When information gatherers use methods that cross a legal or ethical threshold?

Although the first ancient DNA molecules were extracted more than three decades ago, the first ancient nuclear genomes could only be characterized after high-throughput sequencing was invented. Genome-scale data have now been gathered from thousands of ancient archaeological specimens, and the number of ancient biological tissues amenable to genome sequencing is growing steadily. Ancient DNA fragments are typically ultrashort molecules and carry extensive amounts of chemical damage accumulated after death. Their extraction, manipulation and authentication require specific experimental wet-laboratory and dry-laboratory procedures before patterns of genetic variation from past individuals, populations and species can be interpreted. Ancient DNA data help to address an entire array of questions in anthropology, evolutionary biology and the environmental and archaeological sciences. The data have revealed a considerably more dynamic past than previously appreciated and have revolutionized our understanding of many major prehistoric and historic events. This Primer provides an overview of concepts and state-of-the-art methods underlying ancient DNA analysis and illustrates the diversity of resulting applications. The article also addresses some of the ethical challenges associated with the destructive analysis of irreplaceable material, emphasizes the need to fully involve archaeologists and stakeholders as part of the research design and analytical process, and discusses future perspectives.

Introduction

In 1984, short DNA fragments were extracted and sequenced from the dried muscle of a museum specimen of the quagga, a species of zebra that became extinct at the beginning of the twentieth century, marking the birth of ancient DNA [aDNA] research1. Although the data gathered were limited to 229 bp, this was the first time that direct genetic information had travelled through time, adding molecular evidence to the toolkit used by researchers to observe evolution and understand how the modern world came to be.

More than three decades later, the time range amenable to DNA analysis extends to more than half a million years [560,000–780,000 years ago2], and many extinct species have now had their genomes completely sequenced, including woolly mammoths3 and cave bears4, as have human populations from Vikings5 to Paleo-Inuit6 to Neanderthals7,8,9,10. With the addition of aDNA data, our current atlas of genetic variation is not limited to a snapshot of the diversity found in present-day populations across the world. Instead, it is continuously enriched with temporal information tracking changes in the genetic ancestries of human, animal, plant and even microbial populations as they expanded, collapsed and adapted to new local environmental conditions11,12,13.

aDNA has led to the discovery of new branches within the human family tree, including that of the Denisovans, who are close relatives of Neanderthals14,15,16. As a result, the genomic consequences of population decline17,18,19 and the underlying environmental20,21,22 and/or anthropogenic23,24 drivers of extinctions have been revealed and clarified. Applying aDNA techniques to archaeozoological and archaeobotanical remains has also considerably enhanced our understanding of the transition from hunting and gathering to herding and farming, including how past human groups domesticated wild species showing preferred characteristics12. Domestication provided new opportunities for zoonotic transfer of animal pathogens to humans, which have also been analysed using ancient genomic investigation13. More generally, aDNA data have transformed our understanding of the historical emergence, virulence and spread of major infectious diseases25. Beyond the genomes of the hosts and their pathogens, the metagenomic characterization of our microbial self26,27 and the identification of epigenetic marks28,29 have paved the way towards a study of ancient holobiomes, which promise to reveal a deeper understanding of past social, dietary and environmental shifts and their impact on the health of individuals and populations.

None of these developments would have been possible without next-generation sequencing [NGS]30, which remains thus far the most transformative technology in the history of aDNA research, profoundly affecting wet-laboratory and dry-laboratory activities alike [Fig. 1]. At its most basic level, the success of NGS lies in its ability to accommodate the sequencing and analysis of millions of loci in parallel from minute amounts of ultrashort DNA fragments31,32,33. As a result, cumbersome PCR-based analyses of individual loci, which were instrumental in initially establishing the field of aDNA analysis, are now almost entirely superseded by genome-scale studies, with the exception of mini-barcodes that remain useful for characterizing past environmental communities34. NGS also importantly allows the study of DNA in its entirety, a cornerstone in modern authentication approaches of aDNA data. In-solution target-enrichment techniques probing specific genomic regions have become instrumental for the cost-effective recovery of sequence data at the gene35, mitogenome36, chromosomal37 and even whole-genome38 scale.

Fig. 1: Analytical milestones in aDNA research.

Key milestones pertaining to wet-laboratory method improvement [part a] or dry-laboratory advances [part b]. aDNA, ancient DNA; LIMS, laboratory information management systems; mtDNA, mitochondrial DNA; NGS, next-generation sequencing; PCA, principal component analysis; PSMC, pairwise sequentially Markovian coalescent; ROH, runs of homozygosity; USER, uracil–DNA–glycosylase [UDG] and endonuclease VIII [Endo VIII] [New England Biolabs].

Full size image

As a complement to more sensitive techniques for extracting39,40,41,42 and integrating DNA into NGS libraries16,43, new sources of archaeological material have broadened the range of aDNA applications. These include dental calculus26,44, wood45,46, mollusc shells47 and sediments48,49, as well as biocultural archives such as parchments50 and textiles51. The discovery that particular bone types exhibit better molecular preservation52,53 has also facilitated the collection of aDNA time-series data at both population and genome-wide scales. NGS has furthermore revolutionized the analysis of ancient sequence data, providing statistical solutions to overcome vexing problems, such as rampant contamination54,55,56 and post-mortem DNA damage57, while still enabling the detection of subtle changes in population structure by means of genetic drift, admixture and/or selection58.

With the addition of new laboratory and computational methods developed in the past decade, aDNA analysis has now come of age. This Primer provides an up-to-date overview of the most commonly used aDNA methods and tools, as well as their limitations, and anticipates future innovations that will help propel us beyond the current state of the art as we continue to advance the scope of molecular archaeogenetics.

Experimentation

Analysing aDNA requires the destruction of irreplaceable, finite subfossil material that is part of humanity’s biocultural heritage. As such, aDNA studies require careful scientific and ethical planning and a commitment to responsible research. Once a research plan is established, sample analysis proceeds similarly for most sample types. Samples are taken to a dedicated clean laboratory facility, and DNA is typically freed from its parent material by a combination of demineralization and digestion. Often, samples have high levels of degradation and are prone to contamination by other DNA sources. Following purification and concentration steps, aDNA is then constructed into an NGS library and sequenced. In addition to these core steps, there are also optional steps that can be performed at various stages to achieve desired effects, such as the complete or partial removal of post-mortem damage prior to library construction or the enrichment of specific genetic targets prior to sequencing.

Material types

Early aDNA studies considered macroscopic preservation a predictor of molecular preservation and consequently focused on soft tissues from naturally or artificially mummified remains and from stuffed or fluid-preserved museum specimens1,59. Hair shafts provided the DNA source for the first successfully sequenced ancient mammoth60 and human genomes6, in part owing to keratin’s low permeability to contaminants and because it is easier to decontaminate61. Soft tissues, however, rarely preserve and — with the exception of hair — are usually heavily contaminated by environmental microorganisms62. Mineralized tissues are more abundant and typically better preserved than soft tissues, and consequently the focus of more recent aDNA studies has shifted to vertebrate bones and teeth63. The petrous bone has been singled out for retaining a high degree of endogenous DNA preservation52, with ear ossicles64 and teeth providing other suitable alternatives. Although tooth cementum can contain high amounts of host DNA53, dentine is generally preferred for genetic analysis as, in addition to host DNA, it also allows for the recovery of ancient blood-borne pathogens65,66.

In studies of ancient plants, suitable materials for genomic analysis include desiccated, charred, waterlogged or mineralized67 pollen68, cobs69,70,71, pips72, herbarium specimens73 or seeds74,75. Seeds have also been found to preserve RNA76,77. New aDNA reservoirs are also still being discovered in the form of purely biological substrates, such as insects78, feathers79, eggshells80, mollusc shells47 or wood45,46, but also from ‘cultural’ artefacts including livestock skin parchments50 and drinking horns81, pottery82 or birch pitch mastics83,84. Beyond DNA from single focal species, whole communities can also be retrieved from single samples, such as preserved faeces [paleofaeces or coprolites] and calcified dental plaque [calculus], allowing metagenomic analyses of the gut85,86 and oral microbiota26,44, respectively, as well as the detection of pathogens26,44, parasites87 and foods88,89. At larger scales, entire paleoecosystems can be reconstructed from environmental archives, such as sediments22,49,90, ice91 and lake cores34,48,92.

Ethical aDNA research

There are several ethical considerations that researchers must evaluate before embarking on destructive analysis of irreplaceable archaeological material [Fig. 2]. The sampling and destruction of human remains for aDNA research needs to consider cultural, historical and even political implications. Consultation with appropriate local stakeholders, including descendent communities, both prior to the start of a project and during its development, enriches aDNA research and is highly encouraged. This is not limited to human remains but applies to aDNA studies generally, including studies of animals, plants and artefacts.

Fig. 2: Experimental workflow.

A wide range of remains are amenable to ancient DNA [aDNA] analysis. Prior to sample destruction, a research plan should be agreed amongst the different stakeholders. The different wet-laboratory procedures must be carried out in specific aDNA facilities, minimizing environmental contamination, and include all pre-amplification experimental steps, including sample preparation, DNA extraction, optional USER treatment and DNA library construction. Target enrichment and PCR amplification are carried out in regular molecular genetics facilities. Following next-generation sequencing [NGS], the sequence data are processed on computational servers and uploaded to public repositories. Results should be communicated to the stakeholders and any remaining sample should be returned as per the initial agreement. USER, uracil–DNA–glycosylase [UDG] and endonuclease VIII [Endo VIII] [New England Biolabs].

