Skip to content
BY 4.0 license Open Access Published by De Gruyter (O) July 3, 2023

Artefacts from Ban Chiang, Thailand: Pre-metal Age cord-marked pottery

  • Waraporn Tanthanuch , Somchai Tancharakorn , Catleya Rojviriya and Ulrich Bismayer EMAIL logo

Abstract

Pottery excavated from the archeological UNESCO world heritage site Ban Chiang in Thailand stem from distinct periods. Black vessels with cord-mark design from Pre-metal Age (ca. 3000–1000 BC), Bronze Age pottery (ca 1000–300 BC) with yellow-brown surface and Iron Age samples (ca. 300 BC–200 AD) with red pictorial surface patterns. In a previous work [Bismayer U., Srilomsak S., Treekamol Y., Tanthanuch W., Suriyatham K. Artefacts from Ban Chiang, Thailand: pottery with hematite-red geometric patterns. Z. Kristallogr. 2020, 235, 559–568] we studied the mineralogical composition and their surface colour materials of shards from Bronze and Iron Age. In this work we focus on bulk features of the dark Pre-metal Age cord-marked ceramic shard PSN2-S10E13 and compare its elemental and mineralogical composition with bulk composition of sample 5412-S6E15 from Bronze Age. Experimental techniques are electron microprobe, X-ray powder diffraction, FTIR spectroscopy, optical microscopy and X-ray tomographic microscopy (XTM). Sample PSN2-S10E13 contains more quartz than 5412-S6E15. In the bulk of the Pre-metal Age shard, diffraction signals of mullite occur, indicating higher firing temperatures compared to the younger sample. Phyllosilicate signals are seen in FTIR spectra of both shards. E-modes of quartz dominate FTIR spectra of both samples. Optical thin sections show voids around micro-particles in PSN2-S10E13 and XTM indicates that the pore volume percentage of sample PSN2-S10E13 is higher than in 5412-S6E15. Because of the large age gap to younger samples from Ban Chiang, the proper age of our oldest sample PSN2-S10E13 was determined using an accelerator mass spectrometer (AMS) by simultaneous 14C/12C and 13C/12C isotope ratio measurements which yielded a radiocarbon age of 3609 ± 29 BP (resp. 1659 ± 29 BC).

1 Introduction

Because of its extraordinary and characteristic ceramic artefacts with unique coloured patterns on Iron Age pottery and cord-marks on the surface of older Neolithic ceramics the archeological excavation site Ban Chiang in the Udon Thani province, Northeast Thailand, became a UNESCO world heritage in 1992. Intensive studies on ethnic groups living there [1], on their ceramic technology [2], on their archeological classification [3, 4] as well as on the geological setting [5], [6], [7], [8] have been published. Ban Chiang lies on the Indochina micro-plate in the northern part of the Sakon Nakhon Basin. The plate extends to the Khorat Plateau towards the south. Mineral deposits in the area result from Cretaceous granites, granodiorites and pegmatites [8]. Because of repeated flooding combined with chemical weathering of rocks the basin became an extended saline area where Tertiary marine evaporation and non-marine silica-rich sedimentation took place [5], [6], [7]. Several authors discussed possible migration and immigration in the northern-east area of today’s Thailand [1, 3, 9] using isotope analysis and radiocarbon dating of human tooth enamel and bone remains.

Artefacts of Ban Chiang are well recognizable, especially from the late Iron Age, ca. 300 BC to 200 AD. At their surface most of the Iron Age ceramics show characteristic ’zoomorphic’ design and symmetric patterns [10, 11]. The red colour consist of hematite-rich natural pigments [12] and its distinctive style may be considered as precursors of writing with some resemblance to other archeological sites [13, 14]. Ceramics from Bronze Age and early Neolithic Pre-metal Age of Ban Chiang have a different appearance. Bronze Age pottery shows a smooth brown-yellowish surface while Pre-metal Age pottery has handmade impressions, so called cord-marks at the surface also apparent e.g. in other Asian [15], American [16], Near East [17, 18] and African ceramics [19].

In previous work [20] redox-conditions and firing-temperatures had been studied using XANES amongst other techniques and it had been concluded from Fe K-edge X-ray absorption signals that Ban Chiang pottery shards were fired in all periods under reducing conditions but at different temperatures. The authors concluded that Pre-metal shards showing dark to black surface were produced at temperatures above 1300 K. Surface painting and mineral compositions had previously been analysed [12] while bulk properties like grain arrangements, mineral composition and porosity have not been investigated in detail yet. Hence, we focus in this work on compositional and morphological features of the bulk and compare Pre-metal Age pottery with a younger Bronze Age sample excavated in Ban Chiang. Apart from chemical and mineralogical techniques the volume of the samples was studied using X-ray tomographic microscopy to explore more qualitative and quantitative details of texture features of the artefacts.

