Plant Ash

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

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

net/publication/316699759

Plant Ash.

Chapter · September 2017


DOI: 10.1002/9781118941065.ch17

CITATIONS READS

21 2,897

2 authors:

Matthew Canti Jacques Elie Brochier


Historic England French National Centre for Scientific Research, Aix-en-Provence, France
58 PUBLICATIONS 2,566 CITATIONS 127 PUBLICATIONS 2,060 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Matthew Canti on 05 November 2019.

The user has requested enhancement of the downloaded file.


147

17

Plant Ash
Matthew G. Canti1 and Jacques Élie Brochier2
1
Historic England, Eastney, United Kingdom
2
Aix-Marseille Université (AMU), UMR 7269 LAMPEA, Aix‐en‐Provence, France

17.1 Introduction 17.1.2 Ash Formation

On heating, calcium oxalate transforms directly into


Ash is formed when oxidation of plant material in fires calcium carbonate at around 400–500 °C by loss of CO
has progressed beyond charring, to the point where the (Pobeguin 1943; Dollimore 1987; Frost & Weier 2003,
carbon, hydrogen and oxygen components are largely 2004). If fires get hotter, CO2 is driven off leaving CaO,
vaporized. The previously small quantities of mineral which, on cooling becomes Ca(OH)2 then recarbonates
constituents in the plant then become the major with atmospheric CO2 to form CaCO3. Experiments
component left behind as ash. with plant calcium oxalates show that the transformation
into lime occurs around 600 °C, which is lower than the
17.1.1 Ash Composition conversion temperature measured on synthetic calcium
The main mineral constituents of plants are silica (sili- oxalate (Pobeguin 1943; Brochier 1983). Recent experi-
ceous phytoliths) and calcium oxalate (oxalate phytoliths), mentation has, however, measured a higher transforma-
which exist in different forms and quantities according tion temperature (ca. 700 °C) from some plant oxalates
to the plant families. (Shahack‐Gross & Ayalon 2013), suggesting that influ-
Silica is widespread in plants but commonest in the ences from other variables are yet to be clarified for this
monocotyledons, especially grasses and horsetails, where system. The calcium carbonate in ash can be derived
it can be a large percentage of the ash composition. It through either of these transformation (direct or by
takes the form of intercellular and intracellular bodies, recarbonation) routes (Folk 1973; Shahack‐Gross &
either discrete phytoliths (Figure 17.1), multicellular Ayalon 2013).
cuticle structures (Figure 17.2) or silica skeletons Although some changes occur during heating (notably
cementing the whole vascular system (Figure 17.3). the loss of OH groups), silica is mostly unaffected by the
Calcium oxalate (as the monoclinic monohydrate temperatures in normal domestic fires, as long as it does
whewellite, CaC2O4.H2O; or the tetragonal dihydrate not enter into a melt with alkaline salts (see section 17.2).
weddellite, CaC2O4.2H2O) is ubiquitous in most herbs, Variable quantities of mainly potassium and sodium salts
trees and shrubs but very rare in grasses (Metcalfe 1960; are present in fresh ash, and these can play a part in
Prychid & Rudall 1999). It occurs in a variety of shapes subsequent taphonomic changes.
(see for examples Scurfield et al. 1973; Franceschi &
Horner 1980); the commonest types are the prismatic
crystals, which are single or twinned individual crystals, 17.2 Micromorphology
typically 10–20 µm (Figure 17.4) and the druse, which
is a 15–50 µm rosette of radially aligned crystals The transformation of calcium oxalate into calcium
(Figure 17.5). They are generally not preserved in archae- carbonate results in the molecular rearrangement of the
ological sediments, but have occasionally been found relatively large oxalate crystals into microcrystalline
protected in degraded plant tissues (Figure 17.6). aggregates of CaCO3, which retain the varied gross

Archaeological Soil and Sediment Micromorphology, First Edition. Edited by Cristiano Nicosia and Georges Stoops.
© 2017 John Wiley & Sons Ltd. Published 2017 by John Wiley & Sons Ltd.
148 Archaeological Soil and Sediment Micromorphology

100 μm

20 μm

Figure 17.1 Silica phytoliths from garden soil deposits in


Leicester, United Kingdom. PPL.
Figure 17.4 Calcium oxalate crystals lining the veins in a leaf of
birch (Betula pendula). PPL.