Full size image

Ethical issues with regard to aDNA research encompass conceptualization, sampling and communication. First, the conceptualization of the project should proceed in a way that is mindful of historical, cultural and political realities and that involves communication with local stakeholders [excavators, curators/museums, local communities and religious institutions] on equal footing. This should include detailed communication of the relevance and potential outcomes of the analyses as well as possible risks93,94. Second, the strategy and process of sampling should include sample documentation [for example, by photography and/or surface scanning or computed tomography95], the use of minimally destructive analytical techniques96,97,98,99 and a plan for proper export and storage of samples in accordance with official permissions. The remaining sample material should be returned to the museum or appropriate group, with restoration [or production of a copy] of destroyed parts94,100,101 when requested. When working with rare samples, the methods should first be extensively tested and demonstrated to be successful on similar but more commonly available material [for example, faunal remains at hominin sites], and sample material should not be fully exhausted but at least partially conserved for future research. Third, communication and consultation should be ongoing with stakeholders, including descendant communities. Ideally, this communication and consultation will include joint decision-making and local capacity-building, collaboration on publications and agreement about data management102. Wagner et al.103 make five recommendations for establishing successful collaborations between aDNA researchers and descendant communities: consult formally with communities; address cultural and ethical considerations; engage communities and support capacity-building; develop plans to report results and manage data; and develop plans for long-term responsibility and stewardship. Finally, researchers should be mindful of the language they use in publications and specifically the potential for misinterpretation or offence given by technical terms — such as specimen, admixture or inbreeding coefficient — that may imply different meanings in scientific and general contexts. Researchers should use a reflected vocabulary that avoids biological essentialism104 and ensure that their work avoids falling into the pitfalls of racial, nationalist or simplistic narratives105,106,107 [Box 1].

Box 1 Writing archaeogenetic prehistory

In the past decade, archaeogenetic studies have provided radically new insights into prehistory worldwide. However, several early publications did not fully appreciate the related history of research in archaeology. The inherent complexity of relevant terms such as culture, migration or people in the social sciences was insufficiently acknowledged, in part owing to different disciplinary publishing norms and to the strict limitations on word count and number of references imposed by high-profile journals, which resulted in very negative feedback by the archaeological community. Now, sustainable collaborations between archaeogeneticists, archaeologists and historians have been established and even institutionalized [such as with the Max Planck Harvard Research Center for the Archaeoscience of the Ancient Mediterranean], and awareness has been raised on both sides — for the use of challenging terms, on the one hand, and the potential of archaeogenetic historiography on the other. Increasingly, a new generation of scholars is being trained within an interdisciplinary framework that allows narratives of the past to be told in more integrative, nuanced and sophisticated ways that address the complexity of archaeological and archaeoscientific data sets.

aDNA facilities

Preserved aDNA is often limited in quantity and has a high degree of degradation, which can be compounded by contamination with modern DNA. Under ideal conditions, samples would be collected immediately from excavated individuals, with minimal handling to limit potential contamination from modern sources, including staff and storage facilities. However, this is not possible for the large archives of already excavated remains in museums and institutes around the world. As such, the field of aDNA has developed various techniques to identify, remove or reduce contamination introduced during post-excavation storage and handling40,41,42,43,54,55,56,57.

Regardless of sample origins, the extraction and manipulation of aDNA must be carried out in dedicated clean laboratory facilities to minimize further contamination risks. Such facilities are typically access-regulated and located in buildings separate from those where post-amplification DNA is manipulated. They are maintained as sterile environments through HEPA-filtered positive air pressure systems, UV exposure and daily [bleach] decontamination treatment of bench surfaces108. Anterooms allow researchers to dress in suitable personal protective equipment, including disposable full-body suits, gloves, sleeves, face masks and overshoes. The workspace is generally divided into multiple, separate rooms in which specific experimental tasks can be performed so as to parallelize work while limiting cross-contamination risks. Laboratory equipment is routinely decontaminated before and after use by cleaning with bleach and alcohol, whereas laminar flow hoods, with monitored air extraction and filtering systems, help prevent pollen, powder and aerosol contamination. These strict procedures are necessary to minimize modern DNA entering the facilities through reagents, ventilation and staff personnel.

DNA extraction

In order to maximize preservation of the remains’ integrity and allow potential further molecular or morphological analyses, minimally destructive methods have been proposed to sample bones96, teeth97, insects98 or plants99, for example. Optimization at every experimental and computational step of the aDNA pipeline has significantly increased the sensitivity of aDNA methods, thereby decreasing the amounts of sampled material necessary for successful analyses. Small amounts of preserved DNA and extensive DNA fragmentation to lengths  3] are shown in pink, whereas non-significant scores are shown in blue. Individuals from Britain and associated with the Bell Beaker culture and the Chalcolithic/EBA show a significant excess of genetic sharedness with Bronze Age WSH associated with the Yamnaya culture of the Pontic steppes. c | Principal coordinate analysis of genus-level taxonomic frequency profiles of ancient [circles] and modern [diamonds] microbiota reconstructed from faeces, dental calculus, dental plaque, dentine, bone, soil and sediments [data from ref.353]. Ancient dental calculus samples have metagenomic diversity profiles similar to modern dental calculus, whereas paleofaeces resemble modern faeces from non-industrialized populations. Microorganisms colonizing skeletal material generally originate from soil, but some dentine samples show evidence of being decomposed by the dental plaque bacteria. aDNA, ancient DNA.

Full size image

Population ancestry modelling

Whereas PCA results are most often visualized as biplots and in practice analysed in 2D comparisons of two principal components at a time, alternative clustering methods such as those implemented in ADMIXTURE214 make higher-dimensional assessments possible. There is a broad suite of software that implements an explicit genetic model aimed at partitioning individuals into ancestries, fitted to genotype frequencies214,215,216, that has now been extended to model the temporal structure of the data217. However, these approaches are sensitive to [and thus can be biased by] genetic drift between time periods and differences in relative sample size between ancestries. This contrasts with methods based on f statistics217,218,219, which leverage covariation of allele frequencies between populations and can retrieve unbiased admixture proportions. Although more cumbersome than clustering approaches for initial data exploration, f statistic-based methods can, in many cases, also provide a statistical test for proposed ancestry models.

Three individual statistics are central to the f-statistic framework. f2 statistics measure distances between pairs of populations and/or individuals. f3 statistics can reveal unambiguous evidence for situations in which one individual [or population] is formed by the admixture of two others219 but can also provide a measure of shared genetic drift when an outgroup is included220. Finally, f4 statistics test for asymmetries between populations that are indicative of gene flow219 [Fig. 4b]. If a five-population history of a specific topology can be assumed, a ratio of two f4 statistics can be used to estimate ancestry proportions unbiased by genetic drift219. In more generalized developments of the f-statistics approach in the tool qpAdm [and qpWave], the predicted f statistics of a one-source221 or two to four-source158,210 ancestry model can be obtained if a set of reference populations or genomes can be posited that are more distantly related to the ancestry sources than populations defined by users to represent sources. This approach then provides a p value for the ancestry model tested and also, for valid models, an unbiased ancestry proportion estimation.

Genotype imputation

With some exceptions, ancient genomes and genome-wide enrichment of SNPs are sequenced at limited depth of coverage, which precludes the determination of genotypes. Statistical inference of missing genotypes is possible using a process called genotype imputation, which assumes that the distribution of haplotypes present in the population is known and that sufficient coverage is available, typically around 1× [ref.222]. In practice, the haplotype distribution is often approximated from large-scale reference panels such as the Haplotype Reference Consortium223 for human populations. This may be adequate for relatively recent historical time periods224 and/or areas such as Europe where many genomes from modern human populations have been collected222 but may be underpowered in cases of limited reference panels, as is true for many populations from elsewhere in the world225. Imputation can help genotype single loci — such as lactase persistence226 — to full genomes52,222,227 and estimate runs of homozygosity and inbreeding coefficients from low-coverage data227. Promisingly, new approaches relying on linkage information from reference panels of modern haplotypes that are compatible with a minimum of 0.3× coverage data are now emerging228. Patterns of non-random allele association within haplotypes can also improve the inference resolution of population structure and outperform methods using unlinked variation222,229. Nevertheless, imputation can also infer false genotype calls230, and so care must be exercised when using imputation to investigate the evolutionary history of specific traits.

Microbiota profiling

In addition to containing genomic fragments from the focal species, aDNA extracts also generally play host to an entire metagenomic diversity of environmental microorganisms that mainly colonized the subfossil material after death7,231. The presence of often dominant non-host DNA templates within aDNA libraries can considerably increase genome sequencing costs but can be filtered out using target-enrichment techniques or enzymatic digestion of over-represented bacterial sequence motifs7. In studies of pathogens [reviewed elsewhere13] and oral, gastric and faecal microbiota, however, the metagenomic content itself can be the intended research target if pathological lesions, dentine, dental calculus26, stomach contents231,232 or coprolites233,234,235 are analysed.

Taxonomic profiling of library metagenomic content is typically carried out using a sequence identity threshold against comparative databases [such as MALT236; Fig. 4c] or k-mer representativity patterns [for example, Kraken237], which break down the frequency distribution of each sequence motif of a predefined length of k nucleotides. The sensitivity and specificity of taxonomic assignment largely depend on the representation characteristics of the comparative database but are robust to post-mortem DNA damage238,239. Curated databases of microbial markers such as MetaPhlAn2 [ref.240] ensure specificity but may lack sensitivity for environmental and/or non-human associated microbial taxa, whereas large uncurated sequence repositories such as the National Center for Biotechnology Information [NCBI] nucleotide database may lack specificity [reviewed elsewhere27]. Several bioinformatic pipelines automating analyses from mapping to statistical profiling are available, such as HOPS241 and metaBit242. In addition, SourceTracker243 and CoproID233 can estimate mixture proportions from known candidate sources, such as soil, gut and oral microbiota. Authentication of microbial taxonomic profiles is difficult and requires cross-validation through different software238 and analyses of specific sequence characteristics, including through mapping against the genomes of candidate species and recovery of post-mortem signatures of molecular degradation27.

Although 16S meta-barcodes are extensively used for profiling modern microbiota, the approach is impracticable on ancient material because of amplification biases introduced during PCR owing to extensive DNA fragmentation133. 16S ribosomal RNA gene amplification is designed to target hypervariable regions in order to distinguish microbial taxa, and these sequences are clustered based on similarity [≥97%] with bacterial taxa. However, most aDNA sequences are smaller than 200 bp in length, the minimum for 16S ribosomal RNA variable region amplification. For example, a 2015 study133 showed that amplified libraries of the 16S V3 region did not accurately reflect microbial taxa within aDNA samples when compared with shotgun metagenomic data. This was, in part, owing to biased amplification of length polymorphisms whereby taxa with deletions were over-represented and taxa with insertions were under-represented133. Shotgun metagenomic approaches are thus recommended.