2 Experimental

2.1 Samples

Both ceramic shards used in this study stemmed from Ban Chiang and had been made in different periods. The older shard PSN2-S10E13 is a Pre-metal Age cord-marked ceramic piece and has a dark grey surface (Figure 1a). The other shard 5412-S6E15 (Figure 1b) is from Bronze Age (ca. 1000–300 BC) with a yellow-brown surface which had previously been studied with focus on its surface materials [12] and redox- and firing-conditions [20]. For X-ray powder diffraction and FTIR spectroscopy ca. 3–4 mg material was removed from the bulk of the respective samples. Other techniques covered microprobe analysis and the investigation of optical thin sections prepared perpendicular to the sample surface. The chemical composition was analysed using an electron microprobe (Cameca Cambax micro-beam SEM system), technical details have been described by Beirau et al. [21]. The acceleration voltage was 15 kV, the sample current 40 nA and the beam diameter 30 μm. Table 1 shows the result of the analysis.

Figure 1: 
Ceramic shards from Ban Chiang, Thailand. (a) Pre-metal Age cord-marked shard PSN2-S10E13 and (b) Bronze Age sample 5412-S6E15.
Figure 1:

Ceramic shards from Ban Chiang, Thailand. (a) Pre-metal Age cord-marked shard PSN2-S10E13 and (b) Bronze Age sample 5412-S6E15.

Table 1:

Electron microprobe analysis of Ban Chiang ceramics in wt.% oxides. Average chemical composition of (a) bulk of sample 5412-S6E15 (data from [12]) and (b) bulk composition of PSN2-S10E13 (this study). aA thin section of sample PSN2-S10E13 and line along analysed points is shown in Figure 2. Standards and standard deviations are described in [21, 22].

Sample bulk (a) Bronze Age

5412-S6E15
(b) Pre-metal Age

PSN2-S10E13
Oxide wt.% wt.%
Na2O 0.27 0.16
MgO 1.00 0.90
Al2O3 20.09 18.86
SiO2 59.09 63.76
K2O 1.97 1.19
CaO 0.92 0.70
TiO2 0.54 0.89
Cr2O3 <Det. lim <Det. lim
MnO <Det. lim <Det. lim
Fe2O3 4.21 4.37
Total 88.15 90.87
  1. a<Det. lim: MnO and Cr2O3 were below the quantitative detection limit. Total is below 100 wt.% because of pores and unknown amount of OH-groups in the samples.

Texture of the ceramics and mineralogical components of the bulk were studied using thin sections with standard thickness (25 μm), optical microscopy and using the BSE imaging technique of the electron microprobe.

The thin section (Figure 2a) shows brighter quartz grains in a dark matrix of fine clay mineral particles. A few grains show cleavage. Electron microprobe analysis of the thin section (Figure 2b) of sample PSN2-S10E13 was carried out at 100 points along a 1500 μm line starting from the sample surface towards the bulk. No chemical difference between the surface and the bulk occurred. In the bulk material of sample PSN2-S10E13, a few quartz crystals were observed like in the optical thin section (Figure 2a). The average of the microprobe measurements is shown in column b of Table 1. Chemical microprobe analysis of the younger sample 5412-S6E15 was done at 100 points along a line of 1200 μm published previously (column a of Table 1, data from [12]).

Figure 2: 
Inside the Pre-metal Age shard from Ban Chiang, Thailand. (a) Thin section of sample PSN2-S10E13, unpolarized; (b) back-scattering electron image of chemically analysed points along the indicated line of the Pre-metal Age sample. Grey scale levels correspond to relative amount of back-scattered electrons where black areas (cracks, voids) indicate no back-scattering.
Figure 2:

Inside the Pre-metal Age shard from Ban Chiang, Thailand. (a) Thin section of sample PSN2-S10E13, unpolarized; (b) back-scattering electron image of chemically analysed points along the indicated line of the Pre-metal Age sample. Grey scale levels correspond to relative amount of back-scattered electrons where black areas (cracks, voids) indicate no back-scattering.

2.2 Age determination

Age determination of sample PSN2-S10E13 was done at the Leibniz Laboratory for Radiometric Dating and Stable Isotope Research, Christian-Albrechts-University, Kiel, Germany using Accelerator Mass Spectrometry (AMS) and classical calibration [23], [24], [25]. Only bulk material was used for dating in order to avoid incorrect measurements stemming from surface contamination. The selected material was subsequently dissolved in 1 % HCl, 1 % NaOH at 333 K and again 1 % HCl. The CO2 combustion of the alkali residue was performed in an evacuated quartz tube with CuO and silver at 1173 K. The obtained CO2 was reduced at 873 K using H2 and iron powder as catalyst. The resulting iron graphite mixture was pressed into a target holder for AMS measurement. The measurement was carried out using a HVE 3 MV Tandetron 4130 accelerator mass spectrometer. The 14C/12C and 13C/12C isotope ratios were simultaneously measured and then compared to the CO2 measurement standards (oxalic acid II). The result was corrected for effects of exposure to foreign carbon during the sample pretreatment. The resulting 14C-content was corrected for isotope fractionation, related to the hypothetical atmospheric value of 1950 given in pMC (percent Modern Carbon). This value was used to calculate the radiocarbon age according to [23]. The reported uncertainty of the 14C result takes into account the uncertainty of the measured 14C/12C ratios of sample and measurement standard, the uncertainty of the fractionation correction and the uncertainty of the applied blank correction. For calibration the radiocarbon ages were translated into calendar ages using the software package OxCal4 [24] and the Intcal20 dataset [25]. The radiocarbon age of sample PSN2-S10E13 was determined to be 3609 ± 29 BP (BP = AD 1950) corresponding to 1659 ± 29 BC.