30 μm

25 μm
Figure 17.2 Dendriform phytoliths from Graminae. Middle Bronze
Age layers, Cazals cave (Aude). Viewed with interference contrast,
which improves image of isotropic materials. Figure 17.5 Druse, twinned and single prismatic calcium oxalate
crystals from a leaf of downy oak (Quercus pubescens). A few
siliceous tracheids (isotropic) are also visible. Viewed with
interference contrast, which allows birefringent and isotropic
materials to be seen in the same image.

morphologies of the oxalates. Thus, both prismatic and


druse pseudomorphs (Figures 17.7, 17.8 and 17.9) may
be found intact in ash (Brochier 1983; Brochier et al.
1992; Cailleau et al. 2005).
The pseudomorphs are made up of very small calcite
crystals, a few of which can be seen under the highest
magnification, but the remainder are submicron in size.
The smaller the crystals are, the higher is the proportion
of the microscopic view being taken up by the dark crys-
tal outlines. The pseudomorphs are thus darkened
250 μm almost to opacity in plane polarized light due to com-
pound relief produced by multiple overlying crystals
Figure 17.3 Silica skeletons left behind after burning of grasses (Figure 17.9a). Under crossed polars (Figure 17.9b),
(Graminae sp.). PPL. the compound birefringence of the stacked submicron
Plant Ash 149

(a)

5 μm

(b)

Figure 17.7 SEM photograph of a twinned prismatic oxalate


crystal transformed into microcrystalline calcium carbonate.

(c)

Figure 17.8 SEM photograph of a twinned prismatic oxalate


crystal transformed into microsparitic calcium carbonate from
Neolithic layers in Santa Maira rock shelter (Alicante). Image
courtesy of C. Verdasco.
Figure 17.6 (a) Calcium oxalate crystals preserved in degraded
plant tissue in an ash deposit; note the grey colouration of the
microcrystalline calcium carbonate pseudomorphs produced by
compound relief (C), and the clearer light transmission through crystals produces high‐order interference colours but
the intact oxalate crystals (O). PPL. (b) Note the weaker compound the opacifying effect of the compound relief acts to
interference colours of the microcrystalline calcium carbonate
strongly reduce the overall brightness (Canti 2003). The
pseudomorphs (C) and the stronger ordered bands of interference
colours on the intact oxalate crystals (O). XPL. (c) Note the bright submicron crystals are randomly oriented; thus the
diffused white of the calcium carbonate pseudomorphs (C) and pseudomorphs never go into extinction when the stage
the duller transmitted light of the intact oxalate crystals (O). is rotated, although very elongated prisms may occa-
Reflected light. Mesolithic layers, Cuzoul de Gramat rock shelter. sionally show partial extinction. Under reflected light,
Image courtesy of G. Delfour.
they appear as bright white crystalline forms owing to
150 Archaeological Soil and Sediment Micromorphology

(a) (b)

Figure 17.9 (a) Calcium carbonate pseudomorphs of prismatic crystals from modern bonfire ash, PPL. (b) Same as (a), XPL.