DNA methylation

Two main experimental methods have been applied to map DNA methylation signatures on ancient genomes. The first builds on methods commonly used on fresh DNA, such as bisulfite conversion244,245 and immunoprecipitation62,167. Bisulfite conversion uses sodium bisulfite to convert unmethylated cytosines into uracils, similar to the C to U conversion that occurs in post-mortem DNA damage. However, bisulfite conversion is also extremely harmful to DNA and, thus, generally not recommended for aDNA, despite a few successes on limited numbers of samples244,245. Immunoprecipitation is likewise extremely limited in practice owing to the fast post-mortem decay of CpG dinucleotides167 and DNA fragmentation62, which reduce the number of potential DNA methylation targets available per fragment.

The second category of methods for aDNA methylation analysis relies on indirect statistical inference based on patterns of post-mortem DNA decay that are ubiquitous in plants246 and animals28,29. The idea is implemented in open-source statistical packages such as epiPALEOMIX247 and DamMet248. It leverages NGS data generated following USER treatment, which maintains C to T misincorporations only at methylated sites29. Although obtaining reliable estimates at single-nucleotide resolution requires impractical sequence coverage [≥80× coverage248], the methodology has been applied successfully to assess DNA methylation levels within genomic regions.

Applications

The application potential of aDNA analysis has a broad scientific scope and relevance to key archaeological, ecological and evolutionary questions. The following section highlights some of the areas that have received heightened scholarly attention, including human population history, plant and animal domestication, and the origins and evolution of pathogens and microbiomes, as well as the impact of global climatic changes on past faunal and floral communities.

Human genomics

No area of aDNA analysis, except that of microbial archaeology13, has benefited as much from NGS as that of past human population genetics [Fig. 5]. The wealth of newly available ancient human genomic data has been instrumental in transforming our understanding of the human past, from a previous narrative of long-term population continuity and isolation to one in which mobility and population mixture have played much more prominent roles11.

Fig. 5: Geographical and temporal distribution of ancient specimens analysed at the genome scale.

a | Humans [data from Ancient Human DNA uMap]. b | Non-human animals. c | Plants. d | Pathogens [data from ref.353]. e | Microbial metagenomes [data from ref.353]. The minimum [square] and maximum [diamond] age of the samples analysed at a given location are provided. Only squares are shown in case the temporal range overlaps only one time interval. Colours correspond to approximately 1,000-year time intervals, except for the oldest that includes genomes dating to 13,000 years ago or longer, including Late and Middle Pleistocene specimens. Size of squares and diamond is proportional to the number of samples analysed at a given location but is adequately scaled in each panel for clarity and according to the total number of specimens [N] for which both geographical and temporal data are available. bp, before present [that is, prior to 1950 ce].

Full size image

The first ancient human genome was sequenced in 2010 from the hair of an approximately 4,000-year-old Saqqaq Paleo-Inuit man from Greenland6. This made the possibility of sequencing of authentic ancient human genomes a reality, something that was until then considered unlikely owing to pervasive sequence contamination in the pre-NGS era249. The draft genomes of two archaic hominins, the Neanderthals7 [1.3× coverage, composite of 3 individuals] and Denisovans15 [1.9× coverage], were released shortly after the high-coverage Saqqaq Paleo-Inuit genome6 [20× coverage]. Both archaic genomes revealed evidence of admixture amongst various hominin lineages8, in which Neanderthals and Denisovans contributed significant ancestry to modern non-African populations and modern populations of Australasia and Oceania, respectively221, contributing to potential phenotypic and health consequences today250.

Since 2010, ancient human genomics has moved forwards with incredible speed, and current publications typically include genome-scale data and/or whole-genome sequences from dozens to hundreds of individuals5,251,252,253. Current understanding of genomic variation within pre-Holocene populations is lagging behind that of more recent populations owing to the scarcity of the fossil record and relatively limited DNA preservation, although partial nuclear genomes from pre-Neanderthal groups193 and genomes from multiple Upper Palaeolithic modern humans37,254,255,256,257 have been successfully recovered. Europe, and more generally west Eurasia, has been the focus of most research258, although significant progress has also been made in understanding the genomic history of other regions, such as East Asia257,259, the Americas6,220, Oceania190,260, Southeast Asia261,262, the circum-Mediterranean263 and Africa264. These other regions include climatic zones that were previously believed not to be suited for long-term DNA preservation.

Studies of west Eurasia have revealed that the genetic diversity of present-day Europeans is primarily formed by three major ancestry components210, consisting first of early hunter–gatherers194,265 who later admixed with farmers leaving Anatolia approximately 9,000 years ago, and who brought with them agricultural innovations such as domesticated crops and livestock and introduced the Neolithic lifestyle into Europe266. A third genomic component was later overlaid following the migration of pastoralists culturally associated with the Yamnaya horizon from the Pontic steppe around 5,000 years ago158,226 [Fig. 4a,b]. It is now believed that it was this pastoralist migration, and not earlier Neolithic population movements, that likely brought the proto-Indo-European language to Europe158,226. This was then followed by the westward expansion of the Bell Beaker complex that is genetically associated with an almost total population replacement of groups in Britain206.

Further to the east, archaeogenomic studies have shown that Central and Inner Asia underwent a population history very different from that of Europe, even though they were impacted by some of the same events. These areas were initially populated by groups of hunter–gatherers who were among the first to domesticate the horse251 and who were related to contemporary Native Americans. However, in contrast to the hunter–gatherers of Europe, these peoples did not experience migration from Anatolia and never underwent the same processes of Neolithization253. Nevertheless, much of Asia has witnessed major migrations since the Bronze Age251,252,253,267, including from Yamnaya-related and subsequent western steppe populations who spread pastoralism across the continent. These population dynamics resulted in Central and Inner Asia shifting from initially being occupied by groups of peoples related to Native Americans, to being occupied by peoples of mixed western Eurasian genetic ancestry and then, finally, to becoming inhabited by peoples with greater Northeast and East Asian ancestry. In the process, the language topology of the region also changed from being largely Indo-Iranian to becoming predominantly Turkic or Mongolic today252,253.

Beyond Eurasia, studies of ancient human genomes have also substantially contributed to our understanding of the population history of the Americas, from its initial peopling and the subsequent dispersal and diversification of indigenous populations268,269,270 to the impact of European contact and colonialism271. Recent studies are also informing the population history of tropical regions previously thought to be beyond the reach of aDNA owing to extensive DNA degradation conditions, such as the Caribbean, which now appears to have been colonized in at least three waves from migrants originating in both North and South America272. The study of ancient human genomes has also contributed to the reconstruction of events for which the historical record is limited or intentionally concealed. One such example is the application of ancient genomics to the study of the Transatlantic Slave Trade, which has allowed the identification of specific places of origin of African-born enslaved individuals who were forcibly taken from Africa to the Americas, revealing the diversity of their ethno-geographical origins within West Africa and beyond, as well as aspects of their lives in the Americas273,274.

Domestication

The domestication of plants and animals represents a key stage in human history and has long been an area of unresolved debate about the timing, location and number of domestication sources275. Pre-NGS era studies identified key loci under early selection in crops such as maize69 and pigs276, providing glimpses into the process as well as into their likely routes of spread. Now, domestication models are being rewritten as genomic data reveal a dynamic and complex history that includes complete population turnover, adaptive introgression between wild and domesticated forms, an unexpected temporal and spatial distribution of origin and selection, and a larger role for natural dispersal processes [see12,277 for recent reviews on these topics].

Large-scale population replacement occurred in dogs278, wolves196,279,280, horses148,281, pigs282 and crops such as potato283. For example, indigenous American dogs were completely replaced by later European dogs278 with some introgression from Inuit dogs284, and modern domestic horses did not arise from the Eneolithic Botai culture of central Kazakhstan as previously thought281 but from a later Bronze Age expansion148. By contrast, a staggering level of constancy is evident in grapes with more than 900 years of uninterrupted vegetative propagation72.

Adaptive introgression has emerged as a recurrent feature in response to past environmental dynamics. For example, maize assimilated wild adaptive variation enabling northward expansion285, as did flax in Europe286. Cattle were introgressed by wild aurochs287 and with zebus, thereby gaining alleles adaptive for response to drought288. Such findings overturn the concept of domestication as a process of isolation between wild and domesticated populations, and suggest instead complex instances of gene flow, as is evident in the mosaic ancestry of ancient goats289 and in the multiple emergence events of semi-domesticated maize varieties70,74, which were later fully domesticated at secondary centres in a stratified domestication process290. Paleogenomic data also do not support the presence of a strong demographic bottleneck during early domestication stages [see291 for a review in major crops] but, rather, studies of wheat292, sunflowers293 and horses148 support more recent losses of genetic diversity.

Finally, time-stamped genomes have enabled the temporal stratification of both phenotypes and selective regimes. In sorghum, early selection was related to plant architecture and only later switched to increased sugar metabolism294. In maize, complex phenotypes such as days to flowering have been reconstructed from genomes showing latitudinal adaptation to day length295. Early selection of coat colour is evident in pigs282 and goats289, but some traits were acquired later, such as behaviour-related and productivity traits in chickens during the medieval period296,297 and morphotype and speed in horses148,298. Importantly, damaging mutation load was a feature of later selective episodes and breeding but not part of the initial domestication of horses298, maize295 and sorghum294. Continuing to characterize past genomic variation in ancient crops and animal breeds will be instrumental for developing the future sustainability of agriculture.

Pathogens and microbiomes

Together with classical approaches in paleopathology and paleodemography, aDNA from microorganisms, including pathogens and commensals, can provide insights into the health of ancient peoples as well as shifts in diets and disease ecology. Initial studies of ancient microorganisms used PCR-based methods to identify specific pathogens [such as those causing skeletal lesions characteristic of tuberculosis and leprosy]299, to analyse the first historic pathogen genome from the 1918 influenza virus300 and to explore the ancient microbiome27. However, such approaches suffered from an inability to distinguish ancient and modern contaminating microbial DNA27,301. More recently, NGS-based methods have resulted in the successful recovery of authenticated pathogen genomes from not only Mycobacterium tuberculosis and Mycobacterium leprae but also many pathogens that leave no visible evidence in the skeleton, such as Yersinia pestis, Helicobacter pylori, Vibrio cholera, Salmonella enterica serovar Paratyphi C, variola virus, human parvovirus and hepatitis B virus [reviewed elsewhere13]. For example, analyses of the victims of both the Plague of Justinian and the Black Death have shown that Y. pestis was indeed the cause of both pandemics65,302, and further genomic analysis has provided insight into the course of these pandemics over time, including parallel changes in the pathogen itself303,304,305. Surprisingly, extinct strains of Y. pestis have also been identified in ancient individuals dating from as early as the Late Neolithic and Bronze Ages throughout Eurasia, and genomic data show the progression of evolutionary changes that have affected its virulence through time, including the acquisition of a transmission mode via fleas [reviewed elsewhere13].

