Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-06-01T16:12:15.749Z Has data issue: false hasContentIssue false

A ca. 39,000-year record of vegetation and climate change from the margin of the Namib Sand Sea

Published online by Cambridge University Press:  11 July 2023

Brian M. Chase*
Affiliation:
Institut des Sciences de l'Evolution-Montpellier (ISEM), University of Montpellier, Centre National de la Recherche Scientifique (CNRS), EPHE, IRD, Montpellier, France Department of Environmental and Geographical Science, University of Cape Town, South Lane, Upper Campus, 7701 Rondebosch, South Africa
Arnoud Boom
Affiliation:
School of Geography, Geology and the Environment, University of Leicester, Leicester, LE1 7RH, UK
Andrew S. Carr
Affiliation:
School of Geography, Geology and the Environment, University of Leicester, Leicester, LE1 7RH, UK
Michael E. Meadows
Affiliation:
Department of Environmental and Geographical Science, University of Cape Town, South Lane, Upper Campus, 7701 Rondebosch, South Africa School of Geography and Ocean Sciences, Nanjing University, China
Sophak Lim
Affiliation:
Institut des Sciences de l'Evolution-Montpellier (ISEM), University of Montpellier, Centre National de la Recherche Scientifique (CNRS), EPHE, IRD, Montpellier, France
*
*Corresponding author email address: Brian.Chase@umontpellier.fr
Rights & Permissions [Opens in a new window]

Abstract

This paper presents the first continuous multi-proxy record of climate and vegetation change from the central Namib Desert extending over much of the last ca. 39,000 years. Derived from rock hyrax middens, evidence from stable carbon and nitrogen isotopes, pollen, and microcharcoal reveals significant differences between glacial-age and Holocene climates and vegetation types. Although still arid to semi-arid, conditions during Marine Oxygen Isotope Stages (MIS) 2–3 were significantly more humid than in the Late Holocene. Considerable associated vegetation change is apparent, with cooler temperatures and higher/more-regular rainfall promoting the westward expansion of relatively mesic shrubby karroid vegetation during MIS 2–3. With the last glacial–interglacial transition, increasing temperatures and less/less-regular rainfall resulted in marked vegetation changes and the establishment of current xeric grasslands. The inter-plant spacing of the karroid vegetation promoted by wetter conditions does not carry fire effectively, and the microcharcoal record indicates that more extensive fires may develop only with the development of grassier vegetation under drier conditions. As with other terrestrial records from the Namib Desert and environs, no Cape flora elements were found to support previously hypothesised expansion of the Fynbos Biome during the last glacial period.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © University of Washington. Published by Cambridge University Press, 2023

INTRODUCTION

Reliable information regarding past climate and vegetation change in the Namib Desert region is extremely difficult to obtain because the arid environment is not conducive to preservation of organic material and associated proxies (see Lancaster, Reference Lancaster2002; Chase and Meadows, Reference Chase and Meadows2007). Indirect evidence from marine sediments has thus been drawn upon extensively to compensate for this lack of terrestrial archives (e.g., Shi et al., Reference Shi, Schneider, Beug and Dupont2001; Stuut et al., Reference Stuut, Prins, Schneider, Weltje, Jansen and Postma2002; Collins et al., Reference Collins, Schefuß, Govin, Mulitza and Tiedemann2014). However, the influences on such marine records of the region's complex circulation and taphonomic dynamics are difficult to assess, and clear discrepancies exist with those terrestrial records that do exist (Scott et al., Reference Scott, Marais and Brook2004, Reference Scott, Gil Romera, Marais and Brook2018; Lim et al., Reference Lim, Chase, Chevalier and Reimer2016; Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019). As a result, a clear understanding of long-term vegetation change in the Namib Desert region remains to be established.

As an iconic feature of southern Africa's diverse geography, the Namib Desert, which extends over 2000 km along the southwestern coast of Africa, has long been a focus of paleoenvironmental study. Arid conditions in the region likely date to at least the late Cretaceous (ca. 80 Ma) (Ward et al., Reference Ward, Seely and Lancaster1983), while intensification of the South Atlantic Anticyclone and the Benguela upwelling system since the Eocene (ca. 34 Ma) is likely to have induced increasingly arid conditions through higher atmospheric pressure and cooler coastal waters (van Zinderen Bakker, Reference van Zinderen Bakker1975; Tankard and Rodgers, Reference Tankard and Rodgers1978; Etourneau et al., Reference Etourneau, Martinez, Blanz and Schneider2009). Within this context and based on these mechanisms, it has been suggested that the last glacial period, particularly Marine Oxygen Isotope Stages (MIS) 2–4 (12–70 ka), which experienced marked increases in anticyclonic intensity and coastal upwelling (Little et al., Reference Little, Schneider, Kroon, Price, Bickert and Wefer1997; Stuut et al., Reference Stuut, Prins, Schneider, Weltje, Jansen and Postma2002; Farmer et al., Reference Farmer, deMenocal and Marchitto2005; Pichevin et al., Reference Pichevin, Cremer, Giraudeau and Bertrand2005), would have been a period of increased aridity in the Namib Desert (van Zinderen Bakker, Reference van Zinderen Bakker1975). However, studies of relict lacustrine deposits from across the Namib, including sites such as Kannikwa (Beaumont, Reference Beaumont1986), Koichab Pan (Lancaster, Reference Lancaster1984), Narabeb (Teller and Lancaster, Reference Teller and Lancaster1986), Khommabes (Teller and Lancaster, Reference Teller and Lancaster1985), and Gobabeb (Vogel and Visser, Reference Vogel and Visser1981) indicate periods of more/more-effective precipitation (Δprecipitation:potential evapotranspiration) during much of MIS 2–3 (37–31 cal ka BP, 30–24 cal ka BP), as well as the during the late glacial and earliest Holocene (14–10 cal ka BP) (Lancaster, Reference Lancaster2002). Much of these early data are derived from paleolacustrine and paleofluvial (Marker and Müller, Reference Marker and Müller1978; Vogel, Reference Vogel1982; Ward, Reference Ward1982; Heine and Heine, Reference Heine and Heine2002) archives associated with hydrologic networks sourced in the relatively mesic uplands of the Great Escarpment, to the east of the Namib Desert. Thus, an often-posed question has been to what extent does evidence from these archives reflect changes in the uplands or in the desert itself (Ward et al., Reference Ward, Seely and Lancaster1983; Heine, Reference Heine1998; Lancaster, Reference Lancaster2002)?

Questions regarding (1) the significance of phases of increased humidity inferred from discrete lacustrine deposits, and (2) the loci of the rainfall that fostered development of the region's lacustrine and fluvial features are being addressed via a wider network of new paleoenvironmental archives. In recent years, it has been shown that rock hyrax middens (Chase et al., Reference Chase, Scott, Meadows, Gil-Romera, Boom, Carr, Reimer, Truc, Valsecchi and Quick2012) can provide such valuable archives. Several continuous high-resolution records from rock hyrax middens from the Namib Desert have indicated that the last glacial period was indeed relatively humid compared to the Holocene (Lim et al., Reference Lim, Chase, Chevalier and Reimer2016; Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), adding significant detail to our understanding of regional climate change over the last 50,000 years. While some middens, such as those from Pella (Lim et al., Reference Lim, Chase, Chevalier and Reimer2016), are located at the margins of the Namib Desert, and may be considered to reflect conditions in the Great Escarpment, others, such those recovered from Spitzkoppe (Chase et al., Reference Chase, Meadows, Scott, Thomas, Marais, Sealy and Reimer2009, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019) and, particularly, Mirabib (Scott et al., Reference Scott, Gil Romera, Marais and Brook2018), are located to the west and lie within the arid core of the desert. In contrast to the region's fluvio-lacustrine archives, it is generally considered that rock hyrax middens preserve a predominantly local signal that reflects the foraging range of the animals (less than ~60 m from the midden/shelter site) (Sale, Reference Sale1965; Chase et al., Reference Chase, Scott, Meadows, Gil-Romera, Boom, Carr, Reimer, Truc, Valsecchi and Quick2012) and that, accordingly, the indications of increased glacial-age humidity at Spitzkoppe and Mirabib can be interpreted to reflect conditions within the desert itself.

