, and listened Understanding aeolian effects is crucial to correctly interpret
avidly as she went on. “The river gathers them up, from every little desert environments of the past, which is in turn essential for accurate
pocket and crevice along its course. It picks up those that were palaeogeographic and palaeoclimatic reconstructions (Dott, 2003;
flung into the air during the volcanic eruptions at the beginning of Avigad et al., 2005). The peculiarity of the Namib Sand Sea, supplied
the continent's existence. For hundreds of million years it has been from fundamentally one single entry point with a wide spectrum of
carrying the diamonds toward the coast. Those beaches are so rich detrital species including pyroxene and volcanic rock fragments, offers
in diamonds that they are the forbidden area, the Spieregebied.” a unique opportunity to solve such a thorny petrological problem,
Wilbur Smith, Power of the sword, p.78 provided we succeed in quantifying subsidiary detrital sources
(hinterland rivers, recycled Tertiary sediments, deflation areas along
the coast; Lancaster and Ollier, 1983; Besler, 1984) and local effects
1. Introduction of wind-induced sorting.
The aims of the present study are to illustrate in detail the composi-
The Orange has been a powerful river possibly since the Jurassic tional variability of Namib dune sands through high-resolution bulk-
(Cox, 1989). Or at least it was, until man tamed it with dams to ex- petrography and heavy-mineral analysis, to pin-point all sediment
ploit its waters for agricultural and industrial purposes. And extreme- sources and quantify their relative contributions to various parts of the
ly energetic has long been the microtidal South Atlantic shore where erg, to evaluate wind-induced concentrations of heavy minerals, and to
it debouches, subjected to vigorous waves and northward longshore assess morphological and compositional changes caused by mechanical
drift fuelled by persistent southerly winds (Bluck et al., 2007). As abrasion and selective breakdown of detrital grains with variable
well as an open-air diamond mine so rich that it is forbidden land durability during aeolian transport. Quantitative provenance analysis
(Jacob et al., 2006), the Orange Delta is thus widely recognised as a represents an effective way to identify long-term transport paths and
wave-dominated end-member (Spaggiari et al., 2006). But there is sediment modifications in modern deserts, and represents a key require-
more. Incessantly dragged northwards by longshore currents and ment for reconstructing palaeowind patterns and palaeoclimate changes
eventually blown inland under the long-lived wind regime, Orange during the geological evolution of such dynamic geomorphic systems.
Sand has been accumulating since the Miocene at least in the Namib
Sand Sea, a sediment sink detached from the river mouth and dis- 2. The Orange River and the Namib Sand Sea
placed on land several hundreds of kilometres away (Fig. 1; Rogers,
1977; Vermeesch et al., 2010). 2.1. The Orange River
In such a geological setting, changes in sediment composition and
grain morphology caused by multistep transport can be monitored The Orange is one of the largest African rivers, with a drainage
over a cumulative distance summing up to 3000 km. Whereas abra- area of ~970,000 km 2 (Fig. 2). Sourced not far from the Indian
sion and mechanical breakdown have long been established to be Ocean in the Drakensberg mountains of Lesotho (maximum elevation
scarcely effective during long-distance fluvial transport by both stud- 3482 m a.s.l.), it flows westwards for ~ 2200 km across an increasingly
ies of natural river systems and laboratory experiments (Russell and arid interior plateau towards the Atlantic Ocean (Moore et al., 2009).
Taylor, 1937; Kuenen, 1959), their incidence during wind transport The catchment receives primarily summer rainfall (October to April),
has been debated for a full century, and still remains controversial varying from 1200 mm/yr in relatively cold Lesotho highlands (mean
(Twenhofel, 1945; Goudie and Watson, 1981). Laboratory experi- temperatures − 5 °C in winter and 16 °C in summer) to 40 mm/yr
ments have shown that aeolian abrasion is 100–1000 times more near the mouth (Swanevelder, 1981). Rainfall and snowmelt in
effective than fluvial abrasion (Kuenen, 1960), and the generation of Lesotho, occupying only 3% of the catchment, contributes 47% of
quartz-rich sand through selective breakdown of feldspar by aeolian total water flow (Makhoalibe, 1999).
impacts has received theoretical support (Dutta et al., 1993). Studies This ancient drainage system, originated in the Early Jurassic as a
in natural environments, however, have produced ambiguous results result of ~2 km surface uplift associated with emplacement of
(Johnsson, 1993), because detrital sources are generally multiple and Drakensberg lavas during rifting from Antarctica, cuts antecedently
undetermined, and dune sand commonly consists of quartz whose in its lowermost tract across the regional uplift formed during
rounded shape may have resulted from aeolian abrasion as well as Lower Cretaceous rifting of the South Atlantic (Cox, 1989; Moore
from recycling of rounded grains from older sandstones (Garzanti and Blenkinsop, 2002). The oldest basement rocks are widely exposed
et al., 2003; Muhs, 2004; Mehring and McBride, 2007). north of the Vaal River (Neoarchean metavolcanic Ventersdorp
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 175

climates (Beaufort Group). A retroarc foreland-basin setting is indi-


cated by Ecca and lower Beaufort quartzofeldspatholithic volcaniclas-
tic sandstones, fed from a partly dissected continental arc located in
the south, and overlain by feldspathoquartzolithic to feldspatholitho-
quartzose sandstones of the upper Beaufort Group (Johnson, 1991).
Above a major unconformity, lithoquartzose braidplain sandstones
were deposited in the Carnian–Norian (Molteno Fm.), followed by
red floodplain mudstones (Elliot Fm.), and by Lower Jurassic aeolian
feldspathoquartzose sandstones (Clarens Fm.; Johnson et al., 1996;
Catuneanu et al., 2005). The succession is capped by up to 1.8 km
thick Drakensberg continental flood basalts (183 ± 2 Ma; Jourdan et
al., 2005). A vast network of dolerite dykes and sills suggests that
tholeiitic lavas originally covered an area of ~2.5 10 6 km 2. Also wide-
spread are pipe-like bodies, including diamond-bearing kimberlites,
intruded mainly during Jurassic to Cretaceous break-up of Gondwana
(Moore et al., 2008).
Most of the sediment flux is generated by high orographic rainfall
Fig. 1. The Namib Sand Sea: the displaced sediment sink of Orange River sediments.
and topographic relief in Lesotho highlands. Suspended load is mainly
supplied by erodible upper Karoo mudrocks, as reflected by the
Supergroup and Palaeoproterozoic metasedimentary Transvaal chemical composition of river muds (Bremner et al., 1990 p. 251;
Supergroup; de Villiers et al., 2000). The lower Orange downstream Compton and Maake, 2007 p. 344–345), whereas Drakensberg basalts
of the Vaal confluence cuts across Mesoproterozoic metasedimentary supply little silt (Haskins and Bell, 1995). Suspended-sediment
and metavolcanic rocks of the Namaqua Belt (Becker et al., 2006), yields can thus be locally as high as 2000 tons/km2·yr in the upper
whereas the upper Orange and its two main branches (Caledon Caledon catchment largely draining the Elliot Formation, but as
and Senqu Rivers) drain the up to 10 km thick Karoo Supergroup low as 10–70 tons/km 2·yr in the upper Senqu catchment draining
exclusively. basaltic rocks (Makhoalibe, 1984). Various tributaries draining
The Karoo succession includes Upper Carboniferous to lowermost Karoo sediments owe their name to their turbid coloured waters
Permian feldspathoquartzose glacial sediments (Dwyka Group), over- (e.g., in Afrikaans Vaal = pale, grey; Modder = mud), but not the
lain by shale, turbiditic to fluviodeltaic sandstone and coal deposited Orange itself, named in honour of William of Orange.
at wetter lower latitudes (Ecca Group), and by uppermost Permian After construction of major dams on the Vaal (Vaal Dam in 1938;
to Anisian floodplain deposits accumulated in warmer and drier Bloemhof Dam in 1970) and upper Orange (Gariep Dam in 1972;

Fig. 2. Geological sketch map of Namibia, South Africa and Lesotho (compiled after Schlüter, 2006, and other sources cited in text). Inset shows the location of studied aeolian dune
and fluvial samples around the Namib erg. Major dams are also indicated: B = Bloemhof, G = Gariep, H = Hardap, K = Katse, Va = Vaal, Vk = Vanderkloof, W = Welbedacht.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
Author's personal copy

176 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

Vanderkloof Dam in 1977), total Orange suspended-load flux at the changes from northwest to north. Offshore transport of Orange
mouth has progressively dropped from ~90 · 10 6 tons/yr to the River sand at subtidal to intertidal depths continues beyond Walvis
present b17 · 10 6 tons/yr (Rooseboom and Harmse, 1979). Annual Bay (Bluck et al., 2007). Dune sand in the northwestern corner of
water discharge dropped from ~ 11.5 km 3 to ~5.7 km 3, but peaks of the Sand Sea may thus have been blown inland directly from the
26 km 3 are still reached in flood years (maximum discharge rate Atlantic coast. Sands further to the east have travelled overland for
8072 m 3/s, 2nd March 1988; Bremner et al., 1990 p. 262). Peak longer distances. Wind energy and sand-transport rates decrease
suspended-sediment concentrations were drastically reduced from landwards (Lancaster, 1985).
extreme values in early times (up to 63 g/l for the Caledon and
40 g/l for the Orange, second only to the Yellow River in China) to 2.3. The Namib Sand Sea
7.4 g/l in the 1988 flood, when sediment load remained significant
(81 · 10 6 tons in March to May 1988) but dominantly derived from The Namib Desert, part of the passive margin originated during
bank erosion and river-bed scour downstream of major dams Early Cretaceous rifting of the South Atlantic (Goudie and Eckardt,
(Bremner et al., 1990 p. 247, 285). Suspended load consisted of 1999), stretches for nearly 600 km along the SW African coast
~ 20% sand, 60% silt and 20% clay in pre-1970 times, and of ~15% between Lüderitz (~ 27°S) and the Kuiseb River (23°S), and for
sand, 30% silt and 55% clay (25 ± 13% montmorillonite, 63 ± 13% illite, 100–150 km inland to the base of the Great Escarpment at the
12 ± 3% kaolinite) during the 1988 flood (Bremner et al., 1990 p. 273). 1000 m contour. Southerly to southwesterly trade winds dominate
Storage loss due to rapid sedimentation in the Welbedacht and the atmospheric circulation in all but the northeasternmost parts of
Gariep reservoirs (Jordaan and Clark, 1988; Sawadogo, 2008) points the desert, although northeasterly winds are also common in winter
to total sediment yields up to 1000–1100 tons/km 2·yr in both (Lancaster, 1985). Climate is hyperarid, with precipitation ranging
Caledon and Senqu catchments, corresponding to erosion rates up from b50 mm/yr and mostly in the form of fog near the coast, to
to 0.4 mm/yr. Total annual load of the upper Orange could thus ~100 mm/yr at the foot of the Great Escarpment (Lancaster, 2002).
reach 60 · 10 6 tons. If 49 ± 7 10 6 tons are accounted for by suspended Ephemeral rivers draining the Escarpment and the semiarid edge of
load (Rooseboom and Harmse, 1979), then bedload represents ≥10% the southern African plateau (rainfall 200–450 mm/yr) penetrate
of total load in Lesotho, significantly more than the ~5% estimated for ≤50 km into the erg, to empty their sediment-laden waters in flat
the whole catchment (Bremner et al., 1990 p. 260). interdune playas named “vleis”. Floods may last a few weeks, after
Sediment yields of several hundred tons/km 2·yr are reported which the river beds are dry for the rest of the year (Jacobson et al.,
even from lower-relief subcatchments draining Karoo sedimentary 1995). Only exceptionally high floods of the Kuiseb River reach the
bedrock, where development of badlands and gullies was enhanced sea at the northern edge of the Namib (16 times since 1837; Morin
by extensive grazing activities in the last century (Rooseboom and et al., 2009).
Lotriet, 1992). Sediment yields are estimated to be significantly The sand sea (~ 34,000 km 2) is dominated by large linear dunes,
lower for the low-relief Vaal catchment (~150 tons/km2·yr in with areas of star-shaped dunes on its eastern margin and a belt of
1932–1939), and very minor for the arid lower Orange catchment simple and compound transverse and barchanoid dunes along the
downstream of the Vaal confluence (Rooseboom and Harmse, 1979 coast. Optical dating in the northern Namib indicates that linear
p. 463–462). dunes are younger (~5.7 ka; Bristow et al., 2007) than big star
Cosmogenic 10Be and 26Al measurements in quartz carried by the dunes (43 ± 10 ka; C. Bristow pers. comm. 2009). Considering that
Orange River point to catchment-wide erosion rates of only 0.004 ± the present estimated rate of sand input to the Namib Sand Sea is
0.001 mm/yr averaged over millennial time scales (Vermeesch et al., 400,000 m 3/a (Lancaster, 1989), ≥1 Ma would be required to form
2010). Assuming an average surface rock density of 2.7 g/cm 3, this the modern erg (estimated at 375 to 1020 km 3). This is consistent
corresponds to sediment fluxes of only 11 ± 2 · 10 6 tons/yr and sedi- with an average residence time ≥1 Ma for sand grains in the coastal
ment yields of 11 ± 2 tons/km 2·yr for the entire Orange catchment, Namib, as determined by cosmogenic-nuclide measurements
much lower than recorded through the last century. As for the Nile (Vermeesch et al., 2010).
(Garzanti et al., 2006), this may largely reflect human-induced
accelerated erosion caused by intensive land use. 2.4. Geology of southwestern Namibia