Although the analyses of ancient pathogen genomes have shown plague to be older than expected, they suggest that the origins of the tuberculosis complex [M. tuberculosis and related strains] in humans may have evolved more recently [~3,000–6,000 years ago]306,307. PCR data exist that putatively identify tuberculosis in older individuals, but genomic data from bone samples that would distinguish contaminating microbial DNA are not yet available27,301. Additional genomic data from ancient tuberculosis strains could show that the current estimate of time to the most recent common ancestor reflects a bottleneck or other population dynamics in this clonal pathogen. The ancient tuberculosis data also show the significance of pathogen exchange among animals and humans. In South America, before European colonization, aDNA has shown that people were affected by tuberculosis strains that are most closely related to Mycobacterium pinnipedi, a zoonotic form of the pathogen that usually infects seals and sea lions. This strain was likely transmitted from South American seals to humans through consumption of undercooked seal meat or during butchering, after which human to human transmission may have begun306.

Similarly, NGS has facilitated microbiome analyses from a range of ancient contexts including dental calculus, coprolites, latrine sediments and mummies308,309. Comparisons between modern and ancient oral microbiota have suggested possible shifts in relation to increasingly carbohydrate-rich diets during the Neolithic and Industrial revolution44, although further work revealed the need to account for other possible factors driving oral microbiota profiles, such as the biofilm maturation stage310, tooth type and surface311. Likewise, studies of the ancient gut microbiome have revealed previously unknown microbial diversity in the human gut, as well as the loss of key microbial symbionts in present-day industrialized populations85,235. The microbiome is also the source of several recently evolved human pathogens, including the causative agents of diphtheria [Corynebacterium diphtheriae], gonorrhoea [Neisseria gonorrhoeae], bacterial meningitis [Neisseria meningitidis] and pneumonia [Streptococcus pneumoniae and Haemophilus influenzae] [reviewed elsewhere309], and thus understanding its evolution and changing ecology through time is critical to understanding infectious human disease.

Extinction and climate change

aDNA has fed debates not only on the phylogenetic placement of many extinct species but also on the causes of their disappearance. This can be investigated by correlating known human activities or climatic events with population expansions and declines as revealed from serially sampled DNA data. Dynamic population size trajectories can be estimated from calibrated gene genealogies within a serial coalescent statistical framework [reviewed elsewhere312]. Additionally, the regions showing paleoclimatic conditions compatible with the presence of a given species can be inferred from spatio-temporal fossil distribution data and bioclimatic data. Key to all such analyses is the availability of radiocarbon dates for the specimens analysed using genetic techniques. Radiocarbon dating should thus be highly recommended wherever possible.

Evidence currently available in bison, horses, reindeer, musk oxen, woolly mammoths and woolly rhinos suggests that climate change is a common driver of population size changes but also that species show individualistic responses in the face of climatic and anthropogenic pressure20. Human activities have been found to possibly mediate species demise through various mechanisms, including direct over-exploitation, as shown in moas313 and great auks314. Extinctions can also have many indirect causes that can be difficult to disentangle, such as human-mediated habitat disturbance or introduction of pathogens or predators. For example, the occupation of caves by Aurignacian humans competed with the natural homing behaviour of cave bears for places in which they could hibernate to survive the winter23,24. As regards the extinct huia birds of New Zealand, the drastic reduction of forest cover and the translocation of mammal predators during the European settlement may have been fatal to these endemic passerines315.

Complete genome data from single diploid individuals have provided new avenues for reconstructing paleodemographic trajectories using the pairwise316 and multiple317 sequentially Markovian coalescent modelling frameworks. This approach has revealed that archaic hominin populations survived extinction over long periods of time despite highly reduced effective population sizes8,9,10. Analysis of the complete genome from a late-surviving woolly mammoth has identified two episodes of severe demographic collapse in the Early-Middle Pleistocene and at the Pleistocene–Holocene transition3. The resulting reduced genetic diversity and the accumulation of numerous potentially deleterious mutations may have precipitated the population’s extinction through the alteration of important functions in development, reproduction and olfaction17,18,19. Demographic reconstructions can, however, be biased by strong natural selection acting on genomes, especially in extreme cases of large population sizes. This was observed for the passenger pigeon that went from a census population size of several billion to extinction in only decades318. By reconstructing temporal baselines prior to environmental and/or human impact, it is increasingly clear that aDNA from museum and archaeological remains can advantageously complement current genetic diversity assessments to establish extinction risks and conservation priorities319, as previously hinted for Przewalski’s horses320 and arctic foxes321.

Environmental DNA

The retrieval of animal and plant DNA from sediments, ice and lake cores, commonly known as environmental DNA [eDNA], is probably the only research area that traces its origins to the field of aDNA itself rather than being adopted from contemporary genetics. Early pioneering work successfully retrieved DNA fragments from extinct mammals and birds from only a few grams of Siberian permafrost and New Zealand cave sediments90. If requiring large drilling equipment, sampling procedures should be experimentally tested for contamination, for example by introducing synthetic plasmids of known sequences at the surface of the equipment whose penetration into the drilling core can be tracked22,48. The experimental workflow typically includes DNA extraction and the sequencing of biomarkers providing taxonomic resolution322, and special attention must be given to rule out on-site contamination48 and stratigraphic leaching323. Early analyses relied on molecular cloning of PCR-amplified meta-barcodes, but massively parallel sequencing facilitated access to the entire molecular diversity of individual animals and plants324. Current approaches build on shotgun metagenomics48 and more economical techniques involving target enrichment49,325.

One of the great benefits of eDNA is that it allows the detection of all domains of life, from microorganisms to vertebrate species alike, even in the absence of macrofossils325,326. Taxonomic resolution is, however, limited owing to the extensive fragmentation of eDNA, and varies with the completeness of comparative sequence databases. The extent to which quantitative assessments can be obtained is currently unknown, although new biomass proxies aim at this objective136. The exact sources of eDNA fragments and the conditions governing their preservation remain largely unknown. eDNA is assumed to derive from skin cells, faeces, urine and microfossils, and various mineral particles have been suggested to favour its preservation within sediments90. Microscopic bone and tooth fragments that are too small to be identified, and are thus confused with sediments, have also been speculated to potentially represent a significant source of eDNA49.

eDNA has proved a powerful approach to assess the first and last appearance of taxa in the fossil record, and has been applied to investigate the timing of the extinction of woolly mammoths in mainland Alaska326 and Yukon325 and the survival of spruce in Scandinavian refugia during the last Ice Age130. eDNA has also provided insights into the spatial and temporal distribution of animal species, including extinct hominins49, and the impact of global climatic change22,48,90,91,327. All of these studies have in one way or another challenged the validity of climate niche predictions and commonly accepted models of extinction, biogeography and archaeology.

Reproducibility and data deposition

Possible confounding factors

Given that numerous tools and approaches are available for analysing ancient genomic data, reproducibility of results can be hampered by the lack of sufficiently detailed descriptions of each analytical step in the methods sections of publications. Although some guidelines exist for the analysis of ancient genomic data — for example, for mapping162,181,183 or data authentication56,57,328 — and automated tools embedding critical analytical procedures within open-source pipelines are available187,188,241,242, there is still no consensus on a ‘gold standard’ method for carrying out [or reporting] most analyses. More often, the selection of programs and parameters is research group-based, with publications reporting different analytical approaches and criteria. The list of factors to consider extends far beyond the minimal data requirement for specific analyses66 and includes platform-specific sequence error profiles169; the magnitude of post-mortem DNA damage and contamination, and their impact on downstream inference55; base quality rescaling57; read alignment parameters180,182; database selection and taxonomic classifier parameters27,238; awareness of the applicability conditions of statistical methods; and many more. In the face of the complexity and diversity of possible confounding factors, it is of the utmost importance that all relevant analytical parameters, including software versions and parameters [even if default], are described in full detail in the methods section of any scientific article reporting aDNA data.

Public repositories

To be accepted as valid forms of evidence, raw sequence data and alignments against reference genomes must be made freely available through public repositories such as the Sequence Read Archive [SRA] and the European Nucleotide Archive [ENA] in order to ensure research reproducibility and future analyses. Entirely curated data sets are also made available upon publication through individual laboratory websites [for example, the Reich laboratory website]. Full traceability of the underlying experimental procedures, including sequencing chemistry and base calling software, is necessary to account for possible data structure deriving from technical artefacts. We recommend that raw, unfiltered sequencing data be uploaded as compressed fastq files together with the alignment files underlying the analyses presented — in BAM format for individual [mito]genomes and in multifasta format for barcode alignment or phylogenetic reconstruction. Labels — for example, read groups and sample names — should match those provided in the supplemental sections of the original publication.

Long-term legacy

The destructive sampling underpinning ancient genomic work has raised severe concerns pertaining to the long-term viability of the data produced. Current approaches are mainly based on whole-genome shotgun sequencing or targeted SNP capture. Although the latter is both time and cost-effective160, and remains the only methodology showing sufficient sensitivity to gather data from extremely degraded specimens119, it considerably restricts the amount of information retrieved from samples showing better molecular preservation. In ancient human genome studies, this approach typically queries 1.2 million positions known to be polymorphic amongst a worldwide panel of contemporary human populations160, leaving other positions and other DNA material present in the extracts, such as those from pathogens or microbiota, unexplored. Whole-genome shotgun sequencing is instead not limited to any pre-selected target regions or genomic variants but, rather, aims to uniformly cover the entire metagenome. Although more expensive to generate and more demanding to analyse, this untargeted approach can reveal new informative variants and produce data that are applicable to future research questions. Regardless of the approach, however, it is also possible to create from the original extracted aDNA an immortal DNA sequencing library that can serve as a long-term genetic archive of the sample, and which could be stored in a museum or laboratory cryo-facility. If prepared correctly, this library can be theoretically reamplified indefinitely without exhaustion or loss of complexity143. We advise curators to balance the uniqueness of the material considered, known or estimated DNA preservation rates, experimental costs and long-term plans for both DNA library and data archiving when authorizing destructive sampling.