Strong similarities are evident in the stable nitrogen isotope records (interpreted as a reflection of water availability) from middens recovered from Aba Huab/Austerlitz (Chase et al., Reference Chase, Meadows, Carr and Reimer2010), Spitzkoppe (Chase et al., Reference Chase, Meadows, Scott, Thomas, Marais, Sealy and Reimer2009, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), Zizou (Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), and Pella (Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), indicating a generally common climate response signal across a 900-km transect of the Namib Desert (Fig. 1). Evidence of associated changes in vegetation, however, is often restricted to the Holocene (Scott et al., Reference Scott, Cooremans, de Wet and Vogel1991, Reference Scott, Gil-Romera, Marais and Brook2022; Scott, Reference Scott1996; Gil-Romera et al., Reference Gil-Romera, Scott, Marais and Brook2006, Reference Gil-Romera, Scott, Marais and Brook2007), with the only indications of glacial-age vegetation coming from the continuous 50-kyr record from Pella (Lim et al., Reference Lim, Chase, Chevalier and Reimer2016), and shorter MIS 2–3 time slices from the Brandberg (Scott et al., Reference Scott, Marais and Brook2004) and Mirabib (Scott et al., Reference Scott, Gil Romera, Marais and Brook2018). While each record expresses changes specific to its environmental context, evidence from all of these sites contrasts markedly with pollen records recovered from marine cores off the coast of Namibia (Shi et al., Reference Shi, Dupont, Beug and Schneider2000, Reference Shi, Schneider, Beug and Dupont2001; Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015). A distinctive component in these marine records—both in terms of their content and the associated proxy interpretations—is the prevalence of pollen types associated with the Cape Floristic Region during the Pleistocene. This is evinced in levels of Restionaceae pollen as high as 14–20% during MIS 2–3 (Shi et al., Reference Shi, Dupont, Beug and Schneider2000, Reference Shi, Schneider, Beug and Dupont2001; Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015). In some cases, the addition of other pollen types, such as Anthospermum, Stoebe-type, and Artemisia pollen, that are found in, but are not specific to, fynbos augments the percentage of pollen from “fynbos taxa” to as much as 40% of the pollen sum (e.g., Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015). Such conclusions have played a significant role in theories regarding expansions of the southwest African winter rainfall zone and Fynbos Biome to the north during glacial periods (Chase and Meadows, Reference Chase and Meadows2007; Shi et al., Reference Shi, Dupont, Beug and Schneider2000, Reference Shi, Schneider, Beug and Dupont2001; Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015).

Figure 1. (A) Map of study region, indicating location of the Zizou rock hyrax midden site. (B) The Zizou-1 rock hyrax midden (21.3 cm thick), with view of local landscape and vegetation (C) and (D) view eastward from the site towards the Namib Sand Sea (visible in distance).

As early as 2004 the reliability of interpretations of marine pollen records in the region was questioned, based on an absence of Restionaceae and Ericaceae pollen in glacial-age samples from the Brandberg hyrax middens (Scott et al., Reference Scott, Marais and Brook2004). As a single site, this could not provide definitive refutation (see Chase and Meadows, Reference Chase and Meadows2007), but the absences of these taxa in subsequent records (Scott et al., Reference Scott, Gil Romera, Marais and Brook2018), including sites on the southern boundary of the Namib (Lim et al., Reference Lim, Chase, Chevalier and Reimer2016), implies that the microfossils from marine cores in the Benguela system may be significantly influenced by long-distance transport. As such, they cannot therefore be considered a direct reflection of environmental change on the adjacent continent (Scott et al., Reference Scott, Marais and Brook2004, Reference Scott, Gil Romera, Marais and Brook2018; Lim et al., Reference Lim, Chase, Chevalier and Reimer2016; Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019).

To characterise the nature of environmental and vegetation change in the Namib Desert, and to further test the hypothesis of substantial expansions of the Fynbos Biome during the last glacial period, we present new pollen, microcharcoal, and stable isotope records from the Zizou rock hyrax midden site. Located on the plains below the Great Escarpment on the eastern margin of the Namib Sand Sea (Figs. 1, 2), the Zizou rock hyrax midden site encompasses much of the last 39,000 years and allows an analysis of changes in vegetation and fire regimes within the context of an independent perspective (via stable isotope data) on climate variability.

Figure 2. Maps of study region showing Aridity Index gradients (data from Trabucco and Zomer, Reference Trabucco and Zomer2019), an elevation transect across the map at the latitude of the Zizou hyrax midden site, the distribution of ecoregions (from Dinerstein et al., Reference Dinerstein, Olson, Joshi, Vynne, Burgess, Wikramanayake and Hahn2017), and the association of each vegetation type in the region with aridity.

Regional setting

The Zizou rock hyrax midden site is located on the eastern margin of the northern Namib Sand Sea, 20 km SSW of the town of Solitaire (24.07°S, 15.97°E, 1120 m asl) (Figs. 1, 2). Situated at the base of the western flank of the Great Escarpment, the site lies within the steep environmental gradient between the hyper-arid regions of the Namib Desert and the semi-arid interior (Fig. 2). The limited precipitation received at the site (130 mm/yr on average, but with a high degree of interannual variability) falls almost exclusively during the austral summer (Fick and Hijmans, Reference Fick and Hijmans2017). Vegetation at the site is classified as xeric Namibian Savanna Woodland (Fig. 2), characterised by sparse C4 grass cover on the plains surrounding the site, with sporadic shrubs of the Capparaceae family (Boscia, Maerua) and isolated small trees (Vachellia and Searsia) along drainage lines (Fig. 1). Shrubs and arboreal taxa are more common on the rocky slopes surrounding the midden site, including the occurrence of Commiphora. To the west, the Namib Desert ecoregion is characterised by sparse cover of xeric grasses, particularly Stipagrostis, if vegetation is present at all. Here, the water table is generally too low to support even deep-rooting arboreal taxa, and only plants capable of exploiting the region's extremely sporadic rainfall occur. To the east, the uplands of the Great Escarpment and interior plateau support the low-shrub vegetation of the Gariep Karoo ecoregion. Shrubs of the Asteraceae and Aizoaceae families are most common, including Eriocephalus, Pentzia, and Ruschia, as are arid-adapted C4 grasses such as Aristida and Stipagrostis.

Material and methods

The middens considered in this study were collected and analyzed according to the techniques described by Chase et al. (Reference Chase, Scott, Meadows, Gil-Romera, Boom, Carr, Reimer, Truc, Valsecchi and Quick2012, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019). Depending on the architecture of the shelter in which a midden is found, the accumulation may have varying proportions of faecal pellets to urine (known as hyraceum). At Zizou, two midden sections (ZIZ-1-1 and ZIZ-1-3) were found in close proximity (~2 m) as part of the same midden complex. They were selected for analysis because they are composed primarily of hyraceum, and thus have superior stratigraphic integrity compared to more pellet-rich (and thus coarser) middens (Chase et al., Reference Chase, Scott, Meadows, Gil-Romera, Boom, Carr, Reimer, Truc, Valsecchi and Quick2012). Representative portions of the middens (measuring 227 mm and 145 mm in depth for ZIZ-1-1 and ZIZ-1-3, respectively) were extracted using an angle grinder and transported back to the laboratory, where the sections were cleaned using progressively finer grades of sandpaper prior to sub-sampling for radiocarbon, microfossil, and stable isotope samples.

Radiocarbon dating

Radiocarbon age determinations for the ZIZ-1-1 (n = 14) and ZIZ-1-3b (n = 13) middens were processed at the 14CHRONO Centre, Queen's University Belfast using accelerator mass spectrometry (AMS) (Table 1; Fig. 3). Samples were pre-treated with 2% HCl for one hour at room temperature to remove carbonates and dried at 60°C. They were then weighed into quartz tubes with an excess of CuO, sealed under vacuum, and combusted to CO2. The CO2 was converted to graphite on an iron catalyst using the zinc reduction method (Slota et al., Reference Slota, Jull, Linick and Toolin1987). The radiocarbon ages were corrected for isotope fractionation using the AMS measured δ13C. The ages were calibrated using SHCal20 calibration data (Hogg et al., Reference Hogg, Heaton, Hua, Palmer, Turney, Southon and Bayliss2020), and rbacon v.2.5.8 software package (Blaauw and Christen, Reference Blaauw and Christen2011, Reference Blaauw and Christen2020) was used to generate the age-depth models for each midden, employing in each case calculated mean accumulation rate of 100 yr/mm (Fig. 3).

Figure 3. Distribution in time and depth of 14C ages for the Zizou rock hyrax middens with age-depth model (grey), overlying the calibrated distributions of the individual dates calculated using Bacon v2.5.8 (Blaauw and Christen, Reference Blaauw and Christen2011). Red dotted lines indicate the ‘best’ model based on the mean age. The Holocene section of the ZIZ-1-1 age model was calculated in isolation (upper panel) to better constrain the marked post-Pleistocene change in accumulation rate.