2.2. Northward sediment dispersal Namibia is crossed by the Mesoproterozoic Namaqua (“Kibaran”)
and Neoproterozoic Damara (“Pan-African”) orogenic belts, welding
The Orange River mouth is an extreme example of a wave- the Archean Kalahari and Congo Cratons (Fig. 2; Jacobs et al., 2008;
dominated delta. Sand and gravel delivered to the South Atlantic Miller, 2008). Exposed along the lower Orange to the southern
coast are carried northwards by a powerful longshore drift under pre- margin of the Namib erg (Aus–Lüderitz area) is the Namaqua Meta-
vailing southerly winds. High-energy conditions persist throughout morphic Complex, including medium to high-grade metasediments
most of the year along the coast and adjacent shelf (wave heights interpreted as originally deposited on a passive margin (Becker et
90% in the 0.75–3.25 m range, average 1.5 m for summer and al., 2006). Exposed east of the Namib erg is the Mesoproterozoic
1.75 m for winter). Tidal range is 1.8 m along this microtidal coast, Sinclair Group, largely overprinted by Damaran structures in the
where the Orange River is the sole significant source of sediment Rehoboth area. These low-grade volcanic and subordinate sedimenta-
(Rogers and Rau, 2006). ry rocks rest disconformably on medium-grade metasedimentary and
Whereas mud moves largely north and west out to the continental metavolcanic rocks intruded by 1.37 Ga tonalite, and may have
shelf edge, where it forms extensive mud sheets (Spaggiari et al., represented the active margin of the Kalahari Craton facing an east-
2006), much of the sand is retained within the breaker zone and directed oceanic subduction zone. The widespread intrusion of
moves alongshore in a ≤3 km wide belt. Gravel spits and barrier bea- granitoid batholiths was followed by outpouring of continental
ches predominate in proximal settings, passing northwards to linear tholeiites at ~1.1 Ga (Becker et al., 2006).
beaches and finally to pocket beaches. Gravel beaches, found as far The Damara Belt exposed at the northern margin of the Namib erg
as 300 km north of the Orange mouth, include the richest diamond includes 2.0 to 1.2 Ga basement gneisses overlain by Neoproterozoic
placers known (Corbett and Burrell, 2001; Jacob et al., 2006). metasediments. Peak metamorphic conditions, reaching granulite-
At a number of localities sand is blown off the beaches onto the facies and partial melting north of Walvis Bay, were attained at
land, largely bypasses the Sperrgebiet deflation area (Corbett, 1993), 525–504 Ma. Widespread are 560–470 Ma old granites and subordi-
and accumulates north of Lüderitz, where coastline orientation nate diorites/granodiorites (Miller, 1983; Jung and Mezger, 2003).
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 177

Precambrian fold belts are unconformably overlain by the 3 km-thick (e.g., sedimentaclastic, volcaniclastic, plutoniclastic, metamorphiclas-
shallow-marine to fluvial Nama Group, a Neoproterozoic to Cambrian tic). Metamorphic grains were classified according to protolith com-
succession including quartzose to micaceous sandstones, mudrocks, position and metamorphic rank. Average rank for each sample is
and interbedded limestones deposited in the foreland basin of the expressed by the Metamorphic Index (MI; Garzanti and Vezzoli,
Damara orogen (Saylor et al., 1995; Geyer, 2005; Blanco et al., 2003), which varies from 0 in detritus from unmetamorphosed cover
2009). Deformation increases towards the Damara belt, and very- rocks to 500 in detritus from granulite-facies basement rocks.
low metamorphic grade is reached in slates and phyllites of the Heavy Mineral Concentration indices express the volume percent-
Naukluft mountains to the north (Ahrendt et al., 1978). Neoprotero- age of transparent (tHMC) and total (i.e., including opaque and turbid
zoic successions locally hosting volcanic rocks are overprinted by up grains; HMC) heavy minerals in each sample; the SRD (Source Rock
to lower amphibolite-facies metamorphism in the Gariep Belt to the Density) index is the weighted average density of extrabasinal terrige-
south (Frimmel and Frank, 1998). nous grains (Garzanti and Andò, 2007a). The Hornblende Colour
Younger rocks include Upper Palaeozoic Karoo terrigenous Index (HCI) estimates the average metamorphic grade of amphibolite-
sediments, the Lower Jurassic Kalkrand basalts, and Upper Cretaceous to granulite-facies source rocks (Garzanti and Andò, 2007b). Transpar-
kimberlite pipes (Smith et al., 1993; Stollhofen et al., 2000; Davies ent heavy-mineral suites are described as “poor” (0.5 ≤ tHMC b 1),
et al., 2001). “moderately poor” (1≤ tHMC b 2), “moderately rich” (2≤ tHMC b 5),
“rich” (5≤ tHMC b 10), “very-rich” (10 ≤ tHMC b 20), or “extremely
3. Sand petrology and mineralogy rich” (20 ≤ tHMC b 50). Minerals representing ≥10% of the transparent
heavy-mineral suite are listed in order of abundance; rarer but signifi-
3.1. Sampling and analytical methods cant species are also mentioned. The roundness of counted detrital min-
erals was evaluated by visual comparison with home-made standard
Twelve, fine to medium-grained dune-sand samples of 1 kg each images (nomenclature after Powers, 1953).
were collected in summer 2007 along three E–W transects, five Major and trace element concentrations of the sand fraction
along the northern edge of the Namib Desert south of Kuiseb River (63–2000 μm) of Orange sample NAM 14 were measured by ICP-
canyon (NAM 1–5), three in the Sossusvlei area (NAM 6–8), and AES and ICP-MS at the ACME Laboratories, Vancouver (group 4A–
four along the southern edge of the desert between Aus and Lüderitz 4B; for detailed information see http://acmelab.com/). Weathering
(NAM 9–12; Fig. 2). Because dark-red to black patches where heavy indices were calculated using molecular proportions and correcting
minerals (pyroxene, garnet, magnetite) are concentrated by wind for CaO in phosphates and carbonates; extensive alteration is indicated
turbulence commonly occur on dune flanks, all samples were collect- by high CIA (Chemical Index of Alteration; Nesbitt and Young, 1982)
ed on the crest of the largest dunes. Four Orange River sands, sampled and low WIP (Weathering Index; Parker, 1970).
from the mouth (NAM 13), to ~ 80 (S4037) and ~ 100 km upstream
(NAM 14), include one Quaternary terrace (S4038). In order to assess 3.2. Orange River
the petrographic and mineralogical signatures of all potential sources
of Namib dunes, sand samples were also collected from eleven hinter- Orange sand samples range from very-fine to medium-grained,
land rivers across SW Namibia in July 2009. from moderately-poor to moderately-well sorted, and from feld-
Quartered aliquots of bulk-sand samples were impregnated with spatholithoquartzose to lithofeldspathoquartzose (Fig. 3A). Plagioclase
Araldite, cut into standard thin sections, stained with alizarine red prevails over K-feldspar. Rock fragments include equally abundant
to distinguish dolomite and calcite, and analysed by counting 400 volcanic/subvolcanic (largely mafic lathwork types) and sedimentary
points under the petrographic microscope (Gazzi–Dickinson method; grains (shale, quartzose to feldspathoquartzose siltstone/sandstone,
Ingersoll et al., 1984). From the 32–500 μm fraction of each sample, limestone, dolostone, hybrid carbonate), along with common granit-
treated with sodium ditionite–citrate–bicarbonate to remove Fe- oid, metasedimentary (slate, metasandstone, quartz–sericite, quartz–
oxide coatings, heavy minerals were separated by centrifuging in so- mica, schist, marble) and subordinate metaigneous (porphyroid,
dium metatungstate (density ~ 2.90 g/cm 3), and recovered by partial chloritoschist, amphibolite) grains (Fig. 3B). Biotite is abundant in
freezing with liquid nitrogen; 200 to 250 transparent heavy minerals finer-grained samples. Heavy-mineral assemblages are rich and
were counted in grain mounts (“area method”; Mange and Maurer, clinopyroxene-dominated, with subordinate opaque Fe–Ti–Cr oxides,
1992). The reasons for choosing a 4ϕ-wide size-window for heavy- epidote, amphibole, and garnet (Fig. 4). Olivine was never recorded.
mineral analysis are discussed in Appendix A1. Compositional variability is particularly marked for the pyroxene/
In order to determine their provenance, WDS analyses of 134 epidote ratio, which varies from 2 to 32 in modern sands and reaches
pyroxene grains from five selected samples of fluvial (Orange and 37 in the Quaternary sand (Table 1). Hardly ascribed to grain-size ef-
Fish Rivers) and Namib dune sands were carried out with a Cameca fects and hydraulic sorting because pyroxene and epidote do not have
SX-50 microprobe at the CNR-IGAG Laboratories (method illustrated markedly dissimilar density and shape, such strong variability is pos-
in Lustrino et al., 2005). Data on pyroxene chemistry are given in sibly enhanced by anthropogenic effects (irregular sediment flux after
full and discussed in detail in Appendix A2. extensive dam regulation of the upper Orange and Vaal, and exten-
In order to assess the large-scale pattern of sediment transport sive mining along the lower Orange). If this is true, the pyroxene-
and sand residence time in the Namib Sand Sea, the U–Pb ages of rich Quaternary terrace may reflect the original sediment composi-
~ 100 zircon grains were determined for Orange River sample NAM tion more faithfully than modern sand, locally epidote-rich because
13 and for each of the twelve Namib dune samples. Cosmogenic of reduced supply from the headwaters and consequently greater
10
Be and 26Al measurements on detrital quartz were carried out relative contribution from greenschist-facies rocks exposed along
along a longitudinal transect from the mouth of the Orange River to the lowermost course (Frimmel and Frank, 1998).
the northern edge of the Namib erg. Details on the adopted methods, Although the Orange catchment is huge and characterised by a
followed strategy, and obtained results are discussed in Appendix A3 quite varied geology, the chemical composition of detrital pyroxenes
and illustrated in Vermeesch et al. (2010). is remarkably homogeneous, suggesting a similar origin for most ana-
Sands were classified according to their main components lysed grains. Low-Ti augite is dominant, commonly relatively rich in
exceeding 10%QFL (QFL = quartz + feldspars + lithic fragments), listed Cr2O3 or Al2O3, and with Mg# 73 ± 9, suggesting equilibrium with
in order of abundance (e.g., in a quartzolithic sand L > Q > 10%QFL > F, basaltic to basaltic andesite melts. Pigeonite is rare. The composition
in a lithofeldspathoquartzose sand Q > F > L > 10%QFL); an adjective of both augites and pigeonites resembles that of clinopyroxenes
reflecting the most common rock-fragment type may be added commonly found in continental flood basalts, and is virtually
Author's personal copy