Limitations and optimizations

Wet-laboratory methods

Owing to the sensitive nature of aDNA, preparation and manipulation of aDNA must be carried out in laboratory facilities with positive air pressure, UV surface irradiation and strict cleaning procedures to ensure minimal contamination. Clean laboratory facilities must be physically separated from other laboratory areas where DNA is amplified, captured and/or sequenced. Samples must be selected according to the main objectives of the study. For example, whereas petrosal bones are generally preferred for retrieving genome-scale data from host species, their metagenomic potential for pathogen studies is limited66. Rates of aDNA preservation also vary at microscopic scales. Therefore, multiple analyses repeated on small amounts of material can increase the chances of success compared with a single analysis from a large sample, which averages out preservation rates and maximizes destruction. No DNA analyses should be carried out on precious, rare material in the absence of preliminary analyses, suggesting feasibility and supporting molecular preservation on site. Specific chemical treatment of bone powder, for example with bleach40,42, can reduce the exogenous DNA fraction and may be attempted in cases of repeated failure due to contamination; however, such treatment can also artificially age modern contaminant DNA, and so authentication of bleach-treated samples should be performed with care. DNA libraries should contain multiple indices to eliminate the inclusion of spurious recombining DNA templates in downstream analyses142. PCR amplification of DNA libraries must be carried out in conditions maximizing molecular complexity, and no DNA extracts and/or DNA libraries should be exhausted for a single analysis. In the case of target enrichment, two successive rounds of capture generally augment on-target recovery rates166. Library pooling prior to sequencing helps to reduce sequencing costs but requires that the base composition is balanced, especially for Illumina platforms, as a balanced base composition within the first six nucleotides is instrumental for calibrating the fluorescence measurement of nucleotide bases. This can be achieved by spiking library pools with sufficient amounts of a PhiX DNA library prior to sequencing. Large-scale projects can benefit from nascent automation procedures both for the preparation of DNA libraries and for their capture49,126,144. In all circumstances, the experimental procedures implemented from sampling to sequencing must be recorded and fully described [and, ideally, shared through online repositories such as protocols.io] so as to assess the possible impact of experimental differences in the sequence data.

Dry-laboratory methods

With genome-scale sequence data, the number of possible statistical analyses is virtually limitless. Common caveats include the impact of contamination and post-mortem DNA damage and of read alignment parameters. For highly degraded or contaminated material, analyses may be conditioned on reads showing evidence of post-mortem DNA damage55,189,192 and/or limited to transversions7 so as to mitigate both rampant contamination and damage-related nucleotide misincorporation. This approach, however, significantly reduces the amount of data available. Quantitative assessment of damage and contamination followed by explicit simulation180 can help establish the robustness of the conclusions. Microbial taxonomic assignment should be cross-validated through multiple approaches and the presence of potential pathogens must be established by checking that the read edit distance distribution against the reference genomes of close relatives indicates closer proximity with the candidate27. The amount of sequence data necessary for a given analysis depends on the underlying population history and genetic diversity present in a species. For example, a minimal threshold of 10,000–30,000 SNPs is commonly used for assessing genetic relationships of human individuals at intra-continental scales329,330 but molecular sexing of mammal species and identification of first-generation hybrids generally requires no more than 1,000–10,000 mapped reads200. The computational procedures must be fully described with explicit reference to analytical parameters so as to ensure reproducibility [and, ideally, shared through online repositories such as GitHub or Bitbucket].

Outlook

Working together

Within the past decade, aDNA research has come of age and has complemented the toolkits of archaeologists and evolutionary biologists, with techniques providing an unprecedented resolution of detail on past environments, societies and individuals. Perhaps the most important contribution of aDNA is that it offers access to a wealth of biological information that can help shape and test working hypotheses about the past in a quantitative manner. Adopting such a hypothesis-testing framework comes naturally for evolutionary biologists who share similar methodologies, including phylogenetic reconstruction and population genetics. This approach, however, differs from that of archaeologists, whose research is primarily grounded within the epistemological frameworks of the social sciences and humanities. As a fast-evolving and inherently interdisciplinary field, aDNA research is where these distinct disciplinary approaches mix and meet — to both great effect and occasional mutual misunderstanding. For example, most aDNA work to date has focused on big picture studies comprising large temporal and spatial scales. Although necessary for delineating major trends and new working hypotheses, such studies have sometimes divided the archaeological community, as they necessarily oversimplified the inherent complexity of past societies and the archaeological record itself. Moreover, the often unreflected use of problematic terms and the simplistic equivalence of genetic and cultural groupings [Box 1] have found major criticism104,107. Whether seemingly rapid shifts in ancestry detected in aDNA data were the result of sudden massive population replacements and/or long-term individual human mobility owing to institutions such as patrilocality has also been a matter of discussion and debate331,332. However, now that studies have created a basic overview of the population genomic developments in many regions of the world, it is possible to zoom into more local contexts. Future projects can, for instance, study to what extent biological relatedness was decisive for the constitution of social belonging expressed, for example, through the joint burial of individuals in collective graves over time. Paleodemographic profiles augmented with aDNA evidence on the virulence of past pathogens, as well as paleopathological data, markers of violence and activity and dietary isotopic signatures, can help to establish the underlying causes of health changes in past societies. The reconstruction of ancient trade routes and networks represents another area that could greatly benefit from joint research efforts combining aDNA, archaeological fieldwork and remote sensing, for example by integrating genetic markers of migration and metallurgic signatures of material provenance into geographical information system data analysis. Encouragingly, in the past few years, scholars with different backgrounds in the fields of [bio]archaeology have increasingly learned to work together and have developed many initiatives to establish and strengthen the crucial interdisciplinary dialogue and mutual understanding that forms the basis for a future reflective bioarchaeology104. It has become clear that the full community of stakeholders [including descendant communities, archaeologists, biological anthropologists and museum curators] needs to be embraced, from sampling design to data interpretation, to avoid misinterpreting or over-interpreting the evidence and to collaboratively build a strong foundation for future work [Box 2].

Box 2 Who conducts aDNA research?

The field of ancient DNA [aDNA] research has diverse and heterogeneous disciplinary origins. Initial studies were largely led by biochemists and primarily focused on evolutionary questions relating to extinct mammals, but the field quickly diversified to also encompass anthropological and forensics questions. Today, evolutionary biology and anthropology have become the two main disciplinary homes of aDNA research, and there has been tremendous growth in the number of aDNA laboratories over the past decade. Housed in museums, research institutes and university departments, the size of these laboratories ranges from small laboratories serving a single principal investigator to large laboratories with multiple principal investigators supporting dozens of technicians, students and postdoctoral researchers. Funding is also highly varied, with some laboratories receiving core government funding whereas others depend entirely on third-party research grants. Moreover, funding agencies, grant types and funding levels also differ substantially across countries and disciplines, and such differences contribute to large disparities across different aDNA laboratories in terms of their resources and research capabilities. International collaboration and community-building, the establishment of enduring teaching and training networks and the formation of academic societies and research consortia are needed to leverage resources across the field, remove barriers to access for smaller laboratories and support the growth of the field as a whole. Initiatives such as the biannual International Symposium for Biomolecular Archaeology [ISBA] conference and the Standards and Precautions and Advances in Ancient Metagenomics [SPAAM] workshop aim to address some of these challenges, but additional collaborative efforts, such as lobbying of funding agencies and collective bargaining to make large in-solution probe sets available through commercial vendors, are needed.

Future directions

In studies of humans, aDNA research has thus far been most successful when reconstructing ancestry profiles of past individuals and their population affinities. Future work may increasingly leverage patterns of linkage disequilibrium across sites to gain finer-scale genetic resolution. This will prove especially useful for identifying subtle genetic affinities not captured by unlinked SNP information alone229 and when DNA preservation limits the amount of genetic information recovered. However, gaining further insights into the past will require molecular investigations beyond humans, their diseases and their domesticates, to include the full range of species and biomaterials preserved in the archaeological and paleontological record, from textiles and ceramics to sediments. Integrating DNA data with other proxies, including stable and radiogenic isotopes112, microremains and small-molecule metabolites333 and paleoproteomic data113, will greatly improve our knowledge about the lives of past peoples including their diets26,334, environments, mobility335, craft activities336, drug and medicinal use337 and their health status and stress26,338. The successful application of such multi-proxy approaches to the study of the 5,300-year-old stomach contents of the Iceman has, for example, provided unprecedented resolution into the nutritional habits and food-processing methods during the European Copper Age339.

An increasing body of work has shown that genetic data in tandem with archaeological and paleopathological data can illuminate social rules governing past societies, including marital patrilocality and social inequality205, consanguinity222, inbreeding avoidance255 and care for individuals with genetic disorders. Recent case studies [for example, for the Early Bronze Age in Southern Germany205 and the Neolithic in Ireland222] have shed new light on past social institutions and societal transformations by deciphering the complexity of households and individual mobility as well as the dependency of wealth and status on biological relatedness. Here, archaeogenetic studies have the potential to dramatically advance our understanding of past social structures and the interplay of biological relatedness, health and mobility with social practices and worldviews. Armed with ancient epigenetic markers, future aDNA work will likely inform us about age at death of past individuals29,248 and the biological and medical consequences of their social status.

Given that all of these approaches are destructive and archaeological remains are finite, however, it is critical that resources are used with the aim of maximizing the amount of data collected and that efforts to develop multi-proxy methods compatible with minute sampling are explored112,113. Efforts to better understand post-mortem DNA degradation and to repair aDNA lesions109 should be pursued to help further minimize destructive sampling. Such efforts would facilitate DNA analyses into the Upper Paleolithic and Middle Pleistocene, and possibly beyond the current record for the oldest genome around 560,000–780,000 years ago2. Increasing focus on environmental aDNA fragments preserved in sediments rather than in macrofossils will also provide additional opportunities for minimizing destruction while, perhaps, enriching and providing a fuller account of the taxonomic diversity of plant and animal communities in past ecosystems22,34,48,49. In addition, applying these approaches to marine and deep-sea sediments may increasingly inform us about the resilience of the ocean system to past global environmental change340, and reveal ecosystems and past events in landscapes that became submerged during the Holocene134,136. Such studies promise to have a transformative impact on our understanding of the peopling of the Americas and island southeast Asia, as well as early human expansions out of Africa.