(a) (b)

Figure 17.10 (a) Pinus sylvestris L. (b) Tilia platyphyllos Scop. Radial sections showing calcitic pseudomorphs of prismatic oxalate crystals in
the phloem parenchyma cells. Photos: J. É. Brochier and M. Thinon.

light diffusion (see white areas in Figure 17.6c). Some


oxalate pseudomorphs, frequently observed in archaeo-
logical deposits, have slightly larger crystals (micro-
sparite – Figure 17.8). Their shapes are less sharp, and
they are sometimes difficult to recognize. They have
never been produced experimentally and the formation
process is still unclear, although some type of recrystal-
lization seems most likely (Brochier & Thinon 2003).
Although there are numerous forms of calcitic pseu-
domorphs after oxalate (as numerous as the crystal
habits of plant whewellite and weddellite), very few of
them can be attributed to a specific taxon. It is, how-
ever, possible to recognize, on morphometric grounds
(Brochier 1991, 1995), the genus Pinus (very elongated
prisms) and the two species (cordata and platyphyllos)
of the genus Tilia (elongated and short prisms associ-
ated with druses). See Figures 17.10, 17.11 and
Table 17.1.
Figure 17.11 Prismatic pseudomorph from Balma de la
It is often not possible to see the edges of densely Margineda (Andorra, Early Neolithic). The elongation index of
packed siliceous phytoliths in thin sections, but the 3.6 and width of 16 µm indicates the genus Tilia. PPL.
Plant Ash 151

Table 17.1 Statistical parameters of elongation index, length and width for Tilia cordata, T. platyphyllos, Pinus
halepensis and prismatic calcium carbonate pseudomorphs.

Distribution data Pinus halepensis Pinus sylvestris Tilia cordata Tilia platyphyllos

Elongation index (length/width)


Mean ± standard error 5.33 ± 0.29 6.06 ± 0.42 2.76 ± 0.08 3.88 ± 0.16
Standard deviation 1.01 1.57 0.48 1.03
Median 5.20 5.84 2.67 3.77
Quartile 1 4.38 4.73 2.38 3.22
Quartile 3 6.29 6.93 3.05 4.54
Maximum 7.2 10.0 4.0 6.6
Minimum 4.0 4.0 1.8 2.1

Length (µm)
Mean ± standard error 39.0 ± 2.0 30.5 ± 2.3 43.6 ± 1.3 45.0 ± 1.6
Standard deviation 6.8 8.5 8.1 10.7
Median 40.0 29.3 41.8 45.3
Quartile 1 35.6 23.6 37.5 35.0
Quartile 3 43.8 35.6 48.1 52.0
Maximum 50.0 50.0 62.0 70.0
Minimum 25.0 20.0 29.5 25.0

Width (µm)
Mean ± standard error 7.5 ± 0.5 5.0 ± 0.2 16.2 ± 0.6 11.9 ± 0.4
Standard deviation 1.8 0.6 3.7 2.3
Median 7.3 5.0 15.0 10.5
Quartile 1 6.1 4.9 13.9 10.0
Quartile 3 9.4 5.0 19.3 14.6
Maximum 10.0 7.0 25.0 15.5
Minimum 5.0 4.5 10.0 7.0

negative relief of opaline silica (R.I. typically 1.41–1.46) 1967; Mulholland & Prior 1992), but see Santos et al.
is frequently noticeable on the surfaces and sculpturing (2012) for problems with this approach.
of phytoliths in plane polarized light. Under crossed Silica can undergo significant morphological altera-
polars, silica phytoliths are isotropic (Figure 17.12). tion by fusion. This happens when the ash contains a
Although the gross morphology of opaline silica is high enough proportion of fluxing agents such as sodium
largely unchanged in ash, heating phytoliths does alter and potassium, which lower the overall melting point of
the refractive index by removing the water and hydroxyl the mixture. The resulting material ranges from a lightly
groups (Elbaum et al. 2003). The effect must technically fused cast of the original plant silica skeleton through to
be visible in thin sections but the range of RI differences lumps of glassy clinker‐like material, usually vesicular
is quite small (ca. 0.05) and measuring it requires delib- (Figure 17.13) due to gas bubbles developing in the liquid
erate sample preparation in refractive oil. Phytoliths can melt (Vélain 1878; Folk & Hoops 1982; Thy et al. 1995;
be darkened by burning but some are naturally dark Guélat et al. 1998; Brochier & Claustre 2000; Canti 2007).
without any heating (Parr 2006). These colour differ- Fused silica can form considerable accumulations as, for
ences are thought to be due to occluded carbon, which example, in the Deccan ash mounds (Allchin 1963) and
can also provide valuable radiocarbon dates (Wilding burnt kraals in South Africa (Jacobson et al. 2003).
152 Archaeological Soil and Sediment Micromorphology