178 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

Fig. 3. Comparison between fluvial Orange and aeolian Namib sands. (A, B, C) Orange sand contains volcanic lithics (Lv) and angular clinopyroxene (p) from Drakensberg basalts,
and siltstone grains (Lt) from Karoo sedimentary rocks. (D, E, F) Namib dunes have similar composition. Heavy minerals (o = opaque Fe–Ti–Cr oxide; g = garnet) are locally con-
centrated in placer lags (black arrows point at Fe-oxide coatings) and markedly rounded by wind action (white arrows). All photos but E with crossed polars; blue bar = 250 μm.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

indistinguishable from available analyses of pyroxenes in Karoo weathering and limited depletion in alkali and alkaline-earth metals
basalts (Fig. 5; Sweeney et al., 1994; Melluso et al., 2008; Galerne in semiarid subtropical climate (Fig. 6).
et al., 2010). The distribution of U/Pb zircon ages in Orange River sample
Relative to the estimated average concentration of chemical ele- NAM 13 has a log-normal appearance mostly ranging between 2000
ments in the upper continental crust (UCC; Taylor and McLennan, and 500 Ma, with a peak at 1000 Ma and a few grains at 300 Ma
1995; McLennan, 2001; Hu and Gao, 2008), Orange sand contains (Fig. 7).
half or less Na, Cs, Be, Sr, U, Nb, Ta, Mo, Ni, Sn, Pb, Sb, Bi, and is richer
only in Si and particularly Cr. The REE pattern is similar to the UCC but 3.3. Hinterland rivers
with more gently sloping LREE (LaN/SmN 2.9, GdN/LuN 1.4) and less
negative Eu anomaly (Eu/Eu⁎ 0.83 versus Eu/Eu⁎UCC 0.65), reflecting River sands of SW Namibia are derived from Proterozoic to Cam-
significant basaltic sources. Weathering indices (CIA 46, WIP 50), brian sedimentary/metasedimentary, volcanic/metavolcanic, plutonic
lower than in upper Orange suspended load (CIA 60 ± 2, WIP and metamorphic rocks in various proportions (for geological details
38 ± 2; data after Compton and Maake, 2007), indicate minor of catchment areas see Becker et al., 2006). The Swakop River, largely
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 179

Fig. 4. Petrography and mineralogy of coastal Namib dunes document dominant provenance from the Orange River. Orange sand contains more sedimentary and metamorphic
lithics, which are selectively destroyed during high-energy longshore transport. Instead, neither volcanic lithics (which maintain their relative proportion) nor pyroxene are deplet-
ed in coastal Namib dunes. Progressive eastward enrichment in quartz, K-feldspar, epidote and amphibole relative to volcanic lithics and pyroxene documents hinterland contri-
butions by the Kuiseb and Koichab Rivers in the NE and SE Namib, respectively. Conversely, Kuiseb and Tsauchab sands are enriched downstream in volcanic lithics and pyroxene
due to aeolian mixing with dune sand. Note compositional similarity between Orange and Nile sands (data after Garzanti et al., 2006), both largely derived from rift-related basaltic
highlands. Q = quartz; F = feldspar; L = aphanite lithics (Lm = metamorphic; Lv = volcanic; Ls = sedimentary).

draining Damaran granites and amphibolite-facies metasediments Along its course at the northern boundary of the Namib erg, Kuiseb
of the Nosib and Swakop Groups (Jung and Mezger, 2003), carries sand is slightly enriched in clinopyroxene, revealing mixing with dune
feldspathoquartzose sand (MI 440); the moderately-rich amphibole– sand.
garnet–clinopyroxene–staurolite heavy-mineral suite includes tourma- The Tsondab River, largely draining sedimentary and metasedimen-
line, apatite, titanite, and minor kyanite and fibrolitic sillimanite tary rocks of the Nama Group and Naukluft Nappe Complex, carries
(Table 1). The Kuiseb River, largely draining lower-amphibolite-facies feldspathoquartzolithic sand with limestone, dolostone, hybrid carbon-
quartzites to micaceous schists (Khomas Subgroup; Miller, 1983; de ate, sandstone, metasandstone, shale, slate, phyllite and schist grains
Kock, 2001), carries quartzose to feldspathoquartzose metamorphiclas- (MI 157; Fig. 8B), and a moderately-rich epidote–amphibole–clinopyr-
tic sands with abundant biotite (MI 396; Fig. 8A); the moderately-rich oxene heavy-mineral suite. The Zebra/Tsauchab River, largely
amphibole–epidote–garnet suite includes apatite, tourmaline, clinopyr- draining Nama Group sedimentary rocks, carries quartzolithic sedi-
oxene, and rare zircon, rutile, titanite, staurolite, kyanite and monazite. mentaclastic sand dominated by shale/slate, limestone, quartzose
The Gaub River, a Kuiseb tributary largely draining lower-amphibolite- to feldspathoquartzose sandstone and hybrid carbonate grains
facies rocks of the Rehoboth Terrane, carries feldspathoquartzose (MI 30 ± 7; Fig. 8C); the poor clinopyroxene–hornblende–epidote
metamorphiclastic sand with biotite, muscovite, a few amphibolite suite includes garnet, titanite, Ti-oxides, minor apatite, tourmaline,
and calcsilicate grains (MI 385), and a rich epidote–amphibole suite. and trace staurolite. After entering the Namib erg towards Sossusvlei,
Author's personal copy

180 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

Table 1
Bulk-petrography and heavy-mineral modes for Orange River, Namib dune, and hinterland-river sands (mean in bold, standard deviation in italics). No. = number of samples; Q =
quartz (RDN = rounded; sRDN = subrounded); F = feldspar; L = aphanite lithics (Lv = volcanic; Lc = carbonate; Lt = siltstone/sandstone; Lm = metamorphic). MI = Metamor-
phic Index. HMC = Heavy Mineral Concentration. Op = opaque Fe–Ti–Cr oxides; Px = pyroxene; Amp = amphibole; Ep = epidote; Grt = garnet; St = staurolite; Zrn = zircon; Tur
= tourmaline; TiOx = Ti oxides; Ttn = titanite; Ap = apatite; & = others (andalusite, sillimanite, kyanite, monazite). HCI = hornblende-colour index.

No. Q F Lv Lc Lt Lm MI QRDN QsRDN HMC Op Px Amp Ep Grt St Zrn Tur TiOx Ttn Ap & HCI

R. Orange 4 57 21 8 2 7 4 100.0 56 3% 42% 6.6 12 62 7 12 3 0 0 1 3 1 1 0 100.0 6


9 7 5 3 6 2 10 3% 11% 1.9 6 22 2 12 2 0 0 1 3 1 1 0 4
SW Namib 2 66 23 9 0 1 1 100.0 32 7% 41% 10.1 9 76 4 3 6 0 1 0 1 0 0 0 100.0 11
6 5 0 0 1 0 26 6% 2% 2.0 4 12 1 1 7 0 1 0 0 0 0 0 2
NW Namib 3 68 20 10 0 1 1 100.0 38 10% 37% 7.2 11 72 2 4 9 1 0 0 0 1 0 0 100.0 7
2 1 3 0 0 0 17 2% 10% 1.6 2 3 2 0 1 0 0 0 0 1 0 0
SE Namib 2 70 26 3 0 1 0 100.0 31 8% 53% 7.2 34 31 9 6 15 1 2 0 0 2 0 0 100.0 16
6 4 0 0 0 1 11 1% 5% 2.4 23 25 6 0 6 0 1 0 0 0 0 0 11
CE Namib 3 73 19 6 0 1 1 100.0 52 10% 55% 10.4 27 50 6 5 8 0 1 0 1 0 0 0 100.0 6
2 4 0 0 1 1 56 2% 17% 12.3 18 19 5 1 4 0 0 0 1 0 0 0 2
NE Namib 2 79 16 3 0 1 1 100.0 48 7% 42% 1.3 28 13 9 26 17 1 0 1 2 3 0 1 100.0 6
1 2 1 0 0 0 14 4% 9% 0.6 12 13 2 3 6 1 0 0 0 1 0 0 2
Swakop 1 66 33 0 1 0 1 100.0 338 0% 26% 3.0 31 9 21 4 13 6 0 4 3 3 4 2 100.0 3
Kuiseb 2 85 9 0 2 0 5 100.0 396 0% 45% 3.2 39 3 18 15 7 1 2 5 3 2 5 0 100.0 1
Gaub 1 73 20 0 1 0 6 100.0 385 0% 40% 7.8 12 2 38 38 1 0 0 1 2 4 3 0 100.0 0
Tsondab 1 37 24 0 17 5 16 100.0 157 0% 64% 4.4 44 6 13 23 2 0 4 1 2 1 2 0 100.0 6
Tsauchab 2 14 2 1 28 44 12 100.0 27 0% 0% 1.0 15 30 25 18 4 0 0 1 2 4 1 0 100.0 2
Tsaris 1 44 25 4 15 10 2 100.0 36 0% 40% 1.3 37 14 19 18 1 0 2 1 2 3 2 0 100.0 6
Lovedale 1 48 49 1 1 0 1 100.0 60 1% 84% 2.0 58 4 8 20 2 0 5 0 1 2 0 0 100.0 6
Koichab 1 43 41 2 10 4 1 100.0 103 0% 31% 3.6 52 6 23 7 6 0 3 0 1 1 1 0 100.0 25
Konkiep 1 44 28 18 0 4 6 100.0 64 1% 69% 4.3 18 7 16 56 1 0 1 0 0 1 1 1 100.0 15
Fish 1 60 14 5 1 18 3 100.0 16 1% 80% 2.7 12 73 3 6 3 0 1 1 1 0 0 0 100.0 n.d.