Over the past decade, short-read high-throughput DNA sequencing technologies have fundamentally changed the field of aDNA, but, even now, new sequencing technologies are on the horizon341. At first glance, long-read sequencers such as PacBio and Oxford Nanopore seem to have little relevance for highly degraded aDNA; however, concatenating aDNA fragments separated by spacers within single library templates may offer new opportunities for aDNA sequencing, including the direct detection of base modifications. In the longer term, this may even represent the only economical solution for future paleogenomic studies as short-read sequencing technologies are slowly phased out. New advances in de novo genomic and metagenomic assembly likewise hold promise for moving beyond reference-based mapping and towards a more accurate and complete reconstruction of ancient microbial genomes and metagenomes238, including improved strain separation, genomic architecture reconstruction and the identification and recovery of novel genes. The past decade has shown that aDNA researchers can be extremely creative in developing innovative solutions to harness the full power of available high-throughput sequencing technologies. In the future, it will be essential to continue to adapt our research toolkits to incorporate new and emerging technologies as we strive to fully access the complete amount of information preserved in the fossil and sedimentary record.

New developments in inferring function from ancient genetic sequences are also afoot. Genome editing technologies such as CRISPR–Cas9 [ref.342], induced pluripotent stem cells343 and miniaturized organoid systems mimicking simplified organs344 now offer the possibility to investigate the functional consequences of virtually any change in the sequence of a given genome. The comparison of our genome and that of archaic hominins has, for example, provided a preliminary list of the genetic changes that made us humans, and attempts to understand their consequences on human brain development have already started345. Future research will likely increasingly rely on these technologies and genome time-series data of model organisms to assess the adaptive consequences of past genetic changes.

Over almost 40 years, aDNA research has taught us surprising things about the human story and life on this planet. It has moved forwards in parallel with technological and computational advances in the life sciences and has grown and matured in response to new developments in the social sciences and humanities. The discovery that a vast and invisible molecular past survives in the archaeological record has had a transformative effect on the field of archaeology and promises many new and unexpected findings to come. In the future, there is no doubt that our ability to increasingly detect and reconstruct the molecular archaeological record will enrich the human story and contribute to the transdisciplinary research endeavour of understanding our shared human past.

References

  1. Higuchi, R., Bowman, B., Freiberger, M., Ryder, O. A. & Wilson, A. C. DNA sequences from the quagga, an extinct member of the horse family. Nature 312, 282–284 [1984].

    ADS  Google Scholar 

  2. Orlando, L. et al. Recalibrating Equus evolution using the genome sequence of an early Middle Pleistocene horse. Nature 499, 74–78 [2013]. This article presents the oldest sequenced genome to date, from one horse metapodial preserved in permafrost for 560,000–780,000 years.

    ADS  Google Scholar 

  3. Palkopoulou, E. et al. Complete genomes reveal signatures of demographic and genetic declines in the woolly mammoth. Curr. Biol. 25, 1395–1400 [2015].

    Google Scholar 

  4. Barlow, A. et al. Partial genomic survival of cave bears in living brown bears. Nat. Ecol. Evol. 2, 1563–1570 [2018].

    Google Scholar 

  5. Margaryan, A. et al. Population genomics of the Viking world. Nature 585, 390–396 [2020].

    ADS  Google Scholar 

  6. Rasmussen, M. et al. Ancient human genome sequence of an extinct Palaeo-Eskimo. Nature 463, 757–762 [2010]. This article reports the first ancient human genome sequenced from the hair shaft of a 4,000-year-old Paleo-Inuit. The authors find evidence for genetic discontinuity with modern Greenlanders, supporting multiple migration waves into Arctic Greenland.

    ADS  Google Scholar 

  7. Green, R. E. et al. A draft sequence of the Neandertal genome. Science 328, 710–722 [2010]. This article reports the first Neanderthal genome obtained from the DNA extracts of three paleontological bones, establishing the Neanderthal genetic legacy within modern human genomes and describing important wet-laboratory and dry-laboratory methodologies that have shaped the following decade of aDNA research.

    ADS  Google Scholar 

  8. Prüfer, K. et al. The complete genome sequence of a Neanderthal from the Altai Mountains. Nature 505, 43–49 [2014].

    ADS  Google Scholar 

  9. Prüfer, K. et al. A high-coverage Neandertal genome from Vindija cave in Croatia. Science 358, 655–658 [2017].

    ADS  Google Scholar 

  10. Mafessoni, F. et al. A high-coverage Neandertal genome from Chagyrskaya cave. Proc. Natl Acad. Sci. USA 117, 15132–15136 [2020].

    Google Scholar 

  11. Nielsen, R. et al. Tracing the peopling of the world through genomics. Nature 541, 302–310 [2017].

    ADS  Google Scholar 

  12. Frantz, L. A. F., Bradley, D. G., Larson, G. & Orlando, L. Animal domestication in the era of ancient genomics. Nat. Rev. Genet. 21, 449–460 [2020].

    Google Scholar 

  13. Spyrou, M. A., Bos, K. I., Herbig, A. & Krause, J. Ancient pathogen genomics as an emerging tool for infectious disease research. Nat. Rev. Genet. 20, 323–340 [2019].

    Google Scholar 

  14. Krause, J. et al. The complete mitochondrial DNA genome of an unknown hominin from southern Siberia. Nature 464, 894–897 [2010].

    ADS  Google Scholar 

  15. Reich, D. et al. Genetic history of an archaic hominin group from Denisova cave in Siberia. Nature 468, 1053–1060 [2010]. This article reports the discovery of Denisovans, a previously unknown lineage of archaic hominins that lived in southern Siberia at least 50,000 years ago. This was the first time that a representative of the Homo evolutionary tree was described from molecular data and in the absence of key macrofossil remains showing clear morphological characteristics.

    ADS  Google Scholar 

  16. Meyer, M. et al. A high-coverage genome sequence from an archaic Denisovan individual. Science 338, 222–226 [2012]. This article presents the first high-quality genome from an archaic hominin and describes the first experimental procedure for DNA library preparation from single-stranded DNA templates. The approach outperforms competing technologies in sensitivity and complexity, and minimizes loss of authentic DNA molecules.

    ADS  Google Scholar 

  17. Rogers, R. L. & Slatkin, M. Excess of genomic defects in a woolly mammoth on Wrangel island. PLoS Genet. 13, e1006601 [2017].

    Google Scholar 

  18. Lynch, V. J. et al. Elephantid genomes reveal the molecular bases of woolly mammoth adaptations to the Arctic. Cell Rep. 12, 217–228 [2015].

    Google Scholar 

  19. Fry, E. et al. Functional architecture of deleterious genetic variants in the genome of a Wrangel island mammoth. Genome Biol. Evol. 12, 48–58 [2020].

    Google Scholar 

  20. Lorenzen, E. D. et al. Species-specific responses of late quaternary megafauna to climate and humans. Nature 479, 359–364 [2011].

    ADS  Google Scholar 

  21. Orlando, L. & Cooper, A. Using ancient DNA to understand evolutionary and ecological processes. Annu. Rev. Ecol. Evol. Syst. 45, 573–598 [2014].

    Google Scholar 

  22. Willerslev, E. et al. Fifty thousand years of Arctic vegetation and megafaunal diet. Nature 506, 47–51 [2014].

    ADS  Google Scholar 

  23. Fortes, G. G. et al. Ancient DNA reveals differences in behaviour and sociality between brown bears and extinct cave bears. Mol. Ecol. 25, 4907–4918 [2016].

    ADS  Google Scholar 

  24. Gretzinger, J. et al. Large-scale mitogenomic analysis of the phylogeography of the Late Pleistocene cave bear. Sci. Rep. 9, 10700 [2019].

    ADS  Google Scholar 

  25. Key, F. M. et al. Emergence of human-adapted Salmonella enterica is linked to the Neolithization process. Nat. Ecol. Evol. 4, 324–333 [2020].

    Google Scholar 

  26. Warinner, C. et al. Pathogens and host immunity in the ancient human oral cavity. Nat. Genet. 46, 336–344 [2014]. This article reports the first metagenomic and paleoproteomic analysis of ancient dental plaque and demonstrates the preservation of oral microbial signatures, diet content and inflammation markers.

    Google Scholar 

  27. Warinner, C. et al. A robust framework for microbial archaeology. Annu. Rev. Genomics Hum. Genet. 18, 321–356 [2017].

    Google Scholar 

  28. Gokhman, D. et al. Reconstructing the DNA methylation maps of the Neandertal and the Denisovan. Science 344, 523–527 [2014].

    ADS  Google Scholar 

  29. Pedersen, J. S. et al. Genome-wide nucleosome map and cytosine methylation levels of an ancient human genome. Genome Res. 24, 454–466 [2014]. This article is the first report of ancient epigenomes, leveraging post-mortem DNA degradation signatures to statistically infer DNA methylation and nucleosome positioning.

    Google Scholar 

  30. Goodwin, S., McPherson, J. D. & McCombie, W. R. Coming of age: ten years of next-generation sequencing technologies. Nat. Rev. Genet. 17, 333–351 [2016].

    Google Scholar 

  31. Poinar, H. N. et al. Metagenomics to paleogenomics: large-scale sequencing of mammoth DNA. Science 311, 392–394 [2006].

    ADS  Google Scholar 

  32. Green, R. E. et al. Analysis of one million base pairs of Neanderthal DNA. Nature 444, 330–336 [2006].

    ADS  Google Scholar 

  33. Orlando, L., Gilbert, M. T. P. & Willerslev, E. Reconstructing ancient genomes and epigenomes. Nat. Rev. Genet. 16, 395–408 [2015].

    Google Scholar 

  34. Giguet-Covex, C. et al. Long livestock farming history and human landscape shaping revealed by lake sediment DNA. Nat. Commun. 5, 3211 [2014].

    ADS  Google Scholar 

  35. Avila-Arcos, M. C. et al. Application and comparison of large-scale solution-based DNA capture-enrichment methods on ancient DNA. Sci. Rep. 1, 74 [2011].

    Google Scholar 

  36. Maricic, T., Whitten, M. & Pääbo, S. Multiplexed DNA sequence capture of mitochondrial genomes using PCR products. PLoS ONE 5, e14004 [2010].

    ADS  Google Scholar 

  37. Fu, Q. et al. DNA analysis of an early modern human from Tianyuan cave, China. Proc. Natl Acad. Sci. USA 110, 2223–2227 [2013]. This study presents the first application of in-solution target enrichment at the genome scale and describes sequence for a full chromosome of an approximately 40,000-year-old anatomically modern human from China.