Table 1. Radiocarbon ages and calibration information for the Zizou-1-1 and Zizou-1-3b rock hyrax middens.

Stable carbon and nitrogen isotopes

Following the previous works of Chase et al. (Reference Chase, Scott, Meadows, Gil-Romera, Boom, Carr, Reimer, Truc, Valsecchi and Quick2012, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), the stable carbon and nitrogen isotope composition of 271 hyraceum samples (~2 mg) was determined at the School of Geography, Geology and the Environment, University of Leicester, with contiguous/overlapping samples obtained from two offset tracks using a 1-mm drill. Isotope ratios were measured on a Sercon 20-20 continuous flow isotope ratio mass spectrometer. For the stable isotope analyses, the standard deviation derived from replicate analyses of homogeneous reference material (Casein protein CatNo.B2155 Batch no. 114859 Elemental Microanalysis Ltd UK) was better than 0.2‰ for both carbon and nitrogen. Carbon isotope results are expressed relative to Vienna PDB, and nitrogen isotope results are expressed relative to atmospheric nitrogen (Fig. 5).

Pollen and microcharcoal

Following the protocol of Lim et al. (Reference Lim, Chase, Chevalier and Reimer2016), pollen samples were prepared with standard physical (600-μm sieving and decanting) and chemical (HCl, KOH, HF, and acetolysis) methods (Moore et al., Reference Moore, Webb and Collinson1991). Lycopodium tablets (Batch 1031, containing ~20,848 spores/tablet) were added to the weighed sample to estimate pollen concentrations (Stockmarr, Reference Stockmarr1971). A minimum pollen sum of 400 grains was counted at a magnification of ×400 under a light microscope, and identified using (1) the literature (van Zinderen Bakker, Reference van Zinderen Bakker1953, Reference van Zinderen Bakker1956; van Zinderen Bakker and Coetzee, Reference van Zinderen Bakker and Coetzee1959; Scott, Reference Scott1982); (2) the photographic and slide reference collections of the University of the Free State, University of Cape Town, and University of Montpellier, and (3) the atlas of southern African pollen types and their climatic affinities (Chevalier et al., Reference Chevalier, Chase, Quick, Scott, Runge, Gosling, Lézine and Scott2021).

Microcharcoal particles were identified as black, completely opaque, angular fragments that occurred in the pollen slides (Clark, Reference Clark1988). Only charcoal particles >75 μm2 (or longer than 10 μm) were counted under a light microscope at ×400 magnification (Patterson et al., Reference Patterson, Edwards and Maguire1987; Mooney and Tinner, Reference Mooney and Tinner2011). A minimum count of 200 items (given by the sum of charcoal particles and exotic marker grains) was used. Charcoal particles that exceed the mesh-width size of 600 μm are missing from the microscopic charcoal record and particles of less than ~10 μm were not counted in order to ensure correct identification (Mooney and Tinner, Reference Mooney and Tinner2011). Accordingly, the charcoal signal is related primarily to the regional fire signal, specifically with local fires (large particles) and remote, extra-regional fires (<10 μm particles) being excluded. The TILIA program was used to construct the pollen diagrams, and pollen zones were determined by CONISS cluster analysis (Grimm, Reference Grimm2011).

In comparing the stable-isotope and pollen data from middens, differences in sample sizes should be considered since the stable isotope record is based on 1-mm overlapping samples, while pollen samples from the Zizou middens, limited by the brittleness and density of the hyraceum and the larger sample sizes required for pollen analysis, vary in width from 3.15–12.54 mm.

RESULTS

Chronology

The radiocarbon and age-depth modeling results suggest continuous midden accumulation despite changes in accumulation rates. Midden structure and sampling constraints have provided a stable isotope record extending from 2500–37,100 cal yr BP and pollen and microcharcoal records extending from 6000–39,000 cal yr BP (average age of microfossil samples, which measure 6 ± 2 mm in thickness). Notable in the ZIZ-1-1 midden, which extends into the Holocene, is the substantial decrease in accumulation rate, which renders the Holocene record from the middens of significantly lower resolution (two samples comprising 8100 years of accumulated hyraceum) (Fig. 3).

Stable isotopes

The δ13C and δ15N records from the ZIZ-1-1 and ZIZ-1-3b middens have each been integrated into single records based on sample age. The δ13C values of the Zizou middens vary from −18.6 to −24.3‰ (Fig. 5), with the highest values occurring during the Holocene, while the last glacial period is generally characterised by lower values. The δ15N values range from 7.2–13.6‰ (Fig. 5) with Holocene and glacial periods being characterised by generally higher (mean 12.1‰) and lower (mean 10.9‰ values) respectively. The δ13C and δ15N records exhibit a degree of covariance at centennial–millennial timescales, with higher δ13C values occurring, broadly, when δ15N values are elevated. The lowest δ13C and δ15N values occur around 30,000 cal yr BP and at the end of the Last Glacial Maximum (LGM)/earliest glacial-interglacial transition, ca. 16,000–18,000 cal yr BP, with relatively higher values during the LGM (especially 20,000–23,000 cal yr BP) and most notably during the Holocene after 7000 cal yr BP.

Pollen and microcharcoal

As with the stable isotope records, the pollen and microcharcoal records from the ZIZ-1-1 and ZIZ-1-3b middens were integrated into single records based on sample age. Fifty pollen taxa were identified, which then were clustered into four statistically significant pollen assemblage zones (ZIZ 1–ZIZ 4) spanning much of the last 39,000 years (Fig. 4). Poaceae, Asteraceae, Capparaceae, and Zygophyllum are the most common pollen types, with much greater pollen diversity in the Pleistocene. The Holocene pollen spectra are dominated by grass (Poaceae) pollen. This Pleistocene–Holocene dichotomy is less evident in the microcharcoal record, but charcoal concentrations are highest in Pleistocene zones 1 and 3.

Figure 4. Diagram of pollen and microcharcoal from the Zizou rock hyrax middens, including taxa > 2% of composition. Mean ages for samples, including minimum and maximum depths, are provided.

Figure 5. δ15N, δ13C, microcharcoal concentration, and select pollen data from the Zizou rock hyrax middens illustrated along with the Namib composite δ15N record from Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019, and high latitude boreal and low latitude austral summer insolation (Laskar et al., Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). Holocene, Last Glacial Maximum (LGM), and Heinrich stadial (HS1, HS2, HS3; grey bars) chronozones are indicated.

The lowermost pollen Zone (Zone 1; ca. 39,000 to ca. 35,800 cal yr BP) is characterised by early peaks in Searsia-type and Moraceae pollen among the arboreal taxa but is primarily defined by elevated levels of Capparaceae and Zygophyllum pollen (Fig. 4). Asteraceae pollen is relatively rare (average 10%) and Poaceae pollen is at its lowest levels in the sequence (average 31%).

Pollen Zone 2 (ca. 35,800 to ca. 23,900 cal yr BP) is defined by a strong increase in Asteraceae pollen (average 22%) at the expense of arboreal taxa. Poaceae pollen increases slightly (average 36%) but fluctuates around this mean value. Microcharcoal concentrations decrease slightly in Zone 2, from averages of 913 to 704 fragments/gram.

Subsequently, Pollen Zone 3 (ca. 23,900 to ca. 18,000 cal yr BP) sees a strong increase in microcharcoal, while Asteraceae pollen, particularly Felicia, declines. Zygophyllum pollen also declines in this zone, while Capparaceae again becomes a dominant element in the pollen assemblage (average 15%). Poaceae pollen increases again (average 42%) and stabilises around this mean.

Pollen Zone 4 (ca. 18,000 to ca. 5300 cal yr BP) reflects a fundamentally different vegetation that the Pleistocene assemblages. Poaceae pollen is totally dominant, representing 92% of the pollen recovered, with only very low percentages of drought-resistant taxa such as Aizoaceae, Amaranthaceae-type, and Myrothamnus comprising the other apparent elements. Asteraceae and arboreal pollen are found in trace amounts (<2%).