Tsauchab sand is progressively enriched in subrounded to rounded few amphibolite and felsic volcanic grains (MI 103; Fig. 8E); the
quartz, feldspars, volcanic lithic fragments and clinopyroxene, revealing moderately-rich amphibole–epidote–clinopyroxene–garnet suite in-
mixing with dune sand (Fig. 4). Tsaris sand is feldspatholithoquartzose cludes common green–brown hornblende (HCI 25), reflecting erosion
sedimentaclastic (MI 36) with a moderately-poor amphibole–epidote– of amphibolite-facies to granulite-facies rocks exposed in the Aus area.
clinopyroxene suite, reflecting mixed provenance from Nama sedimen- The Fish River, a tributary of the Orange draining the Nama Group
tary and Sinclair magmatic rocks. and the overlying Karoo sedimentary and basaltic rocks (Kalkrand
The Konkiep River, sourced in the very low-grade volcanic and basalts of central Namibia), carries feldspatholithoquartzose sand with
subordinately sedimentary rocks of the Sinclair Group with associat- abundant quartzose or feldspathoquartzose sandstone and shale/slate
ed plutonic intrusions, carries feldspatholithoquartzose sand with grains, common volcanic lithic fragments (MI 16; Fig. 8F), and a
volcanic/metavolcanic, subvolcanic, plutonic, sandstone/metasandstone, moderately-rich clinopyroxene-dominated suite. Pyroxene grains dis-
and shale/slate grains (MI 64); the moderately-rich epidote- play quite similar chemical composition as those carried by the Orange
dominated suite includes amphibole and minor clinopyroxene. River (Fig. 5), indicating provenance dominantly from the Kalkrand
Lovedale sand, entirely derived from such granitoids, is quartzofelds- continental flood basalts widely exposed in the upper course.
pathic, with a moderately-poor epidote–amphibole–clinopyroxene
suite (Fig. 8D). 3.4. Namib dunes
Similarly plutoniclastic is the lithofeldspathoquartzose Koichab sand
derived from the Namaqua Metamorphic Complex, which contains sub- The analysed coastal Namib dune sands are mainly lower
ordinate sedimentary (limestone, hybrid limestone, sandstone) and a medium-grained, well-sorted, symmetrical and mesokurtic (1.9 ±

Fig. 5. Clinopyroxene grains in Orange sand, Fish sand, and Namib dunes (N = NAM 3, NW = NAM 1, SW = NAM 12). Pyroxene grains display the same range of compositions as in
various types of Karoo lavas (data after Sweeney et al., 1994; Melluso et al., 2008; Galerne et al., 2010), indicating common provenance, dominantly from Karoo continental flood
basalts. Pyroxene quadrilateral diagram after Morimoto et al. (1988).
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 181

Fig. 6. Sand chemistry in big African rivers (elements in UCC-normalised multielement diagram arranged following the periodic table group by group). Composition of Orange sand
~100 km upstream of the mouth (NAM 14) compares closely to that of Nile sand upstream Lake Nasser (Padoan et al., 2011). For both passive-margin river systems, formed long
ago by rift-related domal uplift, Cr enrichment and Eu anomaly less negative than UCC reflect focused erosion of distant basaltic highlands. The Cr peak is controlled chiefly by au-
gitic pyroxene, and subordinately by basaltic rock fragments and Fe–Ti–Cr oxides. Only moderate depletion in alkali and alkaline-earth metals suggests limited weathering in semi-
arid tropical climate. In contrast, weathered sand of the equatorial Congo River is turned into a quartz residue, as shown by strong depletion in all elements but Si and by pattern
similar to detrital quartz. In Orange sand, depletion in REE and transition metals, removed from primary minerals and involved in the formation of oxyhydroxides (Quinn et al.,
2006), is largely ascribed to fractionation of tiny phyllosilicates in suspended load. Soluble Mo is strongly depleted. All three analysed sand samples have normal grain density
(SRD 2.71 ± 0.02) and are not markedly affected by selective-entrainment effects. Data for Upper Orange suspended load after Compton and Maake (2007). Quartz separates
obtained from 4 central African and Arabian sand samples.

0.2ϕ, σ 0.46 ± 0.11ϕ, Sk 0.0 ± 0.1, Ku 2.6 ± 0.0). As observed by Dune sands in the coastal Namib are invariably lithofeldspatho-
Lancaster and Ollier (1983 p. 72; Lancaster, 1989), grain size tends quartzose volcaniclastic, with homogeneous composition from
to decrease, and sorting to increase in eastern Namib dunes (2.1 ± Lüderitz to Walvis Bay (Fig. 4). Plagioclase prevails over K-feldspar.
0.2ϕ, σ 0.37 ± 0.02ϕ, Sk 0.0 ± 0.0, Ku 2.7 ± 0.0). Sand colour varies Mainly mafic volcanic rock fragments predominate over granitoid,
across the sand sea from light yellowish brown in coastal areas sedimentary (quartzose to feldspathoquartzose siltstone/sandstone,
(10 YR 6/4), to reddish yellow in central areas (7.5 YR 6/6), to yellow- shale, minor limestone and hybrid carbonate) and metamorphic
ish red in eastern areas (5 YR 5/8). The colour primarily reflects Fe- (quartz–sericite, quartz–mica, quartz–epidote, amphibolite) grains
oxide coatings on detrital grains (White et al., 2007), which are (Fig. 3D). Micas are absent. Rich heavy-mineral assemblages are
visibly thicker in the eastern Namib sands, reflecting both more clinopyroxene-dominated, with subordinate opaque Fe–Ti–Cr oxides,
ancient origin of the dunes and greater moisture content (Fig. 3E; garnet, epidote, and amphibole.
Folk, 1976; Walden and White, 1997). Weathering of heavy minerals Eastern Namib dunes are feldspathoquartzose, with less volcanic
is minor. Pyroxene is occasionally corroded, rarely etched, and rock fragments. K-feldspar prevails over plagioclase. Micas are
exceptionally skeletal in Orange sand; etched pyroxene grains with absent. Heavy-mineral assemblages, generally moderately-rich and
rounded or subrounded outline occur in Namib dunes. Amphibole clinopyroxene-dominated with Fe–Ti–Cr oxides, amphibole, garnet
and epidote grains may be corroded, but rarely etched. and epidote, are locally rich (NAM 9) or even extremely rich (NAM

Fig. 7. Detrital zircon U–Pb age spectra for the Orange River and Namib Sand Sea (modified after Vermeesch et al., 2010). Age spectra display a pronounced east–west dichotomy,
with coastal samples appearing nearly identical to the Orange sample, whereas samples near the edges of the desert show evidence for additional local sources.
Author's personal copy

182 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

Fig. 8. Petrography of hinterland-river sands. Sharply distinct from both Orange and Namib dune sand, composition ranges from quartzose metamorphiclastic (A), to quartzolithic
sedimentaclastic (B, C), and quartzofeldspathic plutoniclastic (D, E). Only the Fish River, an Orange tributary, similarly carries mafic volcanic and sedimentary detritus from Kalk-
rand basalts and underlying Karoo and Nama sedimentary rocks (F). K = K-feldspar; P = plagioclase; m = mica; L = lithics (Lv = volcanic; Lc = carbonate; Lt = siltstone; Lms =
metasandstone; Lm = metamorphic). All photos with crossed polars; blue bar = 250 μm.

8; Fig. 3E), and display a sharp increase in ultradense Fe–Ti–Cr indicating mixing with locally sourced populations of zircon grains
oxides and garnet at the expense of less dense pyroxene. In the NE (Fig. 7).
corner of the Namib, instead, assemblages are moderately poor
and include epidote, Fe–Ti–Cr oxides, garnet, clinopyroxene, amphi- 4. Provenance of Namib dune sand
bole, minor titanite, tourmaline, rutile, staurolite, andalusite, and
apatite (NAM 4–5). Pyroxene grains from dune sands at the SW, The proposed sources of Namib sand beside the Orange River
NW and N edges of the sand sea display very similar chemistry as include ephemeral hinterland rivers, whose detritus may be blown
those in Orange River sands (Fig. 5). to less peripheral parts of the sand sea by the northeasterly winter
U–Pb age distributions of detrital zircons remain virtually identical winds of the NE Namib, recycled ancestral fluvial or aeolian deposits,
to the Orange sample all along the coastal Namib, whereas the and deflation areas along the coast (e.g., Sperrgebiet; Besler, 1984;
distinct age peak at 1000 Ma becomes less pronounced inland. Lancaster, 1985).
The similarity of age populations is greater for samples along longitu- The strict compositional resemblance between Orange sand and
dinal profiles than along latitudinal profiles, indicating northward most desert dunes proves that most of the Namib sand is ultimately
sand migration, parallel to the coast and to the linear-dune system derived from the Orange River (Fig. 9). Specifically, overall abundance
that dominates the desert. Deviations from the Orange-sand and relative percentages of feldspar, volcanic-rock-fragment, and
spectrum are most marked at the northeastern edge of the erg, heavy-mineral species including clinopyroxene are virtually identical
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 183

Significant hinterland contributions are detected also at the SE


corner of the Namib. Landwards of Lüderitz (NAM 11 to NAM 9),
quartz increases at the expense of volcanic detritus, and garnet,
Fe–Ti–Cr oxides, amphibole (including green and green–brown
hornblende), epidote, titanite, staurolite and zircon become more
significant, suggesting contributions from the Koichab River and
locally exposed Namaqua basement rocks. In sample NAM 9, the
HCI index matches that of Koichab sand draining high-temperature
Namaqua rocks; concentration of garnet and Fe–Ti–Cr oxides is
instead ascribed to aeolian sorting (see below).
Quartz is relatively high and volcanic rock fragments low in the
Sossuvlei transect, where amphibole is slightly more abundant locally
Fig. 9. Namib dunes compare compositionally with Orange River sand, and are distinct
from hinterland-river sand. Affinity with Orange sand, greatest in the coastal Namib,
(NAM 6); strong concentration of garnet, Fe–Ti–Cr oxides and epidote
decreases inland and is least for NAM 4–5 dunes at the NE corner of the erg, indicating in sample NAM 8 is an aeolian-sorting effect. Such changes cannot be
mixing with local detrital sources. Cluster analysis performed with Aitchison distance accounted for by contributions from the Tsauchab River, which
(van den Boogaart and Tolosana-Delgado, 2008) on the whole bulk-petrography and carries abundant shale/slate and limestone rock fragments, little
heavy-mineral dataset.
quartz, few heavy minerals, and virtually no garnet.
A quantitative estimate of hinterland contributions shall be
attempted by comparing detrital modes of dune and hinterland-
in Orange and coastal Namib sands. Common mafic volcanic detritus river sands after all other factors controlling compositional variability
and rich clinopyroxene-dominated suites with minor epidote and are carefully considered.
amphibole are signatures that characterise no hinterland river drain-
ing into the Namib erg, and are even more extreme in coastal Namib 4.2. Recycling from Tertiary aeolianites
dunes than in the modern Orange itself. Such signatures best compare
with the Orange Quaternary terrace, suggesting greater contribution A major potential source of recycled sand is the up to 220 m-thick
from the upper Orange catchment in pre-dam times. Modern post- Tsondab Sandstone, underlying much of the modern sand sea (Ward,
dam Orange sediments, displaying detrital modes less dissimilar 1988). This unit displays quite similar morphology, geometry and
from those of hinterland rivers, may be derived in greater proportion mineralogy of modern linear dunes, and represents a Miocene ana-
from the lower catchment (for instance, relative to modern Orange logue of the present erg (Kocurek et al., 1999; Ségalen et al., 2004).
sand amphibole%tHM is half in coastal Namib dunes and Quaternary The Tsondab aeolianites are characterised by rich and pyroxene-
terrace, but invariably twice or more in hinterland rivers). dominated heavy-mineral assemblages (HMC 9 ± 8, Px 85 ±6%tHM),
Only minor and local supply from non-Orange sources is indepen- but for the northern edge of the Namib erg where heavy-mineral con-
dently confirmed by zircon U–Pb age spectra of Namib dune sands, centration is much less (HMC 1.1 ± 0.8), and common garnet is asso-
which compare closely with Orange sand (Fig. 7) rather than with ciated either with staurolite, epidote and tourmaline (south of the
ages of Namaqua or Damara basements exposed nearby. Abundance Kuiseb River downstream of Gobabeb), or with hornblende, zircon,
of zircon grains does not result to be greater in Orange sand (0.03 ± and rutile (Tsauchab and Tsondab valleys as far as the NAM 4 locality;
0.03%) than in hinterland-river sands (0.05 ± 0.05%), indicating that Dickinson and Ward, 1994; Besler, 1996).
relative budgets based on zircon-age spectra are not significantly The close mineralogical similarity between Tertiary aeolianites
biased by markedly different fertility of diverse sources of sediment and modern dunes suggests that the patterns of sand dispersal have
(Moecher and Samson, 2006). Zircon content in dune sands resulted not changed much since the Miocene at least, but prevents us to
to be similar (0.04 ± 0.04%, notably increasing to 0.2% in NAM 9 and establish how much detritus is recycled from Tertiary sandstones.
to 0.3% in NAM 8 due to aeolian-sorting effects, as discussed below), Recycling may be particularly significant in the type area, where
proving that such estimates are not inaccurate, although scarcely Tsondab aeolianites are more widely exposed, and similar petro-
precise because of the limited number of counted zircon grains. graphic changes are observed in Tertiary and modern dunes (Ward,
1988).
4.1. Contributions from hinterland rivers Fe–smectite and zeolite cements, as well as grain-dissolution
features, are commonly observed in Tsondab aeolianites (Dickinson
Virtually unchanged detrital modes from the Orange mouth to and Ward, 1994), but abundant pyroxene and high heavy-mineral
Walvis Bay reveal lack of additional sediment sources along the concentration (Besler, 1996) prove that the abundance of unstable
coast, and irrelevant supply from both Sperrgebiet outcrops in the detrital species is not significantly affected by diagenesis at such
south and Swakop mouth in the north. At the northern edge of the very shallow burial depths. The invariably drastic decrease of
desert, volcanic lithic fragments and pyroxene grains remain as abun- pyroxene close to the eastern margin of the sand sea, both in modern
dant, and epidote and amphibole as scarce, as in Orange or Lüderitz sands and Tertiary aeolianites (Besler, 1996), indicates that basement
dune sand. rocks east of the Namib erg do not represent a major source of
Only at the northeastern edge of the desert, where relict fluvial detrital pyroxene (as hypothesised instead by Lancaster and Ollier,
deposits associated with former courses of the Kuiseb and Tsondab 1983 p. 81).
Rivers occur (Lancaster et al., 2000), does the mineralogy of dune
sands drastically change, revealing the existence of local entry points 5. Physical processes controlling compositional variability
of detritus. Most evident are hinterland contributions at the NE of dune sand
corner of the Namib erg, where quartz is more abundant, volcanic
detritus scarce, and opaque–epidote–garnet assemblages poorer and In order to be properly and quantitatively understood, the
distinct from clinopyroxene-dominated Orange and coastal Namib multiple factors controlling sediment composition should be ideally
suites. Affinity with Kuiseb sand is indicated. Very minor supply singled out and studied one by one. This is easier to attempt in
from modern or ancient hinterland-river sediments cannot be modern desert settings, where mineralogy is only affected by physical
excluded for samples NAM 1, NAM 2 and NAM 3, which include processes, with negligible chemical weathering and no diagenesis.
garnet and trace staurolite. Although compositional variability still reflects the superposed effects
Author's personal copy