    ADS  Google Scholar 

  38. Carpenter, M. L. et al. Pulling out the 1%: whole-genome capture for the targeted enrichment of ancient DNA sequencing libraries. Am. J. Hum. Genet. 93, 852–864 [2013].

    Google Scholar 

  39. Dabney, J. et al. Complete mitochondrial genome sequence of a Middle Pleistocene cave bear reconstructed from ultrashort DNA fragments. Proc. Natl Acad. Sci. USA 110, 15758–15763 [2013]. This study reports a new DNA extraction method from ancient osseous remains that is tailored to the ultrashort and extensively damaged nature of aDNA molecules. This methodology allowed the retrieval of full mitochondrial genome sequences from 300,000-year-old cave bear specimens preserved in Atapuerca, Spain.

    ADS  Google Scholar 

  40. Korlević, P. et al. Reducing microbial and human contamination in DNA extractions from ancient bones and teeth. BioTechniques 59, 87–93 [2015].

    Google Scholar 

  41. Gamba, C. et al. Comparing the performance of three ancient DNA extraction methods for high-throughput sequencing. Mol. Ecol. Resour. 16, 459–469 [2016].

    Google Scholar 

  42. Boessenkool, S. et al. Combining bleach and mild predigestion improves ancient DNA recovery from bones. Mol. Ecol. Resour. 17, 742–751 [2017].

    Google Scholar 

  43. Gansauge, M.-T. & Meyer, M. Selective enrichment of damaged DNA molecules for ancient genome sequencing. Genome Res. 24, 1543–1549 [2014].

    Google Scholar 

  44. Adler, C. J. et al. Sequencing ancient calcified dental plaque shows changes in oral microbiota with dietary shifts of the Neolithic and industrial revolutions. Nat. Genet. 45, 450–455, 455e1 [2013].

    Google Scholar 

  45. Wagner, S. et al. High-throughput DNA sequencing of ancient wood. Mol. Ecol. 27, 1138–1154 [2018].

    Google Scholar 

  46. Lendvay, B. et al. Improved recovery of ancient DNA from subfossil wood—application to the world’s oldest Late Glacial pine forest. New Phytol. 217, 1737–1748 [2018].

    Google Scholar 

  47. Der Sarkissian, C. et al. Ancient DNA analysis identifies marine mollusc shells as new metagenomic archives of the past. Mol. Ecol. Resour. 17, 835–853 [2017].

    Google Scholar 

  48. Pedersen, M. W. et al. Postglacial viability and colonization in North America’s ice-free corridor. Nature 537, 45–49 [2016].

    ADS  Google Scholar 

  49. Slon, V. et al. Neandertal and Denisovan DNA from Pleistocene sediments. Science 356, 605–608 [2017].

    ADS  Google Scholar 

  50. Teasdale, M. D. et al. Paging through history: parchment as a reservoir of ancient DNA for next generation sequencing. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 20130379 [2015].

    Google Scholar 

  51. O’Sullivan, N. J. et al. A whole mitochondria analysis of the Tyrolean Iceman’s leather provides insights into the animal sources of Copper Age clothing. Sci. Rep. 6, 31279 [2016].

    ADS  Google Scholar 

  52. Gamba, C. et al. Genome flux and stasis in a five millennium transect of European prehistory. Nat. Commun. 5, 5257 [2014]. This study reports the first matched empirical evidence establishing better DNA preservation rates in petrosal bones, which paved the way for future studies at the population scale, and also presents complete genomes of Mesolithic hunter–gatherers and Neolithic farmers in the Caucasus, supporting direct early contact between both groups.

    ADS  Google Scholar 

  53. Damgaard, P. B. et al. Improving access to endogenous DNA in ancient bones and teeth. Sci. Rep. 5, 11184 [2015].

    ADS  Google Scholar 

  54. Krause, J. et al. A complete mtDNA genome of an early modern human from Kostenki, Russia. Curr. Biol. 20, 231–236 [2010]. This article establishes important criteria for aDNA data authentication, including the presence of post-mortem cytosine deamination signatures and DNA fragmentation through depurination, and reports a complete mitochondrial genome from an approximately 32,000-year-old anatomically modern human from Kostenki, Russia.

    Google Scholar 

  55. Skoglund, P. et al. Separating endogenous ancient DNA from modern day contamination in a Siberian Neandertal. Proc. Natl Acad. Sci. USA 111, 2229–2234 [2014].

    ADS  Google Scholar 

  56. Renaud, G., Schubert, M., Sawyer, S. & Orlando, L. Authentication and assessment of contamination in ancient DNA. Methods Mol. Biol. 1963, 163–194 [2019].

    Google Scholar 

  57. Jónsson, H., Ginolhac, A., Schubert, M., Johnson, P. L. F. & Orlando, L. mapDamage2.0: fast approximate Bayesian estimates of ancient DNA damage parameters. Bioinformatics 29, 1682–1684 [2013].

    Google Scholar 

  58. Leonardi, M. et al. Evolutionary patterns and processes: lessons from ancient DNA. Syst. Biol. 66, e1–e29 [2017].

    Google Scholar 

  59. Pääbo, S., Gifford, J. A. & Wilson, A. C. Mitochondrial DNA sequences from a 7000-year old brain. Nucleic Acids Res. 16, 9775–9787 [1988].

    Google Scholar 

  60. Miller, W. et al. Sequencing the nuclear genome of the extinct woolly mammoth. Nature 456, 387–390 [2008].

    ADS  Google Scholar 

  61. Gilbert, M. T. P. et al. Resistance of degraded hair shafts to contaminant DNA. Forensic Sci. Int. 156, 208–212 [2006].

    Google Scholar 

  62. Seguin-Orlando, A. et al. Pros and cons of methylation-based enrichment methods for ancient DNA. Sci. Rep. 5, 11826 [2015].

    ADS  Google Scholar 

  63. Hagelberg, E., Sykes, B. & Hedges, R. Ancient bone DNA amplified. Nature 342, 485 [1989].

    ADS  Google Scholar 

  64. Sirak, K. et al. Human auditory ossicles as an alternative optimal source of ancient DNA. Genome Res. 30, 427–436 [2020].

    Google Scholar 

  65. Bos, K. I. et al. A draft genome of Yersinia pestis from victims of the Black Death. Nature 478, 506–510 [2011]. This article reports the first complete genome of an ancient bacterial pathogen, Y. pestis, from osseous human remains of individuals who died from the Black Death in 1347–1348.

    ADS  Google Scholar 

  66. Margaryan, A. et al. Ancient pathogen DNA in human teeth and petrous bones. Ecol. Evol. 8, 3534–3542 [2018].

    Google Scholar 

  67. Brown, T. A., Allaby, R. G., Brown, K. A., O’Donoghue, K. & Sallares, R. DNA in wheat seeds from European archaeological sites. Experientia 50, 571–575 [1994].

    Google Scholar 

  68. Suyama, Y. et al. DNA sequence from a fossil pollen of Abies spp. from Pleistocene peat. Genes. Genet. Syst. 71, 145–149 [1996].

    Google Scholar 

  69. Jaenicke-Després, V. et al. Early allelic selection in maize as revealed by ancient DNA. Science 302, 1206–1208 [2003].

    ADS  Google Scholar 

  70. Vallebueno-Estrada, M. et al. The earliest maize from San Marcos Tehuacán is a partial domesticate with genomic evidence of inbreeding. Proc. Natl Acad. Sci. USA 113, 14151–14156 [2016].

    Google Scholar 

  71. Ramos-Madrigal, J. et al. Genome sequence of a 5,310-year-old maize cob provides insights into the early stages of maize domestication. Curr. Biol. 26, 3195–3201 [2016].

    Google Scholar 

  72. Ramos-Madrigal, J. et al. Palaeogenomic insights into the origins of French grapevine diversity. Nat. Plants 5, 595–603 [2019].

    Google Scholar 

  73. Weiß, C. L. et al. Temporal patterns of damage and decay kinetics of DNA retrieved from plant herbarium specimens. R. Soc. Open Sci. 3, 160239 [2016].

    ADS  Google Scholar 

  74. Goloubinoff, P., Pääbo, S. & Wilson, A. C. Evolution of maize inferred from sequence diversity of an Adh2 gene segment from archaeological specimens. Proc. Natl Acad. Sci. USA 90, 1997–2001 [1993].

    ADS  Google Scholar 

  75. Mascher, M. et al. Genebank genomics bridges the gap between the conservation of crop diversity and plant breeding. Nat. Genet. 51, 1076–1081 [2019].

    Google Scholar 

  76. Fordyce, S. L. et al. Deep sequencing of RNA from ancient maize kernels. PLoS ONE 8, e50961 [2013].

    ADS  Google Scholar 

  77. Smith, O. et al. Small RNA activity in archeological barley shows novel germination inhibition in response to environment. Mol. Biol. Evol. 34, 2555–2562 [2017].

    Google Scholar 

  78. Goldstein, P. Z. & Desalle, R. Calibrating phyfDabnlogenetic species formation in a threatened insect using DNA from historical specimens. Mol. Ecol. 12, 1993–1998 [2003].

    Google Scholar 

  79. Rawlence, N. J., Wood, J. R., Armstrong, K. N. & Cooper, A. DNA content and distribution in ancient feathers and potential to reconstruct the plumage of extinct avian taxa. Proc. Biol. Sci. 276, 3395–3402 [2009].

    Google Scholar 

  80. Oskam, C. L. et al. Fossil avian eggshell preserves ancient DNA. Proc. Biol. Sci. 277, 1991–2000 [2010].

    Google Scholar 

  81. Bro-Jørgensen, M. H. et al. Ancient DNA analysis of Scandinavian medieval drinking horns and the horn of the last aurochs bull. J. Archaeol. Sci. 99, 47–54 [2018].

    Google Scholar 

  82. Foley, B. P., Hansson, M. C., Kourkoumelis, D. P. & Theodoulou, T. A. Aspects of ancient Greek trade re-evaluated with amphora DNA evidence. J. Archaeol. Sci. 39, 389–398 [2012].

    Google Scholar 

  83. Kashuba, N. et al. Ancient DNA from mastics solidifies connection between material culture and genetics of mesolithic hunter-gatherers in Scandinavia. Commun. Biol. 2, 185 [2019].