DISCUSSION

Combining stable carbon and nitrogen data with pollen and microcharcoal data from the same archives provides the opportunity to evaluate vegetation and fire dynamics within the context of an independent climate record derived from midden stable isotope content. As described by Chase et al. (Reference Chase, Scott, Meadows, Gil-Romera, Boom, Carr, Reimer, Truc, Valsecchi and Quick2012, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), rock hyrax δ15N records are interpreted as a proxy for water availability, with higher δ15N indicating drier conditions. As described in several previous studies, and supported by strong correlations with a diverse range of proxy climate records (Chase et al., Reference Chase, Meadows, Scott, Thomas, Marais, Sealy and Reimer2009, Reference Chase, Lim, Chevalier, Boom, Carr, Meadows and Reimer2015a, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), drier conditions induce more N flow to inorganic soil nitrogen pools that are subject to gaseous loss of 15N-depleted products (Austin and Vitousek, Reference Austin and Vitousek1998; Handley et al., Reference Handley, Austin, Stewart, Robinson, Scrimgeour, Raven, Heaton and Schmidt1999; Murphy and Bowman, Reference Murphy and Bowman2009). Plants growing in the resulting 15N-enriched soils exhibit higher δ15N values (Craine et al., Reference Craine, Elmore, Aidar, Bustamante, Dawson, Hobbie and Kahmen2009; Hartman and Danin, Reference Hartman and Danin2010), a signal that is transmitted to the tissues (and excreta) of animals that consume them (Murphy and Bowman, Reference Murphy and Bowman2006; Hartman, Reference Hartman2011; Carr et al., Reference Carr, Chase, Boom and Medina-Sanchez2016).

Hyrax midden δ13C reflects animal diet and, in turn, two aspects of the response of vegetation to climate change: (1) plant photosynthetic pathway of C3 (approximately −34 to −24‰), C4 (approximately −16 to −10‰), and CAM (typically within the range −20 to −10‰) plants (Smith and Epstein, Reference Smith and Epstein1971; Smith, Reference Smith1972; Werger and Ellis, Reference Werger and Ellis1981; Rundel et al., Reference Rundel, Esler and Cowling1999; Boom et al., Reference Boom, Carr, Chase, Grimes and Meadows2014); and (2) changes in C3 plant water-use efficiency (Farquhar and Richards, Reference Farquhar and Richards1984; Ehleringer and Cooper, Reference Ehleringer and Cooper1988; Farquhar et al., Reference Farquhar, Hubick, Condon, Richards, Rundel, Ehleringer and Nagy1989a). Pollen data from Zizou indicate that C3 vegetation in the form of trees and shrubs, was at times abundant during the Pleistocene. Today, C4 grasses are the most dominant vegetation, and these data show this has likely been the case throughout much of the Holocene. Although it is possible that C3 grasses occurred at the site during the Pleistocene, midden δ13C values suggest that C4 grasses were still present. CAM plants also have been abundant at the site (Fig. 5), and as generalist feeders (Hoeck, Reference Hoeck1975; Fourie, Reference Fourie1990) these may have formed part of the hyraxes’ diet, although perhaps not in significant quantities (Fourie, Reference Fourie1983; Carr et al., Reference Carr, Boom and Chase2010, Reference Carr, Chase, Boom and Medina-Sanchez2016).

The Zizou δ15N record (Fig. 5) exhibits a pattern of climate variability that is similar to that observed across the Namib Desert region (Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019). Locally and regionally, these data indicate that the last glacial period was generally more humid than in the Late Holocene. Considering data from a 900 km north-south transect of hyrax middens from the Namib Desert, Chase et al. (Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019) ascribed these changes in humidity to: (1) changes in high latitude boreal and low latitude austral summer insolation, which operate in concert to modify land-sea pressure gradients and the advection of cold air from the Benguela Upwelling System; and 2) mediation of the response to these orbital parameters by Heinrich events and related stadials, which warm the SE Atlantic and result in more humid conditions in the Namib region.

The observed changes in vegetation are largely consistent with changes in climate inferred from the midden δ15N record. Considered within the context of regional ecosystems, wetter conditions may be expected to favor the development of vegetation types similar to either the Angolan Mopane Woodlands or the Gariep Karoo (Fig. 2). Notably, Colophospermum (Mopane) pollen was not recorded in the Zizou middens, while the proportions of pollen from Asteraceae and succulent taxa, the primary plant types of the Gariep and Nama-Karoo ecoregions (sensu Dinerstein et al., Reference Dinerstein, Olson, Joshi, Vynne, Burgess, Wikramanayake and Hahn2017), dominated the glacial-age pollen spectra (pollen zones 1–3) (Fig. 4). Pollen zone 1 (ca. 39,000 to ca. 32,900 cal yr BP), pre-dates the phase of peak humidity from 30,700–28,800 cal yr BP and has an assemblage suggesting a dry karroid vegetation dominated by succulents, with arboreal taxa also being well represented. As conditions became wetter, pollen zone 2 (ca. 32,900 to ca. 22,500 cal yr BP) exhibits a marked decline in succulent pollen, concurrent with peak humidity, and a clear increase in Asteraceae pollen. We interpret these changes as indicating a shift to more mesic Karoo vegetation, comprising grasses and asteraceous shrubs, with a reduced succulent component. Based on the climatic affinities of vegetation types in the region today (Fig. 2), humidity may have been 50–100% higher than the modern situation (e.g., aridity index values increasing from ~0.08 to 0.12–0.16).

A similar sequence of vegetation changes is observed in hyrax middens at Mirabib (Scott et al., Reference Scott, Gil Romera, Marais and Brook2018), on the now grass-dominated plains of the Namib Desert to the north of the Kuiseb River (Fig. 1). As at Zizou, MIS-2–MIS-3 pollen assemblages indicate much higher percentages of Asteraceae, Aizoaceae, and arboreal taxa (particularly Searsia, Euclea, and Capparaceae). These changes in vegetation observed in both the eastern and central Namib Desert, along with the stable isotope data presented here and elsewhere (Chase et al., Reference Chase, Meadows, Scott, Thomas, Marais, Sealy and Reimer2009, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), modify the conclusion of Lancaster's (Reference Lancaster2002) review of Namib Desert paleoclimates, in which it was suggested that there was little evidence for significantly increased rainfall in the Namib Desert during the last glacial period. Our findings are not inconsistent with the view that conditions during this time remained “arid,” since even if humidity at Zizou was doubled, it would still leave the region below the threshold for “semi-arid” conditions (aridity index values of 0.2–0.5; UNEP, 1997). A shift towards drier conditions after ca. 29,000 cal yr BP saw a decline in Asteraceae pollen and an increase in succulent pollen, indicating drier karroid vegetation (Fig. 5).

Pollen zone 3 (ca. 22,500 to ca. 17,700 cal yr BP) spans much of the LGM, which is marked by an abrupt decrease in humidity at Zizou at ca. 24,000 cal yr BP. Vegetation responded to this aridification with a decline in Asteraceae pollen and notable increase in the abundance of Poaceae and arboreal pollen. We interpret this as reflecting a lowering water table, and a change in relative competitive advantage associated with plant rooting depth. As conditions became drier, plant-available water in the near-surface sediments would have been reduced, limiting the occurrence of shallower rooting shrubs. However, arboreal taxa, with their deeper root systems, were able to persist through accessing groundwater resources, while grasses were able to thrive, as they do in the present day, by exploiting superficial soil moisture following occasional rainfall events.

Pollen zone 4 (ca. 17,700 to ca. 5300 cal yr BP) represents the most dramatic shift in the vegetation at Zizou. The regional Namib climate record (Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019) indicates that conditions became more humid after ca. 18,000 cal yr BP, culminating in a period of particularly wet conditions from ca. 15,100–14,400 cal yr BP. This is consistent with a slowing of Atlantic Meridional Overturning Circulation and progressive build-up of heat in the South Atlantic associated with Heinrich stadial 1 (Broecker, Reference Broecker1998; Stocker, Reference Stocker1998; Stocker and Johnsen, Reference Stocker and Johnsen2003; Chase et al., Reference Chase, Boom, Carr, Carré, Chevalier, Meadows, Pedro, Stager and Reimer2015b, Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019). The pollen record from Zizou is of insufficient temporal resolution to address this point but does suggest that after the LGM karroid vegetation and arboreal taxa were largely absent, and opportunistic grasses were the only plants able to persist under the warm, dry conditions that became established. This is inferred to reflect lowering of groundwater resources and dominance of growth forms that were capable of exploiting limited and sporadic rainfall. The better-resolved δ15N values from the Zizou middens vary consistently with other regional records (Stute and Talma, Reference Stute and Talma1998; Chase et al., Reference Chase, Meadows, Scott, Thomas, Marais, Sealy and Reimer2009, Reference Chase, Meadows, Carr and Reimer2010; Carr et al., Reference Carr, Boom and Chase2010) that suggest a relatively mesic Early Holocene. The decreased midden accumulation rate and the pollen assemblages suggest a decrease in plant biomass and a decline in the population of the hyrax colony at the site during the Holocene. In the absence of any significantly better-resolved Holocene pollen sequences from the western slope of the Escarpment to facilitate comparison (Scott, Reference Scott1996; Scott et al., Reference Scott, Gil-Romera, Marais and Brook2022), the samples from Zizou presently can provide only an initial indication of vegetation at the site in the post-glacial period.