184 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

of diverse factors, and ambiguous interpretation may thus ensue (e.g., risk of amplifying analytical errors by such procedure must always
ultradense and durable garnet may increase locally because of selec- be taken into account.
tive destruction of labile components, recycling of underlying sand- Our calculations suggest that aeolian sorting accounts for ~ 20%
stones, contribution from external sources, or aeolian sorting), depletion in quartz, feldspar and lithic fragments in sample NAM 8,
favourable conditions in this respect are offered by the Namib Sand and corresponding four-fold enrichment in heavy minerals. Opaque
Sea, where mechanical abrasion and breakdown of detrital grains Fe–Ti–Cr oxides are estimated to be enriched 10–15 times, zircon ~ 8
with different durabilities can be monitored during long-distance times, garnet ~ 5 times, staurolite, epidote and titanite 3–4 times, py-
aeolian transport from the single principal entry-point in the south roxene and amphibole ~ 2 times. Modifications are lesser for sample
to the northern edges of the desert. Once the localization and compo- NAM 9, with ~ 3% depletion in quartz, feldspar and lithic fragments,
sitional signatures of potential external sources are identified, the and corresponding 30% enrichment in pyroxene and amphibole;
mechanical causes of compositional and morphological variabilities staurolite, epidote and titanite are estimated to be enriched ~ 2
can be quantitatively investigated in Namib dune sands, after assess- times, garnet ~ 3 times, zircon ~ 4 times, and Fe–Ti–Cr oxides ~ 5 times.
ment and removal of aeolian-sorting effects. Besides such locally strong effects, the abundance of dense and
ultradense minerals, average grain density and grain size do not
5.1. Aeolian sorting and placer formation show any systematic change with transport distance as far as the
northern edge of the erg. This lack of evidence for long-term sorting
Aeolian deposits may be enriched in dense minerals to various de- of grains according to density or size during northward transport is
grees by wind turbulence on dune flanks. Dense minerals are also apparently in contrast with the observed decline in abundance and
concentrated up the stoss side of dunes where flow velocity increases, size of diamond grains north of the Orange mouth (Jacob et al.,
whereas low-density minerals are preferentially entrained over the 2006 p. 496).
crest and down the slip face (Komar, 2007). Such effects abide simple
physical rules, and can thus be quantified and predicted (Garzanti et 5.2. Mechanical durability of diverse detrital components
al., 2008; 2009). For instance, epidote, titanite or diamond (densi-
ty ~ 3.5 g/cm 3) should be ~0.4 ϕ smaller than associated quartz, and The mineralogy of the Lüderitz dune (NAM 12) compares closely
enriched by ~ 3 times in placer sand with SRD index ~ 3.0 (diamond with Orange sand, and particularly with the Quaternary-terrace sam-
was never found in our samples, unfortunately). ple. The absence of micas and the only sporadic occurrence of carbon-
Aeolian-sorting effects can be assessed by determining the average ate, terrigenous and metamorphic lithic fragments, common to all
grain density for each sample, which in absence of environmental bias Namib dunes, are the sole notable differences. Compositional modifi-
is expected to be equal to average source-rock density (Garzanti and cations in high-energy beach settings thus appear to be limited to
Andò, 2007a). The SRD index of Orange sand is 2.72 ± 0.04, close to winnowing of slow-settling micas, widely dispersed offshore
the average density of the upper crust and thus of sediments carried (Spaggiari et al., 2006; Bluck et al., 2007), and to selective destruction
by many big rivers including the Nile (Garzanti et al., 2006; 2010). of non-durable sedimentary/metasedimentary rock fragments and
Eight out of twelve Namib dune samples have similar SRD indices alterites (Cameron and Blatt, 1971; McBride and Picard, 1987;
(2.71 ± 0.02). Such remarkable consistency testifies that the care Garzanti et al., 2002). Most severely depleted are shale/slate grains,
taken to collect samples unaffected by environmental bias was largely followed by carbonate, sandstone/metasandstone, and other meta-
successful. Higher indices characterise samples NAM 9 (HMC 9, SRD morphic rock fragments. All types of volcanic rock fragments, as
2.79) and particularly NAM 8 (HMC 25, SRD 3.03), reflecting wind- well as most heavy-mineral species, appear instead to healthily sur-
induced concentration of ultradense detrital components (Fe–Ti–Cr vive longshore transport; possible exceptions are apatite and amphi-
oxides 39–42%HM, garnet 11–19%HM, zircon 1–2%HM; Fig. 3E). bole, which show slight depletion.
Strong concentration of ultradense minerals was much more fre- The composition of dune sand remains virtually identical after an-
quently reported by Lancaster and Ollier (1983). Possibly due to a dif- other 600 km of transport as far as Walvis Bay at least, with the excep-
ferent sampling strategy, they found garnet > 20%HM in 15 out of tion of a slight relative enrichment in quartz and garnet (Fig. 10). Slight
their 26 samples, and even prevailing over clinopyroxene in 9 sam- enrichment in quartz and garnet is also observed in dune sand ~ 60 km
ples (which we only observed in samples NAM 5 and NAM 9, where NE of Lüderitz (NAM 11), whereas compositional changes further east
hinterland contributions are significant). (NAM 10–9) are largely ascribed to hinterland supply and aeolian sort-
Because the degree to which any denser mineral lags behind a less ing. Central eastern Namib sands (NAM 7–6 and NAM 8 SRD-corrected
dense mineral is primarily a function of their density ratios (Trask and for aeolian sorting) are slightly enriched in quartz and garnet, and de-
Hand, 1985), detrital minerals are systematically enriched in propor- pleted in volcanic rock fragments and pyroxene relative to Orange
tion to their density in samples with higher SRD. The observed com- sands and the Lüderitz dune (NAM 12). Along the northern Namib tra-
positions can thus be recalculated for each sample by iteratively verse, detrital modes of dune sand remain virtually identical for
correcting abundances of detrital species in proportion to their densi- 50–100 km (NAM 1–2–3), and next change chiefly because of supply
ties, until the correct SRD index is restored for each (SRD correction of from external sources (NAM 4–5). Micas are virtually absent in dune
Garzanti et al., 2009). This cannot be safely done for samples with sands even close to the Kuiseb River that carries abundant biotite,
lower grain density (SRD b 2.70) where either selective removal of suggesting efficient selective winnowing.
dense grains with anomalous physical behaviour (e.g., biotite flakes Comparing sand composition at the entry point (Lüderitz dune
supplied by the nearby Kuiseb River for samples NAM 4 and NAM NAM 12, where quartz is 56%, volcanic lithic fragments 9%, pyroxene
5) or quartz dilution (e.g., recycling of Tsondab Sandstone for samples 8%, and garnet 0.2% of bulk sand) with dunes showing only minor
NAM 6 and NAM 7) might have taken place. In such cases, the SRD hinterland contributions (NAM 1–2–3, 6–7–8, 10–11, where quartz
correction would restore an excess of densest components. is 65 ± 3%, volcanic lithic fragments 7 ± 3%, pyroxene 4 ± 1%, and gar-
For the other samples, aeolian-sorting effects were compensated net 0.6 ± 0.2% of bulk sand) allows us to assess the maximum possible
for by recalculating detrital modes to an SRD index of 2.715 ± 0.005. effect of selective breakdown during prolonged aeolian transport.
Such corrections are minor for most samples and most minerals Although sedimentary processes along the high-energy Namibia
(b20% even for garnet), but significant for placer sand NAM 8. The coast are held to be extreme, and severe sand abrasion in the defla-
compositional variability in homogeneous sample groups tends to de- tion basin of the Namib erg can destroy naturally exhumed bone ma-
crease after the SRD correction, indicating that even subtle environ- terial in a few months (Corbett and Burrell, 2001 p. 64), our data
mental bias can be detected and removed. On the other hand, the consistently indicate that only trivial compositional changes are
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 185

Fig. 10. Long-distance transport in high-energy coastal and aeolian settings does not selectively destroy volcanic rock fragments and pyroxene grains (compositional changes given
in log ratios). Whereas sedimentary and metamorphic rock fragments are effectively reduced during longshore transport from Orange mouth (OR) to Lüderitz (LÜ), the F/Rv ratio
remains unchanged to Walvis Bay (WB) and the Q/Rv ratio increases only slightly. The same is true during aeolian transport eastwards across the western Namib, whereas Orange-
derived volcanic detritus decreases eastwards across the eastern Namib chiefly by dilution of hinterland-derived sand. Inset shows changing ratios between key parameters. Q =
quartz; F = feldspars (P = plagioclase; K = K-feldspar); RF = rock fragments (Rv = volcanic; Rm = metamorphic; Rs = sedimentary); HMC = heavy-mineral concentration; Px =
pyroxene; Ep = epidote; Amp = amphibole; Grt = garnet; ZTR = zircon + rutile + tourmaline.