(a) (b)

Figure 17.12 Silicified plant remains (Si) in mixed ash from the fuel of a corn drying oven at Grateley, United Kingdom. (a) PPL; (b) XPL.

17.3 Ash Taphonomy


Once deposited, ash can undergo a wide range of tapho-
nomic processes, including trampling, mixing, colluvia-
tion and recrystallization (Brochier 1993, 2002; Karkanas &
Goldberg 2010; also see Mallol et al. 2017, this book).
If retained in a dry, undisturbed environment, ash will
preserve indefinitely, but this is not the commonest
taphonomic outcome. Many outdoor archaeological sit-
uations have enough rainfall to dissolve it completely,
sometimes leaving a fine charcoal layer to mark where
the ash used to be. Areas of restricted rainfall, for exam-
ple Middle Eastern open‐air sites, will retain ash to a
considerable extent where burial is rapid (e.g., Shillito
et al. 2011; Friesem et al. 2014; Gur‐Arieh et al. 2014).
Figure 17.13 Silica ash melt from Northampton, United
Kingdom. PPL. The best preservation is generally found in caves and

(a) (b)

Figure 17.14 (a) Ash from Early Neolithic cave deposits in the Upper Ebro valley with slightly browner central area of phosphatic staining,
PPL. (b) Same as (a), with central area of phosphatic staining producing isotropy. XPL. Image courtesy of A. Polo Díaz.
Plant Ash 153

rock shelters. These have often sustained long periods of Near‐isotropic areas in ashy deposits often signal this
occupation, sometimes with repeated burning of animal process (Figure 17.14).
bedding and manure, resulting in deep ash accumula- Geogenic calcite in the form of either fragments of
tions. However, even caves contain moist air, dripping limestone from cave walls / roof or speleothems is some-
water and bat colonies, which together produce a range times difficult to distinguish from recrystallized calcitic
of effects. The ash itself will sometimes partially dissolve ash without additional analyses (Shahack‐Gross et al.
and recrystallize as neoformed calcite, losing some or all 2008; Shahack‐Gross & Ayalon 2013). Where detrital
of its oxalate pseudomorph characteristics. In the pres- inputs are significant to interpretation, variation in the
ence of phosphates (from either degrading bat guano or abundance of calcitic pseudomorphs dispersed in an
dissolving bones), the ash calcite will convert to apatite archaeological profile can play a useful role (along with
or other phosphates (Karkanas et al. 2000; Shahack‐ other anthropic particles) in helping determine sedimen-
Gross et al. 2004) and therefore loose its birefringence. tation / dissolution rates (Brochier 1993, 1995).