Throughout the record, changes in vegetation are reflected in the midden δ13C values, which track variations in Poaceae (predominantly C4) and to a lesser degree succulent (predominantly CAM) pollen percentages (Figs 4, 5), supporting the evidence that woody/arboreal (C3) vegetation was dominant during a more humid last glacial period. Together with the observation that grass dominated the Holocene vegetation, our observations indicate a limited role of atmospheric CO2 concentrations as a driver of glacial–interglacial vegetation dynamics in the Namib Desert. This differs from observations and interpretations in other contexts, where atmospheric CO2 concentration is thought to play a central role in vegetation change (Bond and Midgley, Reference Bond and Midgley2000; Bragg et al., Reference Bragg, Prentice, Harrison, Eglinton, Foster, Rommerskirchen and Rullkötter2013; Prentice et al., Reference Prentice, Cleator, Huang, Harrison and Roulstone2017; Dupont et al., Reference Dupont, Caley and Castañeda2019) through the competitive advantage that C3 woody plants have under augmented CO2 conditions (Farquhar et al., Reference Farquhar, Hubick, Condon, Richards, Rundel, Ehleringer and Nagy1989a, Reference Farquhar, Ehleringer and Hubickb; Bond and Midgley, Reference Bond and Midgley2000; Donohue et al., Reference Donohue, Roderick, McVicar and Farquhar2013). The evidence here suggests that Holocene high-CO2 conditions were accompanied by the expansion of C4 grasses and a virtual absence of woody C3 taxa. This pattern of vegetation change has been observed elsewhere, for example in the southern Namib at Pella (Lim et al., Reference Lim, Chase, Chevalier and Reimer2016), and in other dryland regions such as the North American southwest (Holmgren et al., Reference Holmgren, Betancourt, Peñalba, Delgadillo, Zuravnsky, Hunter, Rylander and Weiss2014). In summary, this evidence suggests that, while variation in CO2 may influence C3/C4 vegetation dynamics, it is not necessarily the dominant driver in some contexts over glacial-interglacial timescales (see Nolan et al., Reference Nolan, Overpeck, Allen, Anderson, Betancourt, Binney and Brewer2018).

The microcharcoal content of the Zizou middens facilitates consideration of the region's fire history and ecology. Generally speaking, arid and semi-arid regions are considered to have fire regimes that are biomass limited (Marlon et al., Reference Marlon, Bartlein, Daniau, Harrison, Maezumi, Power, Tinner and Vanniére2013). As a result, charcoal concentrations may be expected to increase in these environments under wetter conditions, with denser vegetation providing more biomass, allowing fire to travel more effectively and thereby producing more charcoal. At Zizou, the relationships between climate, vegetation, and fire indicate that this nexus may be more complex. During the last glacial period, phases of relatively wet conditions are shown here to be associated with reduced microcharcoal concentrations (Fig. 5), an apparent anomaly that may be explained by the relative ability of different vegetation types to carry fire due to spacing of individual plants. For example, the relatively wide spacing between plants in karroid vegetation limits its potential to carry fire (Mucina et al., Reference Mucina, Rutherford, Palmer, Milton, Scott, Lloyd, van der Merwe, Mucina and Rutherford2006; Hebbelmann et al., Reference Hebbelmann, O'Connor and du Toit2022). At Zizou, increased charcoal concentrations correspond to phases with higher percentages of grass pollen, implying either that grass expansion increased charcoal concentrations as a function of increased biomass, or that it may play an important role in carrying fire in such environments (Rahlao et al., Reference Rahlao, Milton, Esler, Van Wilgen and Barnard2009). Elevated percentages of arboreal pollen during these phases likely indicates the presence of an important fuel source, further augmenting charcoal concentrations in what may have been more extensive fires. In contrast, the terminal Pleistocene shift to sparse xeric grassland was accompanied by a decline in microcharcoal concentrations, conforming to predictions based on a marked decreased in biomass.

The Zizou pollen record provides further confirmation of the concerns raised by Scott et al. (Reference Scott, Marais and Brook2004) regarding the interpretation of pollen and other terrestrial material (e.g., microcharcoal and leaf waxes) recovered from marine cores in the Benguela Upwelling System. For example, whereas abundant Restionaceae pollen is evident in regional marine cores during MIS 2–3, including evidence used to infer significant expansions of the Fynbos Biome and winter rainfall zone (Shi et al., Reference Shi, Dupont, Beug and Schneider2000, Reference Shi, Schneider, Beug and Dupont2001; Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015), no similar evidence has been recovered from the Namib Desert (Scott et al., Reference Scott, Marais and Brook2004, Reference Scott, Gil Romera, Marais and Brook2018; Lim et al., Reference Lim, Chase, Chevalier and Reimer2016; this paper).

Modeling studies of wind trajectories indicate the primary source area for terrestrial components in marine cores GeoB 1710 (Dupont and Wyputta, Reference Dupont and Wyputta2003), GeoB 1712, (Shi et al., Reference Shi, Schneider, Beug and Dupont2001), MD96-2098 (Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015), and MD08-3016 (Collins et al., Reference Collins, Schefuß, Govin, Mulitza and Tiedemann2014) lies directly to the east of the core sites (Dupont and Wyputta, Reference Dupont and Wyputta2003). This observation, combined with finding that aeolian transport trajectories deviated little during the LGM (Dupont and Wyputta, Reference Dupont and Wyputta2003), has been used to support the use of marine records as evidence of environmental change on adjacent continental regions (Dupont and Wyputta, Reference Dupont and Wyputta2003; Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015). However, there are no terrestrial records from this putative source region that support this interpretation (Scott et al., Reference Scott, Marais and Brook2004, Reference Scott, Gil Romera, Marais and Brook2018; Lim et al., Reference Lim, Chase, Chevalier and Reimer2016; this paper). Records of terrestrial material preserved in marine sediments from this region, for example pollen (Urrego et al., Reference Urrego, Sánchez Goñi, Daniau, Lechevrel and Hanquiez2015), leaf waxes (Collins et al., Reference Collins, Schefuß, Govin, Mulitza and Tiedemann2014), or charcoal (Daniau et al., Reference Daniau, Sánchez Goñi, Martinez, Urrego, Bout-Roumazeilles, Desprat and Marlon2013), require further confirmation of their possible sources. Without such information, reliable reconstructions of the adjacent terrestrial environments are questionable. Possible alternative hypotheses to explain the anomaly between marine and terrestrial records, such as changes in source area driven by demonstrated changes in wind strength (Chase et al., Reference Chase, Niedermeyer, Boom, Carr, Chevalier, He, Meadows, Ogle and Reimer2019), warrant further consideration.

CONCLUSIONS

  • The stable carbon and nitrogen isotope, pollen, and microcharcoal data from the Zizou rock hyrax middens provide evidence for significant differences between MIS 2–3 and Holocene climates and vegetation types in the central Namib Desert.

  • While Pleistocene conditions were likely still arid to semi-arid, they were significantly more humid than in the later Holocene.

  • During MIS 2–3, shrubby karroid vegetation developed under the most humid conditions, while deeper-rooting arboreal taxa were relatively more abundant under intermediate conditions, in a response to lowering water table. The modern xeric grass-dominant vegetation, an opportunistic vegetation that is capable of exploiting sporadic rains, was established after the last glacial period.

  • The Zizou microcharcoal record is most strongly correlated with percentages of grass and arboreal pollen, rather than with humidity. The karroid vegetation that characterises the most humid periods is not conducive to extensive fires because inter-plant spacing does not carry fire well. Under drier conditions, a better-developed grass component carries fire more effectively and the persistence of a significant arboreal component provides fuel, raising charcoal concentrations. This fuel source apparently was absent during the period reflected in our Holocene samples, and charcoal concentrations declined.

  • Evolution of the vegetation composition at Zizou does not indicate that changes in atmospheric CO2 played a dominant role in determining plant functional type. The dominance of grasses (primarily C4) during the Holocene and shrubs and arboreal (C3) taxa during the Pleistocene indicates that climate is the primary driver of vegetation composition.

  • As at other terrestrial sites from the Namib Desert, no elements of the Cape flora (e.g., Restionaceae) were found in the Zizou pollen assemblages. These results differ markedly from pollen records obtained from marine sediment cores, requiring a fundamental reconsideration of interpretations of environmental change in the region based on marine records.