caused by mechanical breakage during transport over a distance of the Mediterranean Sea (Shukri, 1950; Garzanti et al., 2006). We can
600 km and more (Fig. 10). The slight increase in durable quartz thus conclusively contradict the common belief that mafic volcanic
and garnet across the sand sea cannot be ascribed to external sources detritus is readily worn away within a few tens of km of its source,
(hinterland rivers carry similar amounts or less quartz and garnet and that rift-related volcanic rocks are consequently not expected to
than the Orange River), but may be partly accounted for by recycling leave significant trace in the sedimentary record (Blatt, 1978;
of underlying Tertiary aeolianites. Ingersoll, 1984; Ingersoll et al., 1993 p. 942).
Particularly worthy of note is the durability of basaltic rock frag-
ments and pyroxene grains, which do not only endure more than
2000 km of transport from Lesotho highlands along the Orange 5.3. Rounding by aeolian abrasion
River, but survive in virtually unchanged proportions multistep trans-
port in high-energy coastal and aeolian environments for another Although aeolian impacts do not prove to be effective enough to
1000 km, prolonged over a time period as long as 1 Ma or more change sand composition by selectively destroying labile components,
(Vermeesch et al., 2010). Basaltic rock fragments and pyroxene grains they can spectacularly modify the morphology of detrital minerals,
are similarly observed to remain virtually unchanged after 4000 km which may become as well rounded as perfect spheres. Evident at
of fluvial transport along the Nile from Ethiopian rift highlands to first sight (Fig. 3C, F), this effect can be quantified by comparing the
Author's personal copy

186 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

roundness of diverse minerals in Namib dunes and in Orange and the average. For all samples we used a double data string, one with
hinterland river sands. bulk-petrography modes, recalculated to 100% after eliminating
Very few detrital grains are rounded in Orange sands (e.g., micas and labile sedimentary/metasedimentary lithic fragments but
apatite), where quartz, titanite, amphibole, pyroxene, or more rarely including total heavy-mineral content (i.e., HMC index), and the
other minerals may be subrounded. The same is true for hinterland- other with heavy minerals only, recalculated to 100%.
river sediments, where subrounded quartz and heavy minerals are Contrary to chemical processes, which can be safely considered as
observed more frequently in the Zebra/Tsauchab and Fish Rivers, negligible in such a hyperarid environment, physical processes are
recycling Nama Group or Karoo sedimentary rocks. relevant, and have to be evaluated and wherever possible corrected
In Namib dunes, instead, sand grains are mostly subrounded for. Most evident is heavy-mineral enrichment due to wind action
(Goudie and Watson, 1981), and commonly rounded. Only in sample in samples NAM 9 and particularly NAM 8, the detrital modes of
NAM 5 are grains uncommonly rounded (9%), and frequently which have been SRD-corrected to an SRD index of 2.715 ± 0.005.
angular/subangular (23%). Among heavy minerals, titanite appears For simplicity, raw data were used for all of the other samples.
as most commonly rounded (67% grains rounded, 28% grains sub-
rounded), followed by pyroxene (55% rounded, 39% subrounded), 6.1. Quantitative provenance assessment
opaque Fe–Ti–Cr oxides (49% rounded, 45% subrounded), apatite
(40% rounded, 60% subrounded), amphibole (40% rounded, 54% sub- Whereas sand composition varies little from the Orange mouth to
rounded), tourmaline (38% rounded, 50% subrounded), epidote (29% Walvis Bay, along a distance of nearly 1000 km along the coast, all
rounded, 50% subrounded), zircon (13% rounded, 65% subrounded), three studied E–W transects display a systematic variability in detrital
garnet (2% rounded, 79% subrounded), and rutile (8% rounded, 33% modes, with eastward increase in quartz (and particularly of sub-
subrounded). Positive correlation between mineral roundness and rounded to well-rounded grains) and parallel decrease in volcanic
hardness, as empirically measured by the Mohs' scale (Marsland lithic fragments and pyroxene. Such variability can be modelled as
and Woodruff, 1937; Dietz, 1973), is thus indicated (r = 0.72; 2% simple mixing of Orange-delivered sand with detritus produced in
sign. lev). the hinterland. As a measure of statistical similarity, we used the co-
Quartz appears to be least commonly rounded and subrounded at efficient of multiple correlation R between observed and modelled
Lüderitz, but evidence for systematic landward increase in quartz petrographic and heavy-mineral modes (for further methodological
roundness is weak along the southern and central transects. Pyroxene details see Vezzoli and Garzanti, 2009).
tends to become more commonly rounded landwards, but other min- Linear-mixing calculations indicate that dune sand of the coastal
erals display erratic trends. Conversely, angularity tends to increase Namib is virtually entirely Orange-derived, with external contribu-
inland along the northern transect, confirming the influence of hin- tions being ≤15% even at the northernmost edge of the sand sea
terland sources for dune samples NAM 4 and NAM 5. (NAM 1–2–3, R = 0.99). External sources become dominant only at
Sand grains thus appear to be efficiently and rapidly rounded in the northeastern edge of the erg (~75%; NAM 4, R = 0.95; ≤98%,
aeolian settings, whereas the effects of further multistep aeolian NAM 5, R = 0.97), where dune sand is largely derived from the Kuiseb
transport are barely evident and locally dimmed by supply of more River, either directly or indirectly by recycling of relict deposits.
angular detritus from hinterland rivers. Roundness may have been Quartz and heavy minerals are less commonly rounded than in
acquired up to an undetermined degree during high-energy long- other dunes, suggesting that recycling of Tertiary aeolianites is not
shore transport before entering the sand sea, or partly inherited extensive. Contribution from the Tsauchab River to the central east-
from recycling of Tsondab aeolianites. Comparison with sediment ern Namib is minor (≤15%; NAM 7–6 and NAM 8 SRD-corrected,
transport along the Nile coastal cell from the Delta to northern Israel 0.90 b R b 0.95). At the southern edge of the desert, the contribution
(own data) indicates that heavy-mineral roundness does increase from the Koichab River and locally exposed Namaqua basement
over a longshore distance of ~400 km, but such a change is at least rocks increases steadily eastwards from ≤ 15% (NAM 11, R = 0.97),
partly due to reworking of coastal dunes and aeolianites. to 35–40% (NAM 10, R = 0.96) and up to 70–75% (NAM 9 SRD-
corrected, R = 0.88). Because samples were collected from accessible
6. Sediment budget dunes at the edge of the sand sea in areas more prone to receiving
detritus from hinterland sources, external contributions are expected
Having assessed the diverse causes of compositional change, and to decrease in the interior of the erg.
determined the localizations and signatures of all possible external The similarity indices thus obtained (R = 0.95 ± 0.04) can be
sources (i.e., potential contributions from the Great Escarpment via improved by tentatively quantifying the progressive breakdown of
the Koichab, Tsaris, Tsauchab, Tsondab and Kuiseb Rivers in the mechanically less resistant grains, with consequent enrichment in
southeast to northeast, and along shore from the Swakop mouth in quartz. We obtained the best overall fit (R = 0.98 ± 0.01) by assuming
the northwest; Fig. 1), we can now attempt to quantitatively assess that most carbonate lithic fragments, over two/thirds of other sedi-
the Orange versus hinterland contributions to diverse parts of the mentary to low-rank metasedimentary lithic fragments, and 10–15%
Namib Sand Sea with simple statistical techniques (linear mixing of feldspar (Q/F increases from Lüderitz to Walvis Bay) are selectively
model; Weltje, 1997). destroyed in aeolian settings, and that garnet is enriched at the
Orange sand, including sedimentary to mafic volcanic lithic frag- expense of cleavable amphibole and of other heavy minerals to a
ments and clinopyroxene from Carboniferous to Jurassic rocks of the lesser extent. The sediment budget thus calculated is illustrated in
Karoo Supergroup, is readily distinguished from hinterland-river Fig. 11.
sand, containing granitoid and metasedimentary detritus with If these assumptions are correct, the estimated Tsauchab contribu-
epidote, amphibole, garnet and locally staurolite from Proterozoic to tions to the central eastern Namib may reach locally as high as
lowermost Palaeozoic basement rocks of the Namaqua and Damaran 30–40% (NAM 7–6–8, R = 0.98 ± 0.01), whereas hinterland contribu-
Orogens. Considering that the marked compositional variability of tions at the northwestern edge of the Namib become negligible (b5%,
modern Orange sand results at least in part from the modification of NAM 1–2, R = 1.00). The estimated contribution from the Koichab
natural sediment fluxes caused by dams and mining activities, we River and Namaqua rocks exposed along the southern edge of the
chose to discard modern sands of the lowermost course (S4037, erg (≤15%, NAM 11, R = 1.00; 35–40% NAM 10, R = 0.98; ~75%
NAM 13), and thus assumed as Orange end-member the average de- NAM 9, R = 0.96), and from the Kuiseb River at the northeastern
trital modes of samples NAM 14 and S4038. More robust results were edge (b10%, NAM 3, R = 0.98; ~75%, NAM 4, R = 0.97; ≤98%, NAM
obtained when the proximal Lüderitz dune (NAM 12) was included in 5, R = 0.97) remain virtually unchanged.
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 187

to recycling of underlying Tertiary aeolianites, is a slight northward


and eastward enrichment in quartz and garnet, and a slight depletion
in volcanic detritus in the central eastern Namib. Aeolian abrasion, a
relatively inefficient agent of compositional change, meets far greater
success in sculpting detrital grains, invariably much better rounded in
aeolian dunes than in both Orange and hinterland fluvial sands. The
roundness of quartz grains tends to improve with transport distance,
but heavy minerals appear nearly as well rounded in the Lüderitz
dune as in the rest of the erg, where a minority of subangular grains
remains. Roundness is thus acquired readily in the aeolian environ-
ment, with minor further improvement.
Beyond the origin of Namib dune sands, the new data presented
here solve issues that have remained controversial in sedimentary
petrology for long. We demonstrated the inefficacy of selective
mechanical breakdown in modifying sediment composition, even
through a sequence of extremely energetic fluvial, coastal and aeolian
environments, along a routing system of more than 3000 km, and
during time periods up to 1 Ma and more. In particular, we conclu-
sively proved the durability of basaltic rock fragments and pyroxene
grains. Rift-related volcanism thus can and does leave traces in the
sedimentary record, and represents an independent and significant
type of sediment provenance. Finally, we confirmed the efficiency of
wind-induced grain-to-grain collisions in sculpting detrital minerals,
which become rapidly rounded once passed from fluvial and marine
environments to the aeolian domain. Such newly acquired insight
on the effects of physical processes will result in higher resolution
for provenance interpretations of ancient sandstones, and thus in
more accurate reconstructions of climates and landscapes of the past.