References
Allchin, F. R. (1963) Neolithic Cattle Keepers of South pastoral sites. Journal of Anthropological Archaeology
India. A Study of the Deccan Ashmounds. Cambridge 11, 47–102.
University Press, Cambridge. Cailleau, G., Braissan, O., Dupraz C. et al. (2005)
Brochier, J. É. (1983) Bergeries et feux néolithiques dans le Biologically induced accumulations of CaCO3 in orthox
Midi de la France, caractérisation et incidence sur le soils of Biga, Ivory Coast. Catena 59, 1–17.
raisonnement sédimentologique. Quatar 33/34, Canti, M. G. (2003) Aspects of the chemical and
181–193. microscopic characteristics of plant ashes found in
Brochier, J. É. (1991) Géoarchéologie du monde archaeological soils. Catena 54, 339–361.
agropastoral. In: Guilaine, J. (ed.) Pour une Archéologie Canti, M. G. (2007) Geoarchaeological studies at
Agraire,. Armand Colin, Paris, pp. 303–322. Flixborough‐North Conesby: understanding the burial
Brochier, J. É. (1993) Çayönü Tepesi. Domestication, environment, taphonomy and preservation conditions.
rythmes et environnement au PPNB. Paléorient 19/2, In: Loveluck, C. & Atkinson, D. (eds) The Early Medieval
39–49. Settlement Remains from Flixborough, Lincolnshire:
Brochier, J. É. (1995) Estudi geoarqueològic dels dipòsits The Occupation Sequence, c. AD 600–1000. Oxbow
holocens de la Balma de la Margineda: capes de 1 a la 6. Books, Oxford, pp. 17–23.
(Étude géoarchéologique des dépôts holocènes de la Dollimore, D. (1987) The thermal decomposition of
Balma Margineda, Andorre, couches 1 à 6). In: Guilaine, J. oxalates. A review. Thermochimica Acta 117, 331–363.
& Martzluff, M. (eds) Les Excavacions a la Balma de la Elbaum, R., Weiner, S., Albert, R. M. et al. (2003) Detection
Margineda (1979–1991), Vol. 1. Edicions del Govern of burning of plant materials in the archaeological
d’Andorra, Andorra, pp. 56–90. record by changes in the refractive indices of siliceous
Brochier, J. É. (2002) Les sédiments anthropiques. phytoliths. Journal of Archaeological Science 30,
Méthodes d’étude et perspectives. In: Miskovsky, J. C. 217–226.
(ed.) Géologie de la Préhistoire: Méthodes, Techniques, Folk, R. L. (1973) The geological framework of Stobi.
Applications. Géopré Éditions, Paris, pp. 453–477. Studies in the Antiquities of Stobi, 1, 37–59.
Brochier, J. É. & Claustre, F. (2000) Le parcage des bovins Folk, R. L. & Hoops, G. K. (1982) An early Iron‐Age layer
et le problème des litières du Néolithique final à l’Âge du of glass made from plants at Tel Yin’am, Israel. Journal of
Bronze dans la Grotte de Bélesta. In: Vaquer, J., Guilaine, Field Archaeology 9, 455–456.
J. & Martin, H. (eds) Le Néolithique du Nord‐ouest Franceschi, V. R. & Horner, H. T. (1980) Calcium oxalate
Méditerranéen: XXIVe Congrès Préhistorique de France, crystals in plants. The Botanical Review 46, 361–427.
Carcassonne, 1994. Actes Société Préhistorique Friesem, D. E., Zaidner, Y. & Shahack‐Gross, R. (2014)
Française, Paris, pp. 27–36. Formation processes and combustion features at the
Brochier, J. É. & Thinon M. (2003) Calcite Crystals, Starch lower layers of the Middle Palaeolithic open‐air site of
Grains Aggregates or … POCC? Comment on ‘Calcite Nesher Ramla, Israel. Quaternary International 331,
Crystals inside Archaeological Plant Tissues’. Journal of 128–138.
Archaeological Science 30, 1211–1214. Frost, R. L. & Weier, M. L. (2003) Thermal treatment of
Brochier, J. É., Villa, P., Giacomarra, M. et al. (1992) weddellite – a Raman and infrared emission
Shepherds and sediments: Geo‐ethnoarchaeology of spectroscopic study. Thermochimica Acta 406, 221–232.
154 Archaeological Soil and Sediment Micromorphology