Acknowledgments

Funding was received from the European Research Council (ERC) under the European Union's Seventh Framework Programme (FP7/2007-2013)/ERC Starting Grant “HYRAX”, grant agreement no. 258657, and from the Leverhulme Trust.

References

REFERENCES

Austin, A.T., Vitousek, P.M., 1998. Nutrient dynamics on a precipitation gradient in Hawai'i. Oecologia 113, 519529.CrossRefGoogle ScholarPubMed
Beaumont, P.B., 1986. Where did all the young men go during 18O stage 2? Palaeoecology of Africa 17, 7986.Google Scholar
Blaauw, M., Christen, J.A., 2011. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Analysis 6, 457474.CrossRefGoogle Scholar
Blaauw, M., Christen, J.A., 2020. Rbacon: Age-Depth Modelling Using Bayesian Statistics. R package version 2.5.6. https://CRAN.R-project.org/package=rbacon.Google Scholar
Bond, W.J., Midgley, G.F., 2000. A proposed CO2-controlled mechanism of woody plant invasion in grasslands and savannas. Global Change Biology 6, 865869.CrossRefGoogle Scholar
Boom, A., Carr, A.S., Chase, B.M., Grimes, H.L., Meadows, M.E., 2014. Leaf wax n-alkanes and δ13C values of CAM plants from arid southwest Africa. Organic Geochemistry 67, 99102.CrossRefGoogle Scholar
Bragg, F.J., Prentice, I.C., Harrison, S.P., Eglinton, G., Foster, P.N., Rommerskirchen, F., Rullkötter, J., 2013. Stable isotope and modelling evidence for CO2 as a driver of glacial–interglacial vegetation shifts in southern Africa. Biogeosciences 10, 20012010.CrossRefGoogle Scholar
Broecker, W.S., 1998. Paleocean circulation during the last deglaciation: a bipolar seesaw? Paleoceanography 13, 119121.CrossRefGoogle Scholar
Carr, A.S., Boom, A., Chase, B.M., 2010. The potential of plant biomarker evidence derived from rock hyrax middens as an indicator of palaeoenvironmental change. Palaeogeography, Palaeoclimatology, Palaeoecology 285, 321330.CrossRefGoogle Scholar
Carr, A.S., Chase, B.M., Boom, A., Medina-Sanchez, J., 2016. Stable isotope analyses of rock hyrax faecal pellets, hyraceum and associated vegetation in southern Africa: implications for dietary ecology and palaeoenvironmental reconstructions. Journal of Arid Environments 134, 3348.CrossRefGoogle Scholar
Chase, B.M., Boom, A., Carr, A.S., Carré, M., Chevalier, M., Meadows, M.E., Pedro, J.B., Stager, J.C., Reimer, P.J., 2015b. Evolving southwest African response to abrupt deglacial North Atlantic climate change events. Quaternary Science Reviews 121, 132136.CrossRefGoogle Scholar
Chase, B.M., Lim, S., Chevalier, M., Boom, A., Carr, A.S., Meadows, M.E., Reimer, P.J., 2015a. Influence of tropical easterlies in southern Africa's winter rainfall zone during the Holocene. Quaternary Science Reviews 107, 138148.CrossRefGoogle Scholar
Chase, B.M., Meadows, M.E., 2007. Late Quaternary dynamics of southern Africa's winter rainfall zone. Earth-Science Reviews 84, 103138.CrossRefGoogle Scholar
Chase, B.M., Meadows, M.E., Carr, A.S., Reimer, P.J., 2010. Evidence for progressive Holocene aridification in southern Africa recorded in Namibian hyrax middens: implications for African Monsoon dynamics and the “African Humid Period”. Quaternary Research 74, 3645.CrossRefGoogle Scholar
Chase, B.M., Meadows, M.E., Scott, L., Thomas, D.S.G., Marais, E., Sealy, J., Reimer, P.J., 2009. A record of rapid Holocene climate change preserved in hyrax middens from southwestern Africa. Geology 37, 703706.CrossRefGoogle Scholar
Chase, B.M., Niedermeyer, E.M., Boom, A., Carr, A.S., Chevalier, M., He, F., Meadows, M.E., Ogle, N., Reimer, P.J., 2019. Orbital controls on Namib Desert hydroclimate over the past 50,000 years. Geology 47, 867871.CrossRefGoogle Scholar
Chase, B.M., Scott, L., Meadows, M.E., Gil-Romera, G., Boom, A., Carr, A.S., Reimer, P.J., Truc, L., Valsecchi, V., Quick, L.J., 2012. Rock hyrax middens: a palaeoenvironmental archive for southern African drylands. Quaternary Science Reviews 56, 107125.CrossRefGoogle Scholar
Chevalier, M., Chase, B.M., Quick, L.J., Scott, L., 2021. An atlas of southern African pollen types and their climatic affinities. In: Runge, J., Gosling, W.D., Lézine, A.-M., Scott, L. (Eds.), Quaternary Vegetation Dynamics – The African Pollen Database. CRC Press, London, 442 pp.Google Scholar
Clark, J.S., 1988. Stratigraphic charcoal analysis on petrographic thin section: application to fire history in northwestern Minnesota. Quaternary Research 30, 8191.CrossRefGoogle Scholar
Collins, J.A., Schefuß, E., Govin, A., Mulitza, S., Tiedemann, R., 2014. Insolation and glacial–interglacial control on southwestern African hydroclimate over the past 140 000 years. Earth and Planetary Science Letters 398, 110.CrossRefGoogle Scholar
Craine, J.M., Elmore, A.J., Aidar, M.P.M., Bustamante, M., Dawson, T.E., Hobbie, E.A., Kahmen, A., et al., 2009. Global patterns of foliar nitrogen isotopes and their relationships with climate, mycorrhizal fungi, foliar nutrient concentrations, and nitrogen availability. New Phytologist 183, 980992.CrossRefGoogle ScholarPubMed
Daniau, A.-L., Sánchez Goñi, M.F., Martinez, P., Urrego, D.H., Bout-Roumazeilles, V., Desprat, S., Marlon, J.R., 2013. Orbital-scale climate forcing of grassland burning in southern Africa. Proceedings of the National Academy of Sciences 110, 50695073.CrossRefGoogle ScholarPubMed
Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N.D., Wikramanayake, E., Hahn, N., et al., 2017. An ecoregion-based approach to protecting half the terrestrial realm. BioScience 67, 534545.CrossRefGoogle ScholarPubMed
Donohue, R.J., Roderick, M.L., McVicar, T.R., Farquhar, G.D., 2013. Impact of CO2 fertilization on maximum foliage cover across the globe's warm, arid environments. Geophysical Research Letters 40, 30313035.CrossRefGoogle Scholar
Dupont, L.M., Caley, T., Castañeda, I.S., 2019. Effects of atmospheric CO2 variability of the past 800 kyr on the biomes of southeast Africa. Climate of the Past 15, 10831097.CrossRefGoogle Scholar
Dupont, L.M., Wyputta, U., 2003. Reconstructing pathways of aeolian pollen transport to the marine sediments along the coastline of SW Africa. Quaternary Science Reviews 22, 157174.CrossRefGoogle Scholar
Ehleringer, J.R., Cooper, T.A., 1988. Correlations between carbon isotope ratio and microhabitat of desert plants. Oecologia 76, 562566.CrossRefGoogle ScholarPubMed
Etourneau, J., Martinez, P., Blanz, T., Schneider, R., 2009. Pliocene–Pleistocene variability of upwelling activity, productivity, and nutrient cycling in the Benguela region. Geology 37, 871874.CrossRefGoogle Scholar
Farmer, E.C., deMenocal, P.B., Marchitto, T.M., 2005. Holocene and deglacial ocean temperature variability in the Benguela upwelling region: implications for low-latitude atmospheric circulation. Paleoceanography 20, PA2018. https://doi.org/10.1029/2004PA001049.CrossRefGoogle Scholar
Farquhar, G.D., Ehleringer, J.R., Hubick, K.T., 1989b. Carbon isotope discrimination and photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology 40, 503537.CrossRefGoogle Scholar
Farquhar, G.D., Hubick, K.T., Condon, A.G., Richards, R.A., 1989a. Carbon isotope fractionation and plant water-use efficiency. In: Rundel, P.W., Ehleringer, J.R., Nagy, K.A. (Eds.), Stable Isotopes in Ecological Research. Springer, New York, pp. 2140.CrossRefGoogle Scholar
Farquhar, G., Richards, R., 1984. Isotopic composition of plant carbon correlates with water-use efficiency of wheat genotypes. Functional Plant Biology 11, 539552.CrossRefGoogle Scholar
Fick, S.E., Hijmans, R.J., 2017. WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. International Journal of Climatology 37, 43024315.CrossRefGoogle Scholar
Fourie, L.J., 1983. The population dynamics of the rock hyrax Procavia capensis (Pallas, 1766) in the Mountain Zebra National Park. PhD dissertation, Rhodes University, Grahamstown, South Africa, 397 pp.Google Scholar
Fourie, L.J., 1990. Quantitative and qualitative aspects of the diet of the rock hyrax (Procavia capensis Pallas, 1766) in the Mountain Zebra National Park. Revue de Zoologie Africaine 103, 171177.Google Scholar
Gil-Romera, G., Scott, L., Marais, E., Brook, G.A., 2006. Middle- to Late-Holocene moisture changes in the desert of northwest Namibia derived from fossil hyrax dung pollen. The Holocene 16, 10731084.CrossRefGoogle Scholar
Gil-Romera, G., Scott, L., Marais, E., Brook, G.A., 2007. Late Holocene environmental change in the northwestern Namib Desert margin: new fossil pollen evidence from hyrax middens. Palaeogeography, Palaeoclimatology, Palaeoecology 249, 117.CrossRefGoogle Scholar
Grimm, E.C., 2011. Tilia software version 1.7.16. Illinois State Museum, Research and Collection Center, Springfield, Illinois. Available: http://intra.museum.state.il.us/pub/grimm/tilia/.Google Scholar
Handley, L.L., Austin, A.T., Stewart, G.R., Robinson, D., Scrimgeour, C.M., Raven, J.A., Heaton, T.H.E., Schmidt, S., 1999. The 15N natural abundance (∂15N) of ecosystem samples reflects measures of water availability. Functional Plant Biology 26, 185199.CrossRefGoogle Scholar
Hartman, G., 2011. Are elevated δ15N values in herbivores in hot and arid environments caused by diet or animal physiology? Functional Ecology 25, 122131.CrossRefGoogle Scholar
Hartman, G., Danin, A., 2010. Isotopic values of plants in relation to water availability in the Eastern Mediterranean region. Oecologia 162, 837852.CrossRefGoogle ScholarPubMed
Hebbelmann, L., O'Connor, T.G., du Toit, J.C.O., 2022. Fire as a novel disturbance and driver of vegetation change in Nama-Karoo rangelands, South Africa. Journal of Arid Environments 203, 104777. https://doi.org/10.1016/j.jaridenv.2022.104777.CrossRefGoogle Scholar
Heine, K., 1998. Climate change over the past 135 000 years in the Namib Desert (Namibia) derived from proxy data. Palaeoecology of Africa 25, 171198.Google Scholar
Heine, K., Heine, J.T., 2002. A paleohydrologic reinterpretation of the Homeb Silts, Kuiseb River, central Namib Desert (Namibia) and paleoclimatic implications. Catena 48, 107130.CrossRefGoogle Scholar
Hoeck, H.N., 1975. Differential feeding behaviour of the sympatric hyrax: (Procavia johnstoni and Heterohyrax brucei). Oecologia 22, 1547.CrossRefGoogle ScholarPubMed