2.3 XRD and FTIR studies

Powder X-ray diffraction patterns were collected using a STOE powder diffractometer with Bragg Brentano geometry and Cu-Kα1 radiation with an asymmetric Ge 111 monochromator. The diffraction patterns of material removed from the bulk of both ancient pottery fragments is shown in Figure 3. Diffraction was done in the 2-theta range 3–70° for samples PSN2-S10E13 and 5412-S6E15.

Figure 3: 
X-ray powder diffraction patterns of bulk material of ceramic samples, top: PSN2-S10E13, bottom 5412-S6E15. Labelled reflections indicate minerals.
Figure 3:

X-ray powder diffraction patterns of bulk material of ceramic samples, top: PSN2-S10E13, bottom 5412-S6E15. Labelled reflections indicate minerals.

Infrared (IR) spectroscopy was performed using an N2-purged Bruker FTIR Vertex 70 spectrometer equipped with a RT-DLaTGS detector. MIR-FIR powder infrared absorption spectra with instrumental resolution of 2 cm−1 were measured using KBr pellets. Figure 4 shows IR spectra of material extracted from the shards bulk and displays the absorbance of PSN2-S10E13 and for comparison that of 5412-S6E15.

Figure 4: 
IR absorbance of the bulk material of shards PSN2-S10E13 (this study) and 5412-S6E15. The upper spectrum (a) shows the entire range measured, the lower spectrum (b) shows the expanded region towards the far infrared.
Figure 4:

IR absorbance of the bulk material of shards PSN2-S10E13 (this study) and 5412-S6E15. The upper spectrum (a) shows the entire range measured, the lower spectrum (b) shows the expanded region towards the far infrared.

2.4 X-ray tomographic microscopy (XTM)

X-ray Tomographic Microscopy (XTM) measurements were performed at beamline BL1.2W at the Synchrotron Light Research Institute (Public Organization), Nakhon Ratchasima, Thailand. The storage ring energy was 1.2 GeV, generating a ring current ranging between 150 and 80 mA. In order to get a monochromatic X-ray energy of 11.5 keV for X-ray imaging and reduced scattering as well as beam hardening artifacts, a polychromatic X-ray attenuator made of 400 μm thick aluminum foil was used. The X-ray projectors of each pottery shard were collected for 180° with 0.1° angular increment to create a dataset. X-ray images were taken by 2560 × 2160 pixels CMOS camera (PCO.EDGE 5.5, Optique Peter, Lentilly, France) with 3.61-μm resolution. The data were pre-processed and reconstructed in 3D based on a filtered-back projection algorithm using Octopus Reconstruction software (Octopus 8.9.4, TESCAN-XRE, Ghent, Belgium). Volume visualization, sculpting, and transfer function blending of the 3D XTM was processed using the open source software Drishti version 2.6.4 [26]. The value-gradient magnitude (2D) histogram was used for setting different transfer functions [26, 27], including the lowest density peak representing pore topology, middle density peak representing clay pottery fabric and highest density peak representing mineral inclusions (Figures 5 and 6c–e). Transfer functions map voxel information to optical properties and allow to distinguish between homogeneous regions and their neighbourhood by using colour information to indicate volumetric data, respectively density gradients [26]. Hence, these transfer functions were digitally coloured green (pores), dark yellow to dark brown (clay particle fraction) and red (mineral inclusions). The volume analysis of pores and mineral inclusions was performed using the Octopus image program.

Figure 5: 
XTM volume analysis of shard PSN2-S10E13 using the Drishti program, using different transfer functions. (a) Original ceramic, (b) micro-tomographic surface, (c) pottery fabric (clay matrix) in the middle absorption region, (d) pores topology in the low absorption region, (e) mineral inclusions in the high absorption region. The different transfer functions were selected in specific regions of the value-gradient magnitude (2D) histogram presented in the bottom row.
Figure 5:

XTM volume analysis of shard PSN2-S10E13 using the Drishti program, using different transfer functions. (a) Original ceramic, (b) micro-tomographic surface, (c) pottery fabric (clay matrix) in the middle absorption region, (d) pores topology in the low absorption region, (e) mineral inclusions in the high absorption region. The different transfer functions were selected in specific regions of the value-gradient magnitude (2D) histogram presented in the bottom row.