Acknowledgments

The article benefited from comments and advice by Asish R. Basu


and an anonymous reviewer. Giles Wiggs and Charlie Bristow kindly
Fig. 11. Sediment budget for the Namib Sand Sea. Estimated contributions of hinterland
provided laser grain-size and geochronological data on Namib
sources are given for all studied samples across the erg. The coefficient of multiple cor- dunes. Marta Padoan and Alberto Resentini helped in the SEM
relation R measures the similarity between the observed and modelled detrital modes. analyses of pyroxene grains and statistical analysis.
Main paths of longshore and aeolian sediment transport are shown (wind directions
and potential annual sand flux after Lancaster, 1985; Bristow and Lancaster, 2004).
Appendix A. Supplementary material

7. Conclusion Supplementary material to this article, found online at doi:10.


1016/j.earscirev.2012.02.008, includes discussions on alternative
The Namib Sand Sea is chiefly fed by the Orange River, as docu- heavy-mineral counting methods (Appendix A1), on composition
mented by bulk petrography, heavy-mineral suites, percentages of and provenance of detrital pyroxene (Appendix A2) and on age-
feldspar species and volcanic-rock-fragment types, pyroxene chemis- patterns of detrital zircon (Appendix A3).
try, and U/Pb age populations of detrital zircons, which all remain
remarkably uniform all along the western, coastal part of the erg. References
Hinterland sediment sources do exist along the eastern margin of
the dune field, but major external contributions occur only locally at Ahrendt, H., Hunziker, J.C., Weber, K., 1978. Age and degree of metamorphism and time
of nappe emplacement along the southern margin of the Damara Orogen, Namibia
its NE and SE corners, where sand is largely provided by the Kuiseb (SW-Africa). Geologische Rundschau 67, 719–742.
and Koichab Rivers, respectively. Avigad, D., Sandler, A., Kolodner, K., Stern, R.J., McWilliams, M., Miller, N., Beyth, M.,
The peculiarity of the Namib erg, with essentially one single entry 2005. Mass-production of Cambro-Ordovician quartz-rich sandstone as a conse-
quence of chemical weathering of Pan-African terranes: environmental implica-
point of detritus supplying a variety of labile rock fragments and tions. Earth and Planetary Science Letters 240, 818–826.
heavy minerals, offers a unique opportunity to verify whether and Becker, T., Schreiber, U., Kampunzu, A.B., Armstrong, R., 2006. Mesoproterozoic rocks of
to what extent mechanical abrasion and selective breakdown of less Namibia and their plate tectonic setting. Journal of African Earth Sciences 46,
112–140.
durable components can effectively modify grain morphology and Besler, H., 1984. The development of the Namib dune field according to sedimentolog-
sand composition. Once local provenance and aeolian-sorting effects ical and geomorphological evidence. In: Vogel, J.C. (Ed.), The Late Cainozoic Palaeo-
are identified and compensated for, the composition of Orange- climates of the Southern Hemisphere. Balkema, Rotterdam, pp. 445–453.
Besler, H., 1996. The Tsondab sandstone in Namibia and its significance for the Namib
derived sand and specifically the abundance of basaltic lithic
erg. South African Journal of Geology 99, 77–87.
fragments and pyroxene grains are observed to remain virtually Blanco, G., Rajesh, H.M., Germs, G.J.B., Zimmermann, U., 2009. Chemical composition
unchanged along the Namibia coast and as far as the northern edge and tectonic setting of chromian spinels from the Ediacaran–Early Paleozoic
Nama Group, Namibia. Journal of Geology 117, 325–341.
of the erg. Substantial compositional modifications due to mechanical
Blatt, H., 1978. Sediment dispersal from Vogelsberg Basalt, Hessen, West Germany.
processes do take place before entering the erg in high-energy coastal Geologische Rundschau 67, 1009–1015.
settings, where micas are readily winnowed by waves and deposited Bluck, B.J., Ward, J.D., Cartwright, J., Swart, R., 2007. The Orange River, southern Africa:
offshore, and sedimentary to low-rank metasedimentary rock an extreme example of a wave-dominated sediment dispersal system in the South
Atlantic Ocean. Journal of the Geological Society of London 164, 341–351.
fragments (shale/slate, limestone) are selectively destroyed. Within Bremner, J.M., Rogers, J., Willis, J.P., 1990. Sedimentological aspects of the 1988 Orange
the sand sea, the only systematic change, possibly ascribed in part River floods. Transactions of the Royal Society of South Africa 47, 247–294.
Author's personal copy

188 E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189

Bristow, C.S., Lancaster, N., 2004. Movement of a small slipfaceless dome dune in the Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., Sares, S.W., 1984. The
Namib Sand Sea, Namibia. Geomorphology 59, 189–196. effect of grain size on detrital modes: a test of the Gazzi–Dickinson point-
Bristow, C.S., Duller, G.A.T., Lancaster, N., 2007. Age and dynamics of linear dunes in the counting method. Journal of Sedimentary Petrology 54, 103–116.
Namib desert. Geology 35, 555–558. Ingersoll, R.V., Kretchmer, A.G., Valles, P.K., 1993. The effect of sampling scale on
Cameron, K.L., Blatt, H., 1971. Durabilities of sand size schist and “volcanic” rock actualistic sandstone petrofacies. Sedimentology 40, 937–953.
fragments during fluvial transport, Elk Creek, Black Hills, South Dakota. Journal of Jacob, J., Ward, J.D., Bluck, B.J., Scholz, R.A., Frimmel, H.E., 2006. Some observations on
Sedimentary Petrology 41, 565–576. diamondiferous bedrock gully trapsites on Late Cainozoic, marine-cut platforms
Catuneanu, O., Wopfner, H., Eriksson, P.G., Cairncross, B., Rubidge, B.S., Smith, R.M.H., of the Sperrgebiet, Namibia. Ore Geology Reviews 28, 493–506.
Hancox, P.J., 2005. The Karoo basins of south-central Africa. Journal of Asian Jacobs, J., Pisarevsky, S., Thomas, R.J., Becker, T., 2008. The Kalahari Craton during the
Earth Sciences 43, 211–253. assembly and dispersal of Rodinia. Precambrian Research 160, 142–158.
Compton, J.S., Maake, L., 2007. Source of the suspended load of the upper Orange River, Jacobson, P.J., Jacobson, K.M., Seely, M.K., 1995. Ephemeral Rivers and their Catch-
South Africa. South African Journal of Geology 110, 339–348. ments: Sustaining People and Development in Western Namibia. Desert Research
Corbett, I., 1993. The modern and ancient pattern of sandflow through the southern Foundation of Namibia, Windhoek. 160 pp.
Namib deflation basin. International Association of Sedimentologists. Special Johnson, M.R., 1991. Sandstone petrography, provenance and plate tectonic setting in
Publication 16, 45–60. Gondwana context of the southeastern Cape–Karoo Basin. South African Journal
Corbett, I., Burrell, B., 2001. The earliest Pleistocene(?) Orange River fan-delta: an of Geology 94, 137–154.
example of successful exploration delivery aided by applied Quaternary research in Johnson, M.R., van Vuuren, C.J., Hegenberger, W.F., Key, R., Shoko, U., 1996. Stratigraphy
diamond placer sedimentology and palaeontology. Quaternary International 82, 63–73. in the Karoo Supergroup in southern Africa: an overview. Journal of African Earth
Cox, K.G., 1989. The role of mantle plumes in the development of continental drainage Sciences 23, 3–15.
patterns. Nature 342, 873–877. Johnsson, M.J., 1993. The system controlling the composition of clastic sediments. In:
Davies, G.R., Spriggs, A.J., Nixon, P.H., 2001. A non-cognate origin for the Gibeon Johnsson, M.J., Basu, A. (Eds.), Processes Controlling the Composition of Clastic
Kimberlite megacryst suite, Namibia: implications for the origin of Namibian Sediments: Geol. Soc. America, 284, pp. 1–19.
kimberlites. Journal of Petrology 42, 159–172. Jordaan, J.M., Clark, D., 1988. Regime changes in the Caledon River associated with
de Kock, G., 2001. A reappraisal of the Namibian Damara stratigraphy in part of the Welbedacht Barrage: physical model, prototype and theoretical correlations. In:
Southern Swakop Terrane and its implications to basin evolution. South African White, W.R. (Ed.), Int. Conf. on River Regime. Hydraulic Research Ltd., J. Wiley,
Journal of Geology 104, 115–136. Wallingford, England, pp. 375–384. 18–20 May 2008.
de Villiers, S., Compton, J.S., Lavelle, M., 2000. The strontium isotope systematics of the Jourdan, F., Féraud, G., Bertrand, H., Kampunzu, A.B., Tshoso, G., Watkeys, M.K., Le Gall,
Orange River, Southern Africa. South African Journal of Geology 103, 237–248. B., 2005. Karoo large igneous province: brevity, origin, and relation to mass extinc-
Dickinson, W.W., Ward, J.D., 1994. Low depositional porosity in eolian sands and sand- tion questioned by new 40Ar/39Ar age data. Geology 33, 745–748.
stones, Namib desert. Journal of Sedimentary Research 64, 226–232. Jung, S., Mezger, K., 2003. Petrology of basement-dominated terranes: I. Regional meta-
Dietz, V., 1973. Experiments on the influence of transport on shape and roundness of morphic T–t path from U–Pb monazite and Sm–Nd garnet geochronology (Central
heavy minerals. Schweizerbart, Stuttgart. Contributions to Sedimentology 1, 69–102. Damara Orogen, Namibia). Chemical Geology 198, 223–247.
Dott, R.H., 2003. The importance of eolian abrasion in supermature quartz sandstones Kocurek, G., Lancaster, N., Carr, M., Frank, A., 1999. Tertiary Tsondab Sandstone Forma-
and the paradox of weathering on vegetation-free landscapes. Journal of Geology tion: preliminary bedform reconstruction and comparison to modern Namib sand
111, 387–405. sea dunes. Journal of African Earth Sciences 29, 629–642.
Dutta, P.K., Zhou, Z., dos Santos, P.R., 1993. A theoretical study of mineralogical matu- Komar, P.D., 2007. The entrainment, transport and sorting of heavy minerals by
ration of eolian sand. Geological Society of America, Special Paper 284, 203–209. waves and currents. In: Mange, M.A., Wright, D.T. (Eds.), Heavy Minerals in Use.
Folk, R.L., 1976. Reddening of desert sands: Simpson Desert, Northern Territory, Aus- Developments in Sedimentology Series, 58. Elsevier, Amsterdam, pp. 3–48.
tralia. Journal of Sedimentary Petrology 46, 604–615. Elsevier, Amsterdam.
Frimmel, H.E., Frank, W., 1998. Neoproterozoic tectono-thermal evolution of the Gariep Kuenen, P.H., 1959. Experimental abrasion; 3, fluviatile action on sand. American
Belt and its basement, Namibia and South Africa. Precambrian Research 90, 1–28. Journal of Science 257, 172–190.
Galerne, C.Y., Neumann, E.-R., Planke, S., 2010. Magmatic differentiation processes in Kuenen, P.H., 1960. Experimental abrasion; 4, Eolian action. Journal of Geology 68,
saucer-shaped sills: evidences from the Golden Valley Sill in the Karoo Basin, 427–449.
South Africa. Geosphere 6, 163–188. Lancaster, N., 1985. Winds and sand movement in the Namib Sand Sea. Earth Surface
Garzanti, E., Andò, S., 2007a. Heavy-mineral concentration in modern sands: implica- Processes and Landforms 10, 607–619.
tions for provenance interpretation. In: Mange, M.A., Wright, D.T. (Eds.), Heavy Lancaster, N., 1989. The Namib Sand Sea: Dune Forms, Processes, and Sediments. Balkema,
Minerals in Use, Developments in Sedimentology Series, 58. Elsevier, Amsterdam, Rotterdam. 200 pp.
pp. 517–545. Lancaster, N., 2002. How dry was dry? Late Pleistocene palaeoclimates in the Namib
Garzanti, E., Andò, S., 2007b. Plate tectonics and heavy-mineral suites of modern sands. desert. Quaternary Science Reviews 21, 769–782.
In: Mange, M.A., Wright, D.T. (Eds.), Heavy Minerals in Use, Developments in Lancaster, N., Ollier, C.D., 1983. Sources of sand for the Namib Sand Sea. Zeitschrift für
Sedimentology Series, 58. Elsevier, Amsterdam, pp. 741–763. Geomorphologie 45, 71–83 Supplement band.
Garzanti, E., Vezzoli, G., 2003. A classification of metamorphic grains in sands based on Lancaster, N., Schaber, G.G., Teller, J.T., 2000. Orbital radar studies of paleodrainages in
their composition and grade. Journal of Sedimentary Research 73, 830–837. the Central Namib desert. Remote Sensing of Environment 71, 216–225.
Garzanti, E., Canclini, S., Moretti Foggia, F., Petrella, N., 2002. Unraveling magmatic and Lustrino, M., Melluso, L., Brotzu, P., Gomes, C.B., Morbidelli, L., Muzio, R., Ruberti, E.,
orogenic provenances in modern sands: the back-arc side of the Apennine thrust- Tassinari, C.C.G., 2005. Petrogenesis of the Early Cretaceous Valle Chico igneous
belt (Italy). Journal of Sedimentary Research 72, 2–17. complex (SE Uruguay): relationships with Paraná–Etendeka magmatism. Lithos
Garzanti, E., Andò, S., Vezzoli, G., Dell' Era, D., 2003. From rifted margins to foreland 82, 407–434.
basins: investigating provenance and sediment dispersal across desert Arabia Makhoalibe, S., 1984. Suspended sediment transport measurement in Lesotho.
(Oman, UAE). Journal of Sedimentary Research 73, 572–588. Challenges in African Hydrology and Water Resources: IAHS Publ., 144,
Garzanti, E., Andò, S., Vezzoli, G., Ali Abdel Megid, A., El Kammar, A., 2006. Petrology of pp. 313–321.
Nile River sands (Ethiopia and Sudan): sediment budgets and erosion patterns. Makhoalibe, S., 1999. Management of water resources in the Maloti/Drakensberg
Earth and Planetary Science Letters 252, 327–341. mountains of Lesotho. Ambio 28, 460–461.
Garzanti, E., Andò, S., Vezzoli, G., 2008. Settling-equivalence of detrital minerals and Mange, A., Maurer, H.F.W., 1992. Heavy Minerals in Colour. Chapman and Hall, London.
grain-size dependence of sediment composition. Earth and Planetary Science 147 pp.
Letters 273, 138–151. Marsland, P.S., Woodruff, J.G., 1937. A study of the effects of wind transportation on
Garzanti, E., Andò, S., Vezzoli, G., 2009. Grain-size dependence of sediment composi- grains of several minerals. Journal of Sedimentary Petrology 7, 18–30.
tion and environmental bias in provenance studies. Earth and Planetary Science McBride, E., Picard, D.M., 1987. Downstream changes in sand composition, roundness
Letters 277, 422–432. and gravel size in a short-headed, high-gradient stream, Northwestern Italy.
Garzanti, E., Andò, S., France-Lanord, C., Vezzoli, G., Censi, P., Galy, V., Najman, Y., 2010. Journal of Sedimentary Petrology 57, 1018–1026.
Mineralogical and chemical variability of fluvial sediments. 1. Bedload sand (Ganga- McLennan, S.M., 2001. Relationships between the trace element composition of
Brahmaputra, Bangladesh). Earth and Planetary Science Letters 299, 368–381. sedimentary rocks and upper continental crust. Geochemistry, Geophysics, Geo-
Geyer, G., 2005. The Fish River Subgroup in Namibia: stratigraphy, depositional envi- systems 2 2000GC000109.
ronments and the Proterozoic–Cambrian boundary problem revisited. Geological Mehring, J.L., McBride, E.F., 2007. Origin of modern quartzarenite beach sands in a
Magazine 142, 465–498. temperate climate, Florida and Alabama, USA. Sedimentary Geology 201,
Goudie, A.S., Eckardt, F., 1999. The evolution of the morphological framework of the 432–445.
central Namib desert, Namibia, since the Early Cretaceous. Geografiska Annaler Melluso, L., Cucciniello, C., Petrone, C.M., Lustrino, M., Morra, V., Tiepolo, M.,
81A, 443–458. Vasconcelos, L., 2008. Petrology of Karoo volcanic rocks in the southern Lebombo
Goudie, A.S., Watson, A., 1981. The shape of desert sand dune grains. Journal of Arid monocline, Mozambique. Journal of African Earth Sciences 52, 139–151.
Environments 4, 185–190. Miller, R.M., 1983. The Pan-African Damara Orogen of South West Africa/Namibia. In:
Haskins, D.R., Bell, F.G., 1995. Drakensberg basalts: their alteration, breakdown and Miller, R.M. (Ed.), Evolution of the Damara orogen of South West Africa/Namibia:
durability. Quarterly Journal of Engineering Geology 28, 287–302. Geol. Soc. South Africa, Spec. Publ., 11, pp. 431–515.
Hu, Z., Gao, S., 2008. Upper crustal abundances of trace elements: a revision and Miller, R.M., 2008. The Geology of Namibia, 3 vol. Ministry of Mines and Energy,
update. Chemical Geology 253, 205–221. Windhoek.
Ingersoll, R.V., 1984. Felsic sand/sandstone within mafic volcanic provinces: rift sand/ Moecher, D.P., Samson, S.D., 2006. Differential zircon fertility of source terranes and
sandstone shows no basaltic provenance. Geological Society of America, Abstracts natural bias in the detrital zircon record: implications for sedimentary provenance
with Programs 16 548 pp. analysis. Earth and Planetary Science Letters 247, 252–266.
Author's personal copy