Hogg, A.G., Heaton, T.J., Hua, Q., Palmer, J.G., Turney, C.S.M., Southon, J., Bayliss, A., et al., 2020. SHCal20 Southern Hemisphere Calibration, 0–55,000 years cal BP. Radiocarbon 62, 759778.CrossRefGoogle Scholar
Holmgren, C.A., Betancourt, J.L., Peñalba, M.C., Delgadillo, J., Zuravnsky, K., Hunter, K.L., Rylander, K.A., Weiss, J.L., 2014. Evidence against a Pleistocene desert refugium in the Lower Colorado River Basin. Journal of Biogeography 41, 17691780.CrossRefGoogle Scholar
Lancaster, N., 1984. Paleoenvironments in the Tsondab Valley, central Namib Desert. Palaeoecology of Africa 16, 411419.Google Scholar
Lancaster, N., 2002. How dry was dry?—Late Pleistocene palaeoclimates in the Namib Desert. Quaternary Science Reviews 21, 769782.CrossRefGoogle Scholar
Laskar, J., Robutel, P., Joutel, F., Gastineau, M., Correia, A.C.M., Levrard, B., 2004. A long-term numerical solution for the insolation quantities of the Earth. Astronomy & Astrophysics 428, 261285.CrossRefGoogle Scholar
Lim, S., Chase, B.M., Chevalier, M., Reimer, P.J., 2016. 50,000 years of vegetation and climate change in the southern Namib Desert, Pella, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 451, 197209.CrossRefGoogle Scholar
Little, M.G., Schneider, R.R., Kroon, D., Price, B., Bickert, T., Wefer, G., 1997. Rapid palaeoceanographic changes in the Benguela Upwelling System for the last 160,000 years as indicated by abundances of planktonic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology 130, 135161.CrossRefGoogle Scholar
Marker, M.E., Müller, J., 1978. Relict vlei silts of the Kuiseb River valley, South West Africa. Madoqua 11, 151162.Google Scholar
Marlon, J.R., Bartlein, P.J., Daniau, A.-L., Harrison, S.P., Maezumi, S.Y., Power, M.J., Tinner, W., Vanniére, B., 2013. Global biomass burning: a synthesis and review of Holocene paleofire records and their controls. Quaternary Science Reviews 65, 525.CrossRefGoogle Scholar
Mooney, S.D., Tinner, W., 2011. The analysis of charcoal in peat and organic sediments. Mires and Peat 7, 9. http://www.mires-and-peat.net/pages/volumes/map07/map0709.php.Google Scholar
Moore, P.D., Webb, J.A., Collinson, M.E., 1991. Pollen Analysis. Blackwell Scientific Publications, London.Google Scholar
Mucina, L., Rutherford, M.C., Palmer, A.R., Milton, S.J., Scott, L., Lloyd, J.W., van der Merwe, B., et al., 2006. Nama-Karoo Biome. In: Mucina, L., Rutherford, M.C. (Eds.), The Vegetation of South Africa, Lesotho and Swaziland. South African National Biodiversity Institute, Pretoria, pp. 325347.Google Scholar
Murphy, B.P., Bowman, D.M.J.S., 2006. Kangaroo metabolism does not cause the relationship between bone collagen ∂15N and water availability. Functional Ecology 20, 10621069.CrossRefGoogle Scholar
Murphy, B.P., Bowman, D.M.J.S., 2009. The carbon and nitrogen isotope composition of Australian grasses in relation to climate. Functional Ecology 23, 10401049.CrossRefGoogle Scholar
Nolan, C., Overpeck, J.T., Allen, J.R.M., Anderson, P.M., Betancourt, J.L., Binney, H.A., Brewer, S., et al., 2018. Past and future global transformation of terrestrial ecosystems under climate change. Science 361, 920923.CrossRefGoogle ScholarPubMed
Patterson, W.A., Edwards, K.J., Maguire, D.J., 1987. Microscopic charcoal as a fossil indicator of fire. Quaternary Science Reviews 6, 323.CrossRefGoogle Scholar
Pichevin, L., Cremer, M., Giraudeau, J., Bertrand, P., 2005. A 190 kyr record of lithogenic grain-size on the Namibian slope: forging a tight link between past wind-strength and coastal upwelling dynamics. Marine Geology 218, 8196.CrossRefGoogle Scholar
Prentice, I.C., Cleator, S.F., Huang, Y.H., Harrison, S.P., Roulstone, I., 2017. Reconstructing ice-age palaeoclimates: quantifying low-CO2 effects on plants. Global and Planetary Change 149, 166176.CrossRefGoogle Scholar
Rahlao, S.J., Milton, S.J., Esler, K.J., Van Wilgen, B.W., Barnard, P., 2009. Effects of invasion of fire-free arid shrublands by a fire-promoting invasive alien grass (Pennisetum setaceum) in South Africa. Austral Ecology 34, 920928.CrossRefGoogle Scholar
Rundel, P.W., Esler, K.J., Cowling, R.M., 1999. Ecological and phylogenetic patterns of carbon isotope discrimination in the winter-rainfall flora of the Richtersveld, South Africa. Plant Ecology 142, 133148.CrossRefGoogle Scholar
Sale, J.B., 1965. The feeding behaviour of rock hyraces (genera Procavia and Heterohyrax) in Kenya. African Journal of Ecology 3, 118.CrossRefGoogle Scholar
Scott, L., 1982. Late Quaternary fossil pollen grains from the Transvaal, South Africa. Review of Palaeobotany and Palynology 36, 241278.CrossRefGoogle Scholar
Scott, L., 1996. Palynology of hyrax middens: 2000 years of palaeoenvironmental history in Namibia. Quaternary International 33, 7379.CrossRefGoogle Scholar
Scott, L., Cooremans, B., de Wet, J.S., Vogel, J.C., 1991. Holocene environmental changes in Namibia inferred from pollen analysis of swamp and lake deposits. The Holocene 1, 813.CrossRefGoogle Scholar
Scott, L., Gil-Romera, G., Marais, E., Brook, G.A., 2022. Holocene environmental change along the central Namib Desert escarpment derived from hyrax and owl dung. Review of Palaeobotany and Palynology 305, 104746. https://doi.org/10.1016/j.revpalbo.2022.104746.CrossRefGoogle Scholar
Scott, L., Gil Romera, G., Marais, E., Brook, G.A., 2018. Pollen in fossil hyrax dung from Marine Isotope Stages 2 and 3 reveals past environments in Namibia. Quaternary International 464, 260272.CrossRefGoogle Scholar
Scott, L., Marais, E., Brook, G.A., 2004. Fossil hyrax dung and evidence of Late Pleistocene and Holocene vegetation types in the Namib Desert. Journal of Quaternary Science 19, 829832.CrossRefGoogle Scholar
Shi, N., Dupont, L.M., Beug, H.-J., Schneider, R., 2000. Correlation between vegetation in southwestern Africa and oceanic upwelling in the past 21,000 years. Quaternary Research 54, 7280.CrossRefGoogle Scholar
Shi, N., Schneider, R., Beug, H.-J., Dupont, L.M., 2001. Southeast trade wind variations during the last 135 kyr: evidence from pollen spectra in eastern South Atlantic sediments. Earth and Planetary Science Letters 187, 311321.CrossRefGoogle Scholar
Slota, P.J., Jull, A.J.T., Linick, T.W., Toolin, L.J., 1987. Preparation of small samples for 14C accelerator targets by catalytic reduction of CO. Radiocarbon 29, 303306.CrossRefGoogle Scholar
Smith, B.N., 1972. Natural abundance of the stable isotopes of carbon in biological systems. BioScience 22, 226231.CrossRefGoogle Scholar
Smith, B.N., Epstein, S., 1971. Two categories of 13C/12C ratios for higher plants. Plant Physiology 47, 380384.CrossRefGoogle Scholar
Stocker, T.F., 1998. The seesaw effect. Science 282, 6162.CrossRefGoogle Scholar
Stocker, T.F., Johnsen, S.J., 2003. A minimum thermodynamic model for the bipolar seesaw. Paleoceanography 18, 1087. https://doi.org/10.1029/2003PA000920.CrossRefGoogle Scholar
Stockmarr, J., 1971. Tablets with spores used in absolute pollen analysis. Pollen et Spores 13, 614621.Google Scholar
Stute, M., Talma, A.S., 1998. Glacial temperatures and moisture transport regimes reconstructed from noble gas and ∂18O, Stampriet aquifer, Namibia (IAEA-SM-349/53). In: Isotope Techniques in the Study of Environmental Change. Proceedings of an International Symposium on Isotope Techniques in the Study of Past and Current Environmental Changes in the Hydrosphere and the Atmosphere Organized by the International Atomic Energy Agency and Held in Vienna, 14–18 April 1997. International Atomic Energy Agency, Vienna, Austria, pp. 307318.Google Scholar
Stuut, J.-B.W., Prins, M.A., Schneider, R.R., Weltje, G.J., Jansen, J.H.F., Postma, G., 2002. A 300 kyr record of aridity and wind strength in southwestern Africa: inferences from grain-size distributions of sediments on Walvis Ridge, SE Atlantic. Marine Geology 180, 221233.CrossRefGoogle Scholar
Tankard, A.J., Rodgers, J., 1978. Late Cenozoic palaeoenvironments on the west coast of southern Africa. Journal of Biogeography 5, 319337.CrossRefGoogle Scholar
Teller, J.T., Lancaster, N., 1985. History of sediments at Khommabes, central Namib Desert. Madoqua 14, 267278.Google Scholar
Teller, J.T., Lancaster, N., 1986. Lacustrine sediments at Narabeb in the central Namib Desert, Namibia. Palaeogeography, Palaeoclimatology, Palaeoecology 56, 177195.CrossRefGoogle Scholar
Trabucco, A., Zomer, R., 2019. Global Aridity Index and Potential Evapotranspiration (ET0) Climate Database v2. figshare. Dataset. https://doi.org/10.6084/m9.figshare.7504448.v3.Google Scholar
UNEP (United Nations Environment Program), 1997. World Atlas of Desertification, 2nd ed. John Wiley & Sons, New York.Google Scholar
Urrego, D.H., Sánchez Goñi, M.F., Daniau, A.L., Lechevrel, S., Hanquiez, V., 2015. Increased aridity in southwestern Africa during the warmest periods of the last interglacial. Climate of the Past 11, 14171431.CrossRefGoogle Scholar
van Zinderen Bakker, E M, 1975. The origin and palaeoenvironment of the Namib Desert biome. Journal of Biogeography 2, 6573.CrossRefGoogle Scholar
van Zinderen Bakker, E.M., 1953. South African Pollen Grains and Spores. Part I. Balkema, Amsterdam/Cape Town.Google Scholar
van Zinderen Bakker, E.M., 1956. South African Pollen Grains and Spores. Part II. Balkema, Amsterdam/Cape Town.Google Scholar
van Zinderen Bakker, E.M., 1975. The origin and palaeoenvironment of the Namib Desert biome. Journal of Biogeography 2, 6573.CrossRefGoogle Scholar
van Zinderen Bakker, E.M., Coetzee, J.A., 1959. South African Pollen Grains and Spores. Part III. Balkema, Amsterdam/Cape Town.Google Scholar
Vogel, J.C., 1982. The age of the Kuiseb River silt terrace at Homeb (Namibia). Palaeoecology of Africa 15, 201209.Google Scholar
Vogel, J.C., Visser, E., 1981. Pretoria radiocarbon dates II. Radiocarbon 23, 4380.CrossRefGoogle Scholar
Ward, J.D., 1982. Aspects of a suite of Quaternary conglomeratic sediments in the Kuiseb valley, Namibia. Palaeoecology of Africa 15, 211216.Google Scholar
Ward, J.D., Seely, M.K., Lancaster, N., 1983. On the antiquity of the Namib. South African Journal of Science 79, 175183.Google Scholar
Werger, M.J.A., Ellis, R.P., 1981. Photosynthetic Pathways in the Arid Regions of South Africa. Flora 171, 6475.CrossRefGoogle Scholar
Figure 0