Figure 6: 
XTM volume analysis of pottery shard 5412-S6E15 using the Drishti program, using different transfer functions. (a) Original ceramic, (b) micro-tomographic surface, (c) pottery fabric in the middle absorption region, (d) pores topology in the low absorption region, (e) mineral inclusions in the high absorption region. The different transfer functions were selected in specific regions of the value-gradient magnitude (2D) histogram presented in the bottom row.
Figure 6:

XTM volume analysis of pottery shard 5412-S6E15 using the Drishti program, using different transfer functions. (a) Original ceramic, (b) micro-tomographic surface, (c) pottery fabric in the middle absorption region, (d) pores topology in the low absorption region, (e) mineral inclusions in the high absorption region. The different transfer functions were selected in specific regions of the value-gradient magnitude (2D) histogram presented in the bottom row.

3 Discussion

From burial contexts, non-burial artefacts, cross-dating, samples excavation depths and their surface appearance Ban Chiang pottery had been arranged chronologically [4, 20, 28, 29]. Macroscopic and local chemical and physical probes can be used on ancient pottery to identify its mineralogical composition in order to determine which raw materials were used in certain ceramic bodies of ancient potters and other characteristic features of artefacts produced at different times. Ban Chiang artefacts from the Bronze Age (ca. 1000–300 BC) like sample 5412-S6E15 have a design free surface of yellow-brown colour. It was stated [30], [31], [32] that changes in the ceramic assemblages may indicate cultural changes in the historical course. A drastic change can be observed between Pre-metal Age and younger pottery from Ban Chiang. The surface of the Pre-metal Age shard PSN2-S10E13 with an age of 3609 ± 29 BP (BP = AD 1950) corresponding to 1659 ± 29 BC appears dark-grey to black and the outside of the ceramic shard shows cord-marks (Figure 1a). Mechanically applied cord marks have been found in ceramics from archeological sites at different continents [15], [16], [17], [18], [19]. Cord marks in Figure 1a cross under approximately 70° and were assumed to be made with cord or clay coils pressed or rolled onto the surface.

The raw material for ancient ceramic production in Ban Chiang seems to consists mainly of local clay mineral resources in all periods. Upon firing, they undergo characteristic reactions such as dehydroxylation, decomposition, or transformation. Figure 2 shows a thin section of the bulk of the Pre-metal Age sample with some quartz grains and a few feldspar micro-particles. Feldspar is expected to undergo hydrothermal alteration which converts it to clay during weathering. Despite weathering of sample PSN2-S10E13 over a period of ca. 3600 years, feldspar X-ray diffraction signals are still visible (Figure 3). The mineral might have been added to the matrix as temper to protect the pottery from cracking. FTIR indicates dominating quartz absorbance among other signals of phyllosilicates, alkali feldspar and weak absorption signals of mullite in the bulk material of PSN2-S10E13. IR absorbance spectra of bulk material of both shards (Figure 4) show several superimposed IR excitations of minerals they contain. Because the number of bands is a complex superposition of signals from minerals and impurities of very different quantities, they were neither de-convoluted nor quantitatively evaluated in this work. The band positions of strongest signals and shoulders observed in both shards were assigned qualitatively and are listed in Table 2. Strong IR absorbance peaks result from A2 and E modes of quartz (near 370 cm−1, 397 cm−1, 460 cm−1, 495 cm−1, 694 cm−1, 778 cm−1, 797 cm−1, 1083 cm−1 and 1168 cm−1) [33, 34]. Further bands which appear in both spectra come from A1 and E1 modes of alkali feldspars (albite, K-feldspar) as shoulders near 329 cm−1, 580 cm−1 and an intensive band near 1045 cm−1 [35, 36] as well as signals of possibly different phyllosilicates (muscovite, biotite, montmorillonite) around 3680 cm−1 (OH – stretching), 3440 cm−1 (adsorbed water), 1636 cm−1 (adsorbed water), 1035 cm−1 (ν Si–O–Si), 622 cm−1 (ν (Si–O–Si)), 526 cm−1 ((δ (Si–O–Si))) and 469 cm−1 ((δ (Si–O–Si))) [37, 38]. Phyllosilicate signals shown in Figure 4a are weaker in shard PSN2-S10E13 than those of 5412-S6E15. The IR spectrum of the Pre-metal Age shard PSN2-S10E13 shows weak and broad bands of mullite near 565 cm−1, 740 cm−1 and 870 cm−1 [39] which is confirmed by XRD reflections of mullite displayed in Figure 3.

Table 2:

Infrared band positions observed in bulk material of shard PSN2-S10E13 and 5412-S6E15.