Frost, R. L. & Weier, M. L. (2004) Thermal treatment of Pobeguin, T. (1943) Les oxalates de calcium chez quelques
whewellite – a thermal analysis and Raman angiospermes. Annales des Sciences Naturelles,
spectroscopic study. Thermochimica Acta 409, 79–85. Botanique et Biologie Vegetale (11th series) 4, 1–95.
Guélat, M., Paccolat, O. & Rentzel, P. (1998) Une étable Prychid, C. J. & Rudall, P. J. (1999) Calcium oxalate crystals
gallo‐romaine à Brigue‐Glis VS, Waldmatte. Annuaire in Monocotyledons : A review of their structure and
de la Société Suisse de Préhistoire et d’Archéologie 81, systematics. Annals of Botany 84, 725–739.
171–182. Santos, G. M., Alexandre, A., Southon, J. R. et al. (2012)
Gur‐Arieh, S., Shahack‐Gross, R., Maeir, A. M. et al. (2014) Possible source of ancient carbon in phytolith
The taphonomy and preservation of wood and dung concentrates from harvested grasses. Biogeosciences 9,
ashes found in archaeological cooking installations: case 1873–1884.
studies from Bronze and Iron Age Israel. Journal of Scurfield, G., Michell, A. J. & Silva, S. R. (1973) Crystals in
Archaeological Science 46, 50–67. woody stems. Botanical Journal of the Linnean Society
Jacobson, L., Loock, J. C., van der Westhuizen, W. A. et al. 66, 277–289.
(2003) The occurrence of vitrified dung from the Shahack‐Gross, R., Ayalon, A, Goldberg, P. et al. (2008)
Kamdeboo district, southern Karoo, and Den Staat, Formation processes of cemented features in karstic cave
Limpopo Valley, South Africa. South African Journal of sites revealed using stable oxygen and carbon isotopic
Science 99, 26–28. analyses: A case study at Middle Paleolithic Amud Cave,
Karkanas, P., Bar‐Yosef, O., Goldberg, P. et al. (2000) Israel. Geoarchaeology 23, 43–62.
Diagenesis in prehistoric caves: The use of minerals Shahack‐Gross, R. & Ayalon, A. (2013) Stable carbon and
that form in situ to assess the completeness of the oxygen isotopic compositions of wood ash:
archaeological record. Journal of Archaeological An experimental study with archaeological implications.
Science 27, 915–929. Journal of Archaeological Science 40, 570–578.
Karkanas, P. & Goldberg, P. (2010) Site formation processes Shahack‐Gross, R., Berna, F., Karkanas, P. et al. (2004) Bat
at Pinnacle Point Cave 13B (Mossel Bay, Western Cape guano and preservation of archaeological remains in cave
Province, South Africa): resolving stratigraphic and sites. Journal of Archaeological Science 31, 1259–1272.
depositional complexities with micromorphology. Shillito, L‐M., Bull, I. D., Matthews, W. et al. (2011)
Journal of Human Evolution 59, 256–273. Biomolecular and micromorphological analysis of
Mallol, C., Mentzer, S. M. & Miller C. E. (2017) suspected faecal deposits at Neolithic Çatalhöyük,
Combustion features. In: Nicosia, C. & Stoops, G. (eds) Turkey. Journal of Archaeological Science 38, 1869–1877.
Archaeological Soil and Sediment Micromorphology. Thy, P., Segobye, A. K. & Ming, D. W. (1995) Implications
John Wiley & Sons, Ltd, Chichester, pp. 299–330. of prehistoric glassy biomass slag from East‐Central
Metcalfe, C. R. (1960) Anatomy of the Monocotyledons. I. Botswana. Journal of Archaeological Science 22,
Graminae. Clarendon Press, Oxford. 629–637.
Mulholland, S. C. & Prior, C. (1992) Processing of Vélain, C. H. (1878) Étude microscopique des verres
phytoliths for radiocarbon dating by AMS. The résultant de la fusion des cendres de graminées. Bulletin
Phytolitharien Newsletter 7, 7–9. de la Société Française de Minéralogie 1, 113–125.
Parr, J. F. (2006) Effect of fire on phytolith coloration. Wilding, L. P. (1967) Radiocarbon dating of biogenic opal.
Geoarchaeology 21, 171–185. Science 156, 66–67.

View publication stats

You might also like