E. Garzanti et al. / Earth-Science Reviews 112 (2012) 173–189 189

Moore, A., Blenkinsop, T., 2002. The role of mantle plumes in the development of Schlüter, T., 2006. Geological Atlas of Africa, with Notes on Stratigraphy, Tectonics, Eco-
continental-scale drainage patterns: the southern African example revisited. nomic Geology, Geohazards and Geosites of Each Country. Springer, Heidelberg.
South African Journal of Geology 105, 353–360. 272 pp.
Moore, A., Blenkinsop, T., Cotterill, F., 2008. Controls on post-Gondwana alkaline volca- Ségalen, L., Rognon, P., Pickford, M., Senut, B., Emmanuel, L., Renard, M., Ward, J., 2004.
nism in Southern Africa. Earth and Planetary Science Letters 268, 151–164. Reconstitution of dune morphologies and palaeowind regimes in the Proto-Namib
Moore, A., Blenkinsop, T., Cotterill, F., 2009. Southern African topography and erosion since the Miocene. Bulletin de la Societe Géologique de France 175, 537–546.
history: plumes or plate tectonics? Terra Nova 21, 310–315. Shukri, N.M., 1950. The mineralogy of some Nile sediments. Quarterly Journal Geolog-
Morimoto, N., Fabries, J., Ferguson, A.K., Ginzburg, I.V., Ross, M., Seifert, F.A., Zussman, J., ical Society of London 105, 511–534.
Aoki, K., Gottardi, G., 1988. Nomenclature of pyroxenes. Mineralogy and Petrology Smith, R.M.H., Eriksson, P.G., Botha, W.J., 1993. A review of the stratigraphy and sedi-
39, 55–76. mentary environments of the Karoo-aged basins of Southern Africa. Journal of Af-
Morin, E., Grodek, T., Dahan, O., Benito, G., Kulls, C., Jacoby, Y., Van Langenhove, G., rican Earth Sciences 16, 143–169.
Seely, M., Enzel, Y., 2009. Flood routing and alluvial aquifer recharge along the Spaggiari, R.I., Bluck, B.J., Ward, J.D., 2006. Characteristics of diamondiferous Plio-
ephemeral arid Kuiseb River, Namibia. Journal of Hydrology 368, 262–275. Pleistocene littoral deposits within the palaeo-Orange River mouth, Namibia. Ore
Muhs, D.R., 2004. Mineralogical maturity in dunefields of North America, Africa and Geology Reviews 28, 475–492.
Australia. Geomorphology 59, 247–269. Stollhofen, H., Stanistreet, I.G., Bangert, B., Grill, H., 2000. Tuffs, tectonism and glacially
Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred related sea-level changes, Carboniferous–Permian, southern Namibia. Palaeogeo-
from major element chemistry of lutites. Nature 299, 715–717. graphy, Palaeoclimatology, Palaeoecology 161, 127–150.
Padoan, M., Garzanti, E., Harlavan, Y., Villa, I.M., 2011. Tracing Nile sediment sources by Swanevelder, C.J., 1981. Utilising South Africa's largest river: the physiographic back-
Sr and Nd isotope signatures (Uganda, Ethiopia, Sudan). Geochimica et Cosmochi- ground to the Orange River scheme. GeoJournal Suppl. 2, 29–40.
mica Acta 75, 3627–3644. Sweeney, R.J., Duncan, A.R., Erlank, A.J., 1994. Geochemistry and petrogenesis of
Parker, A., 1970. An index of weathering for silicate rocks. Geological Magazine 107, Central Lebombo basalts of the Karoo igneous province. Journal of Petrology 35,
501–504. 95–125.
Powers, M.C., 1953. A new roundness scale for sedimentary particles. Journal of Sedi- Taylor, S.R., McLennan, S.M., 1995. The geochemical evolution of the continental crust.
mentary Petrology 23, 117–119. Reviews of Geophysics 33, 241–265.
Quinn, K.A., Byrne, R.H., Schijf, J., 2006. Sorption of yttrium and rare earth elements by Trask, C.B., Hand, B.M., 1985. Differential transport of fall-equivalent sand grains, Lake
amorphous ferric hydroxide: influence of pH and ionic strength. Marine Chemistry Ontario, New York. Journal of Sedimentary Petrology 55, 226–234.
99, 128–150. Twenhofel, W.H., 1945. The rounding of sand grains. Journal of Sedimentary Petrology
Rogers, J., 1977. Sedimentation on the continental margin off the Orange River and the 15, 59–71.
Namib Desert. Geol. Survey/Univ. Cape Town Marine Geosci. Group Bull., 7. 162 pp. van den Boogaart, K.G., Tolosana-Delgado, R., 2008. “Compositions”: a unified R pack-
Rogers, J., Rau, A.J., 2006. Superficial sediments of the wave-dominated Orange River age to analyze compositional data. Computers & Geosciences 34, 320–338.
delta and the adjacent continental margin off southwestern Africa. African Journal Vermeesch, P., Fenton, C.R., Kober, F., Wiggs, G.F.S., Bristow, C.S., Xu, S., 2010. Sand res-
of Marine Science 28, 511–524. idence times of one million years in the Namib Sand Sea from cosmogenic nuclides.
Rooseboom, A., Lotriet, H.H., 1992. The new sediment yield map for southern Africa. Nature Geoscience. doi:10.1038/NGEO985.
Erosion and Sediment Transport Monitoring Programmes in River Basins: IAHS Vezzoli, G., Garzanti, E., 2009. Tracking paleodrainage in Pleistocene foreland basins.
Publ., 210, pp. 527–538. Journal of Geology 117, 445–454.
Rooseboom, A., Harmse, H.J. von M., 1979. Changes in sediment load of the Orange Walden, J., White, K., 1997. Investigation of the controls on dune colour in the Namib
River during the period 1929–1969. Hydrology of Areas of Low Precipitation: sand sea using mineral magnetic analyses. Earth and Planetary Science Letters
IAHS Publ., 128, pp. 459–479. 152, 187–201.
Russell, R.D., Taylor, R.E., 1937. Roundness and shape of Mississippi River sands. Journal Ward, J.D., 1988. Eolian, fluvial and pan (playa) facies of the Tertiary Tsondab Sand-
of Geology 45, 225–267. stone Formation in the central Namib Desert, Namibia. Sedimentary Geology 55,
Sawadogo, O. 2008. Analysis of observed reservoir sedimentation rates in South Africa. 143–162.
Thesis, University of Stellenbosch, http://www.aims.ac.za/resources/archive/2007/ Weltje, G.J., 1997. End-member modelling of compositional data: numerical–statistical
ousmane.pdf. algorithms for solving the explicit mixing problem. Journal of Mathematical Geol-
Saylor, B.Z., Grotzinger, J.P., Germs, G.J.B., 1995. Sequence stratigraphy and sedimentol- ogy 29, 503–549.
ogy of the Neoproterozoic Kuibis and Schwarzrand Subgroups (Nama Group), White, K., Walden, J., Gurney, S.D., 2007. Spectral properties, iron oxide content and
southwestern Namibia. Precambrian Research 73, 153–171. provenance of Namib dune sands. Geomorphology 86, 219–229.

View publication stats

You might also like