Observed minerals and modes (cm−1) Shard 5412-S6E15 Shard PSN2-S10E13
Mullite [39] +
565 F1u +
740 F2 +
870 A1 +
Quartz [33, 34] + +
370 A2 + +
397 E + +
460 E + +
495 E + +
694 E + +
778 A2 + +
797 E + +
1083 E + +
1168 E + +
Alkali feldspar [35, 36] + +
329 Au + +
510 Au + +
1045 Au + +
Phyllosilicates (A1, E1) [37, 38] + +
469 + +
526 + +
622 + +
1035 + +
1636 + +
3440 + +
3680 + +

While the material of the Bronze Age shard 5412-S6E15 was found to contain some amount of calcite at the surface [12], no calcite could be detected in the bulk of both samples investigated in this work. Dominating X-ray powder diffraction intensities of quartz occur in the bulk of both shards (Figure 3). Sample PSN2-S10E13 shows additional weak signals of a minor phase of mullite near 2-theta 23.2°, 40.9° and 60.6° as well as those of alkali-feldspar.

The 3D image of pottery shards from both periods show their surface details and inner structural details on a macroscopic scale. The outer surface of PSN2-S10E13 presents an image of exterior press pattern of the pottery by hand-made rolling instruments (Figure 5b). The inner volume was fractionated by selecting the specific area density of the 2D histogram. The clay pottery (Figure 5c) appears to be heterogeneous, however PSN2-S10E13 seems to show oriented ceramic texture probably due to hand forming (Figure 5c). The pottery shard 5412-S6E15 (Figure 6c) has a less pronounced orientation of fine particles than PSN2-S10E13.

The pores topology presented by green colour in the image volume fractionated from the lowest density regions (Figures 5d and 6d) display significant differences between shards of the two periods. The PSN2-S10E13 shard contains a higher amount of pore areas than that of the 5412-S6E15 shard whereby the pore size in both shards can reach 100 μm. Table 3 shows the volume percentage of pores for each shard where the PSN2-S10E13 shard has approximately 1.23 times higher pores quantity than the 5412-S6E15 shard.

Table 3:

Volume percentage and porosity of both samples using the Octopus software.

Sample Pores volume percentage Mineral inclusion percentage
PSN2-S10E13 3.09 ± 0.44 0.15 ± 0.03
5412-S6E15 2.51 ± 0.49 0.21 ± 0.10

Mineral inclusions fractioned from the highest density of the 2D histogram are represented by red spots (Figures 5e and 6e). Micro-meter size crystalline particles show slightly different quantity in both shards. The PSN2-S10E13 sample contains a slightly smaller quantity of mineral particles than that of the 5412-S6E15 shard indicated by the corresponding volume percentage deduced from Octopus frame data (Table 3).

XTM studies indicate that the PSN2-S10E13 shard had a higher pore percentage than 5412-S6E15 shard. This finding supports previous work [20] that the PSN2-S10E13 shard was fired at higher temperatures than the 5412-S6E15 shard causing larger pores from the high thermal expansion and sintering of clay minerals. This is in good agreement with the voids around clay agglomerates and cavities shown in Figure 2a as well as the appearance of mullite peaks in the XRD pattern of PSN2-S10E13 (Figure 3) which are expected to appear in such a mixture of minerals at firing temperatures above 1350 K [40, 41].

The production of shards studied in this work happened in a large interval of time and in a limited geographical area. The appearance and texture of both artefacts deviates heavily and such differences had been discussed in literature e.g. see Santacreu [42] and references therein. Technological changes of pottery production might have been promoted by cultural contacts or migration, functional aspects or technological progress like a more economical firing method. It should be noted that the abandonment of the cord marks was followed by a period of pottery without surface design and more than 1000 years later in the Iron Age replaced by hematite-red painted patterns. It is still unknown if the pottery with cord mark patterns or that with coloured pictorial patterns was made to transfer information about their place of origin, however the origin of the respective pottery investigated in this work is well recognizable.

4 Conclusions

Major differences between the Bronze Age shard (5412-S6E15) and the Pre-metal Age shard (PSN2-S10E13) can be seen by the naked eye. It is the cord marked surface and dark grey colour of the ca. 3600 years old sample. Another difference elucidated using chemical and mineralogical investigations is apart from its higher content of quartz (Table 1) and its slightly lower content of phyllosilicates (Figure 4a) the presence of mullite in the Pre-metal Age shard. Optical and X-ray tomographic microscopy yield a larger porosity of sample PSN2-S10E13. The latter two findings indicate higher firing temperature (>1350 K) of sample PSN2-S10E13 compared with the younger Bronze Age shard. We do not know if the cord mark pattern at the surface of the Pre-metal Age pottery can lead to recognizability of other samples prepared in Ban Chiang during that period or in other words, it is not clear yet if the producers intended to convey group-specific aestheitc “meaning” by their patterns [43]. A comparison with cord marked pottery from other places should help to provide a better understanding of the transition of prehistoric societies from the Neolithic to the Bronze Age and beyond.