Figure 1. (A) Map of study region, indicating location of the Zizou rock hyrax midden site. (B) The Zizou-1 rock hyrax midden (21.3 cm thick), with view of local landscape and vegetation (C) and (D) view eastward from the site towards the Namib Sand Sea (visible in distance).

Figure 1

Figure 2. Maps of study region showing Aridity Index gradients (data from Trabucco and Zomer, 2019), an elevation transect across the map at the latitude of the Zizou hyrax midden site, the distribution of ecoregions (from Dinerstein et al., 2017), and the association of each vegetation type in the region with aridity.

Figure 2

Figure 3. Distribution in time and depth of 14C ages for the Zizou rock hyrax middens with age-depth model (grey), overlying the calibrated distributions of the individual dates calculated using Bacon v2.5.8 (Blaauw and Christen, 2011). Red dotted lines indicate the ‘best’ model based on the mean age. The Holocene section of the ZIZ-1-1 age model was calculated in isolation (upper panel) to better constrain the marked post-Pleistocene change in accumulation rate.

Figure 3

Table 1. Radiocarbon ages and calibration information for the Zizou-1-1 and Zizou-1-3b rock hyrax middens.

Figure 4

Figure 4. Diagram of pollen and microcharcoal from the Zizou rock hyrax middens, including taxa > 2% of composition. Mean ages for samples, including minimum and maximum depths, are provided.

Figure 5

Figure 5. δ15N, δ13C, microcharcoal concentration, and select pollen data from the Zizou rock hyrax middens illustrated along with the Namib composite δ15N record from Chase et al., 2019, and high latitude boreal and low latitude austral summer insolation (Laskar et al., 2004). Holocene, Last Glacial Maximum (LGM), and Heinrich stadial (HS1, HS2, HS3; grey bars) chronozones are indicated.