Corresponding author: Ulrich Bismayer, Cluster of Excellence, Understanding Written Artefacts, Universität Hamburg, Warburgstr. 26, 20354 Hamburg, Germany; and Mineralogisch-Petrographisches Institut, Universität Hamburg, Grindelallee 48, 20146 Hamburg, Germany, E-mail:

Funding source: Khon Kaen University

Award Identifier / Grant number: Unassigned

Funding source: Universität Hamburg

Award Identifier / Grant number: Unassigned

Funding source: Synchrotron Light Research Institute

Award Identifier / Grant number: Unassigned

Acknowledgments

The authors acknowledge experimental support from S. Heidrich, J. Ludwig and P. Stutz. U.B. thanks B. Mihailova, R. Vinx, T. Malcherek and J. Schlüter for helpful discussion. We are thankful to Chr. Hamann, Christian-Albrechts-University, Kiel for AMS measurements and S. Srilomsak, Department of Physics, Faculty of Science, Khon Kaen University, Khon Kaen as well as K. Suriyatham from The Fine Arts Department, National Museum, Ban Chiang, Thailand for help with the samples. The research was conducted within the scope of the Centre for the Study of Manuscript Cultures (CSMC) at Universität Hamburg where U.B. appreciates valuable discussion with M. Friedrich and C. Berns (CSMC). W. Tanthanuch, S. Tancharakorn, C. Rojviriya acknowledge support from the Synchrotron Light Research Institute (Public Organization), Thailand.

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This research was funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy – EXC 2176 ‘Understanding Written Artefacts: Material, Interaction and Transmission in Manuscript Cultures’, project no.390893796 and supported by the Synchrotron Light Research Institute (Public organization), Thailand.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

1. Douglas, M. T., Pietrusewsky, M. V. Vulcan: skilled village craftsman of Ban Chiang, Thailand. In The Bioarcheology of Individuals; Stodder, A. L. W., Palkovich, A. M., Eds. University Press of Florida: Gainsville, 2012; pp. 177–192. https://www.academia.edu/3475996/Vulcan_skilled_village_craftsman_of_Ban_Chiang_Thailand.10.5744/florida/9780813038070.003.0012Search in Google Scholar

2. White, J. C., Vernon, W. W., Fleming, S. J., Glanzman, W. D., Hancock, R. G. V., Pelcin, A. Preliminary cultural implications from initial studies of the ceramic technology at Ban Chiang. Bull. Indo-Pacific Prehist. Assoc. 1991, 11, 188–203.10.7152/bippa.v11i0.11384Search in Google Scholar

3. Higham, C. F. W., Douka, K., Higham, T. F. G. A new chronology for the Bronze Age of northeast Thailand and its implications for southeast Asian prehistory. PloS One 2015, 10, 1–20. https://doi.org/10.1371/journal.pone.0137542.Search in Google Scholar PubMed PubMed Central

4. White, J. C. A Revision of the Chronology of Ban Chiang and its Implications for the Prehistory of Northeast Thailand. Ph.D. Thesis, Department of Anthropology, University of Pennsylvania, USA, 1986. https://repository.upenn.edu/edissertations/4285/.Search in Google Scholar

5. Tabakh, M. E., Utha-Aroon, C., Schreiber, B. C. Sedimentology of the cretaceous maha sarakham evaporites in the Khorat Plateau of northeast Thailand. Sediment. Geol. 1999, 123, 31–62. https://doi.org/10.1016/S0037-0738(98)00083-9.10.1016/S0037-0738(98)00083-9Search in Google Scholar

6. Saminpanya, S., Duangkrayom, J., Jintasakul, P., Hant, R. Petrography, mineralogy and geochemistry of cretaceous sediment samples from western Khorat Plateau, Thailand, and considerations on their provenance. J. Asian Earth Sci. 2014, 83, 13–34. https://doi.org/10.1016/j.jseaes.2014.01.007.Search in Google Scholar

7. Dew, R. E. C., King, R., Collins, A. S., Morley, C. K., Arboit, F., Glorie, S. Stratigraphy of deformed permian carbonate reefs in Saraburi Province, Thailand. J. Geol. Soc. 2018, 175, 163–175. https://doi.org/10.1144/jgs2017-035.Search in Google Scholar

8. Horiuchi, Y., Charusiri, P., Hisada, K. Identification of an anastomosing river system in the early cretaceous Khorat Basin, northeastern Thailand, using stratigraphy and paleosols. J. Southeast Asian Earth Sci. 2012, 61, 62–77. https://doi.org/10.1016/j.jseaes.2012.08.0229.Search in Google Scholar

9. King, C., Tayles, N., Higham, C., Strand-Vidarsdottir, U., Bentley, R. A., Macpherson, C. G., Nowell, G. Using isotopic evidence to assess the impact of migration and the two-layer hypothesis in prehistoric northeast Thailand. Am. J. Biol. Anthropol. 2015, 158, 141–150. https://doi.org/10.1002/ajpa.22772.Search in Google Scholar PubMed

10. Lerner, M., Kossak, S. The arts of south and southeast asia. Metrop. Mus. Art Bull. 1994, LI, 1–92.10.2307/3269200Search in Google Scholar

11. Van Esterik, P. Symmetry and symbolism in Ban Chiang painted pottery. J. Anthropol. Res. 1979, 35, 495–508. https://doi.org/10.1086/jar.35.4.3629544.Search in Google Scholar

12. Bismayer, U., Srilomsak, S., Treekamol, Y., Tanthanuch, W., Suriyatham, K. Artefacts from Ban Chiang, Thailand: pottery with hematite-red geometric patterns. Z. Kristallogr. 2020, 235, 559–568. https://doi.org/10.1515/zkri-2020-0076.Search in Google Scholar

13. Henshilwood, C. S., d’Errico, F., Watts, I. Engraved ochres from the middle stone age levels at blombos cave, South Africa. J. Hum. Evol. 2009, 57, 27–47. https://doi.org/10.1016/j.jhevol.2009.01.005.Search in Google Scholar PubMed

14. Rifkin, R. F. Processing ochre in the Middle Stone Age: testing the inference of prehistoric behaviours from actualistically derived experimental data. J. Anthropol. Archaeol. 2012, 31, 174–195. https://doi.org/10.1016/j.jaa.2011.11.004.Search in Google Scholar

15. Liou, Y. S., Liu, Y. C., Huang, H. Y. Micro‐Raman spectroscopic study of cord‐marked pottery decorated with red coatings from Taiwan, ca 2600–1700 B.C. J. Raman Spectrosc. 2010, 42, 1062–1068. https://doi.org/10.1002/jrs.2813.Search in Google Scholar

16. Quimby, G. I. Cord marked versus fabric impressing of woodland pottery. Am. Antiq. 1961, 26, 426–428. https://doi.org/10.2307/277411.Search in Google Scholar

17. Nieuwenhuyse, O. P., Daskiewicz, M., Schneider, G. Investigating late neolithic ceramics in the northern levant: the view from shir. Levant 2020, 52, 15–33. https://doi.org/10.1080/00758914.2018.1453213.Search in Google Scholar

18. Baldi, J. S. Between and end and a new beginning. Paleorient 2022, 48-1, 41–72. https://doi.org/10.4000/paleorient.1539.Search in Google Scholar

19. Epossi Ntah, Z. L., Sobott, R., Fabbri, B., Bente, K. Characterization of some archaeological ceramics and clay samples from Zamala – far-northern part of Cameroon (West Central Africa). Ceramica 2017, 63, 413–422. https://doi.org/10.1590/0366-69132017633672192.Search in Google Scholar

20. Tanthanuch, W., Pattanasiriwisawa, W., Somphon, W., Srilomsak, S. Synchrotron studies of Ban Chiang ancient pottery. Suranaree J. Sci. Technol. 2011, 18, 15–28.Search in Google Scholar

21. Beirau, T., Paulmann, C., Bismayer, U. Recrystallisation of metamict allanite. Mineral. Mag. 2011, 75, 2393–2399. https://doi.org/10.1180/minmag.2011.075.4.2393.Search in Google Scholar

22. Potts Philip, J., Tindle Andrew, G., Isaacs, M. C. On the precision of electron microprobe data; a new test for the homogeneity of mineral standards. Am. Mineral. 1983, 68, 1237–1242.Search in Google Scholar

23. Stuiver, M., Polach, H. Discussion reporting of 14C data. Radiocarbon 1977, 19, 355–363. https://doi.org/10.1017/S0033822200003672.Search in Google Scholar

24. Ramsey, C. B., Lee, S. Recent and planned developments of the program OxCal. Radiocarbon 2013, 55, 720–730. https://doi.org/10.1017/S0033822200057878.Search in Google Scholar

25. Reimer, P., Austin, W., Bard, E., Baylis, A., Blackwell, P., Ramsey, B., Butzin, M., Cheng, H., Edwards, R. L., Friedrich, M., Grootes, P. M., Guilderson, T. P., Hajdas, I., Heaton, T. J., Hogg, A. G., Hughen, K. A., Kromer, B., Manning, S. W., Muscheler, R., Palmer, J. G., Charlotte Pearson, C., Van der Plicht, J., Reimer, R. W., Richards, D. A., Scott, E. M., Southon, J. R., Turney, C. S. M., Wacker, L., Adolphi, F., Büntgen, U., Capano, M., Fahrni, S. M., Fogtmann-Schulz, A., Friedrich, R., Köhler, P., Kudsk, S., Miyake, F., Olsen, J., Reinig, F., Sakamoto, M., Sookdeo, A., Talamo, S. The IntCal20 northern hemisphere radiocarbon age calibration curve (0–55 cal kBP). Radiocarbon 2020, 62, 725–757. https://doi.org/10.1017/RDC.2020.41.Search in Google Scholar

26. Limaye, A. Drishti, a volume exploration and presentation tool. In Proc. SPIE; Stock, S. R., Ed.; Developments in X-Ray Tomography VIII, 2012, p. 85060X-1.10.1117/12.935640Search in Google Scholar

27. Barron, A., Denham, T. A micro CT protocol for the visualisation and identification of domesticated plant remains within pottery sherds. J. Archaeol. Sci. Rep. 2018, 21, 350–358. https://doi.org/10.1016/j.jasrep.2018.07.024.Search in Google Scholar

28. White, J. C. Changing paradigms in southeast Asian archaeology. J. Indo-Pac. Archaeol. 2017, 41, 66–77; https://doi.org/10.7152/jipa.v41i0.15024.Search in Google Scholar

29. Higham, C. F. W. In the Ban Chiang Culture in Wider Perspective, Vol. 69; British Academy: London, 1984; pp 229–261. https://searchworks.stanford.edu/view/1643426.Search in Google Scholar

30. Boyd, W. E., Chang, N. Integrating social and environmental change in prehistory: a discussion of the role of landscape as a heuristic in defining prehistoric possibilities in NE Thailand. In Terra Australis 21, Altered Ecologies: Fire, Climate and Human Influence on Terrestrial Landscapes; AANU E Press: Canberra, Australia, 2010, pp. 273–297. https://www.jstor.org/stable/j.ctt24h8rj.17.10.22459/TA32.11.2010.15Search in Google Scholar

31. Garcea, E. A. A., Hildebrand, E. A. Shifting social networks along the nile: middle holocene ceramic assemblages from sai island, Sudan. J. Anthropol. Archaeol. 2009, 28, 304–322. https://doi.org/10.1016/j.jaa.2009.05.002.Search in Google Scholar

32. Santacreu, D. A. Change and stability in pottery production. In Materiality, Techniques and Society in Pottery Production; De Gruyter Open: Warsaw Berlin, 2014, pp. 262–278.Search in Google Scholar

33. Etchepare, J., Merian, M., Smetankine, L. Vibrational normal modes of SiO2. I. α and β quartz. J. Chem. Phys. 1974, 60, 1873–1876. https://doi.org/10.1063/1.1681287.Search in Google Scholar

34. He, M., Yan, W., Chang, Y., Liu, K., Liu, X. Fundamental infrared absorption features of α-quartz: an unpolarized single-crystal absorption infrared spectroscopic study. Vib. Spectrosc. 2019, 101, 52–63. https://doi.org/10.1016/j.vibspec.2019.02.003.Search in Google Scholar

35. Zhang, M., Salje, E. K. H., Carpenter, M. A., Parsons, I., Kroll, H., Reed, S. J. B. Graeme-Barber A. Exsolution and Al–Si disorder in alkali feldspars: their analysis by infrared spectroscopy. Am. Mineral. 1997, 82, 849–857. https://doi.org/10.2138/am-1997-9-1002.Search in Google Scholar

36. Reitze, M. P., Weber, I., Kroll, H., Morlok, A., Hiesinger, H., Helbert, J. Mid-infrared spectroscopy of alkali feldspar samples for space application. Mineral. Petrol. 2020, 114, 453–463. https://doi.org/10.1007/s00710-020-00709-9.Search in Google Scholar

37. Singha, M., Singh, L. Vibrational spectroscopic study of muscovite and biotite layered phyllosilicates. Indian J. Pure Appl. Phys. 2016, 54, 116–122. http://nopr.niscpr.res.in/handle/123456789/33795.Search in Google Scholar

38. Vaculikova, L., Plevova, E. Identification of clay minerals and micas in sedimentary rocks. Acta Geodyn. Geomater. 2005, 2, 167–175.Search in Google Scholar

39. MacKenzie, K. J. D. Infrared frequency calculations for ideal mullite (3Al2O3. 2SiO2). J. Am. Ceram. Soc. 1972, 55, 86–71. https://doi.org/10.1111/j.1151-2916.1972.tb11211.x.Search in Google Scholar

40. Park, K. S., Milke, R., Efthimiopoulos, I., Pausewein, R. R., Reinhold, S. Pyrometamorphic process of ceramic composite materials in pottery production in the Bronze/Iron Age of the Northern Caucasus (Russia). Sci. Rep. 2019, 9, 1–14. https://doi.org/10.1038/s41598-019-47228.Search in Google Scholar

41. Gao, Y., Liu, W., Song, X., Liu, Q., Yao, S., Wang, J., Cai, Q., Ma, Y. Preparation, characterization and mechanical properties of continuous mullite fibers derived from the diphasic sol-gel route. J. Sol-Gel Sci. Technol. 2019, 92, 75–83. https://doi.org/10.1007/s10971-019-05078-0.Search in Google Scholar

42. Santacreu, D. A. Change and stability in pottery production. In Materiality, Techniques and Society in Pottery Production; DeGruyter Open Poland: Warsaw, 2014; pp. 262–278.10.2478/9783110410204.21Search in Google Scholar

43. Iliopoulos, A. Early body ornamentation as Ego-culture: tracing the co-evolution of aesthetic ideals and cultural identity. Semiotica 2020, 232, 187–233. https://doi.org/10.1515/sem-2019-0073.Search in Google Scholar

Received: 2023-04-10
Accepted: 2023-06-20
Published Online: 2023-07-03
Published in Print: 2023-07-26

© 2023 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/zkri-2023-0015/html
Scroll to top button