    Google Scholar 

  84. Jensen, T. Z. T. et al. A 5700 year-old human genome and oral microbiome from chewed birch pitch. Nat. Commun. 10, 5520 [2019].

    ADS  Google Scholar 

  85. Tito, R. Y. et al. Phylotyping and functional analysis of two ancient human microbiomes. PLoS ONE 3, e3703 [2008].

    ADS  Google Scholar 

  86. Hagan, R. W. et al. Comparison of extraction methods for recovering ancient microbial DNA from paleofeces. Am. J. Phys. Anthropol. 171, 275–284 [2020].

    Google Scholar 

  87. Søe, M. J. et al. Ancient DNA from latrines in Northern Europe and the Middle East [500 BC–1700 AD] reveals past parasites and diet. PLoS ONE 13, e0195481 [2018].

    Google Scholar 

  88. Poinar, H. N. et al. Molecular coproscopy: dung and diet of the extinct ground sloth Nothrotheriops shastensis. Science 281, 402–406 [1998]. This article reports the first application of next-generation DNA sequencing to ancient specimens and describes the metagenomic nature of paleontological remains together with megabase-scale data from the woolly mammoth genome.

    ADS  Google Scholar 

  89. Bon, C. et al. Coprolites as a source of information on the genome and diet of the cave hyena. Proc. Biol. Sci. 279, 2825–2830 [2012].

    Google Scholar 

  90. Willerslev, E. et al. Diverse plant and animal genetic records from Holocene and Pleistocene sediments. Science 300, 791–795 [2003]. This article reports the first analysis of environmental DNA prior to the advent of next-generation DNA sequencing and establishes the long-term DNA persistence of paleocommunities within sediments.

    ADS  Google Scholar 

  91. Willerslev, E. et al. Ancient biomolecules from deep ice cores reveal a forested southern Greenland. Science 317, 111–114 [2007].

    ADS  Google Scholar 

  92. Coolen, M. J. & Overmann, J. Analysis of subfossil molecular remains of purple sulfur bacteria in a lake sediment. Appl. Environ. Microbiol. 64, 4513–4521 [1998].

    Google Scholar 

  93. Bardill, J. et al. Advancing the ethics of paleogenomics. Science 360, 384–385 [2018].

    ADS  Google Scholar 

  94. Prendergast, M. E. & Sawchuk, E. Boots on the ground in Africa’s ancient DNA ‘revolution’: archaeological perspectives on ethics and best practices. Antiquity 92, 803–815 [2018].

    Google Scholar 

  95. Alberti, F. et al. Optimized DNA sampling of ancient bones using computed tomography scans. Mol. Ecol. Resour. 18, 1196–1208 [2018].

    Google Scholar 

  96. Sirak, K. A. et al. A minimally-invasive method for sampling human petrous bones from the cranial base for ancient DNA analysis. BioTechniques 62, 283–289 [2017].

    Google Scholar 

  97. Harney, É. et al. A minimally destructive protocol for DNA extraction from ancient teeth. Preprint at bioRxiv //doi.org/10.1101/2020.08.19.256412[2020].

  98. Patzold, F., Zilli, A. & Hundsdoerfer, A. K. Advantages of an easy-to-use DNA extraction method for minimal-destructive analysis of collection specimens. PLoS ONE 15, e0235222 [2020].

    Google Scholar 

  99. Sugita, N. et al. Non-destructive DNA extraction from herbarium specimens: a method particularly suitable for plants with small and fragile leaves. J. Plant Res. 133, 133–141 [2020].

    Google Scholar 

  100. Austin, R. M., Sholts, S. B., Williams, L., Kistler, L. & Hofman, C. A. Opinion: to curate the molecular past, museums need a carefully considered set of best practices. Proc. Natl Acad. Sci. USA 116, 1471–1474 [2019].

    Google Scholar 

  101. Pálsdóttir, A. H., Bläuer, A., Rannamäe, E., Boessenkool, S. & Hallsson, J. H. Not a limitless resource: ethics and guidelines for destructive sampling of archaeofaunal remains. R. Soc. Open Sci. 6, 191059 [2019].

    ADS  Google Scholar 

  102. Claw, K. G. et al. A framework for enhancing ethical genomic research with Indigenous communities. Nat. Commun. 9, 2957 [2018].

    ADS  Google Scholar 

  103. Wagner, J. K. et al. Fostering responsible research on ancient DNA. Am. J. Hum. Genet. 107, 183–195 [2020].

    Google Scholar 

  104. Eisenmann, S. et al. Reconciling material cultures in archaeology with genetic data: the nomenclature of clusters emerging from archaeogenomic analysis. Sci. Rep. 8, 13003 [2018].

    ADS  Google Scholar 

  105. Frieman, C. J. & Hofmann, D. Present pasts in the archaeology of genetics, identity, and migration in Europe: a critical essay. World Archaeol. 51, 528–545 [2019].

    Google Scholar 

  106. Furholt, M. De-contaminating the aDNA — archaeology dialogue on mobility and migration: discussing the culture-historical legacy. CSA 27, 53–68 [2019].

    Google Scholar 

  107. Hakenbeck, S. E. Genetics, archaeology and the far right: an unholy Trinity. World Archaeol. 51, 517–527 [2019].

    Google Scholar 

  108. Fulton, T. L. & Shapiro, B. Setting up an ancient DNA laboratory. Methods Mol. Biol. 1963, 1–13 [2019].

    Google Scholar 

  109. Dabney, J., Meyer, M. & Pääbo, S. Ancient DNA damage. Cold Spring Harb. Perspect. Biol. 5, a012567 [2013].

    Google Scholar 

  110. Höss, M. & Pääbo, S. DNA extraction from Pleistocene bones by a silica-based purification method. Nucleic Acids Res. 21, 3913–3914 [1993].

    Google Scholar 

  111. Green, R. E. et al. The Neandertal genome and ancient DNA authenticity. EMBO J. 28, 2494–2502 [2009].

    Google Scholar 

  112. Korlević, P., Talamo, S. & Meyer, M. A combined method for DNA analysis and radiocarbon dating from a single sample. Sci. Rep. 8, 4127 [2018].

    ADS  Google Scholar 

  113. Fagernäs, Z. et al. A unified protocol for simultaneous extraction of DNA and proteins from archaeological dental calculus. J. Archaeol. Sci. 118, 105135 [2020].

    Google Scholar 

  114. Rohland, N. & Hofreiter, M. Ancient DNA extraction from bones and teeth. Nat. Protoc. 2, 1756–1762 [2007].

    Google Scholar 

  115. Rohland, N., Glocke, I., Aximu-Petri, A. & Meyer, M. Extraction of highly degraded DNA from ancient bones, teeth and sediments for high-throughput sequencing. Nat. Protoc. 13, 2447–2461 [2018].

    Google Scholar 

  116. Glocke, I. & Meyer, M. Extending the spectrum of DNA sequences retrieved from ancient bones and teeth. Genome Res. 27, 1230–1237 [2017].

    Google Scholar 

  117. Wales, N., Andersen, K., Cappellini, E., Ávila-Arcos, M. C. & Gilbert, M. T. P. Optimization of DNA recovery and amplification from non-carbonized archaeobotanical remains. PLoS ONE 9, e86827 [2014].

    ADS  Google Scholar 

  118. Wales, N. et al. in Ancient DNA: Methods and Protocols [eds. Shapiro, B. et al.] 45–55 [Springer, 2019].

  119. Meyer, M. et al. A mitochondrial genome sequence of a hominin from Sima de los Huesos. Nature 505, 403–406 [2014].

    ADS  Google Scholar 

  120. Höss, M., Jaruga, P., Zastawny, T. H., Dizdaroglu, M. & Pääbo, S. DNA damage and DNA sequence retrieval from ancient tissues. Nucleic Acids Res. 24, 1304–1307 [1996].

    Google Scholar 

  121. Briggs, A. W. et al. Patterns of damage in genomic DNA sequences from a Neandertal. Proc. Natl Acad. Sci. USA 104, 14616–14621 [2007]. This article presents the first statistical model of post-mortem DNA degradation and its impact on nucleotide misincorporation patterns, which provides the basis of important criteria for data authentication.

    ADS  Google Scholar 

  122. Sawyer, S., Krause, J., Guschanski, K., Savolainen, V. & Pääbo, S. Temporal patterns of nucleotide misincorporations and DNA fragmentation in ancient DNA. PLoS ONE 7, e34131 [2012].

    ADS  Google Scholar 

  123. Ho, S. Y. W., Heupink, T. H., Rambaut, A. & Shapiro, B. Bayesian estimation of sequence damage in ancient DNA. Mol. Biol. Evol. 24, 1416–1422 [2007].

    Google Scholar 

  124. Axelsson, E., Willerslev, E., Gilbert, M. T. P. & Nielsen, R. The effect of ancient DNA damage on inferences of demographic histories. Mol. Biol. Evol. 25, 2181–2187 [2008].

    Google Scholar 

  125. Briggs, A. W. et al. Removal of deaminated cytosines and detection of in vivo methylation in ancient DNA. Nucleic Acids Res. 38, e87 [2010]. This article presents an experimental procedure based on enzymatic digestion that eliminates nucleotide misincorporations from ancient sequence data, which provided the basis for robust statistical inference, minimally impacted by spurious allelic variation, and has paved the way to aDNA methylation maps.

    Google Scholar 

  126. Rohland, N., Harney, E., Mallick, S., Nordenfelt, S. & Reich, D. Partial uracil–DNA–glycosylase treatment for screening of ancient DNA. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 20130624 [2015]. This study reports the first method for DNA library preparation that is compatible with full automation and also describes various approaches aimed at authenticating data while minimizing the impact of post-mortem DNA misincorporation on downstream analyses.

    Google Scholar 

  127. Hebert, P. D. N., Cywinska, A., Ball, S. L. & deWaard, J. R. Biological identifications through DNA barcodes. Proc. Biol. Sci. 270, 313–321 [2003].

    Google Scholar 

  128. Taberlet, P. et al. Power and limitations of the chloroplast trnL [UAA] intron for plant DNA barcoding. Nucleic Acids Res. 35, e14 [2007].

    Google Scholar 

  129. Hofreiter, M., Mead, J. I., Martin, P. & Poinar, H. N. Molecular caving. Curr. Biol. 13, R693–R695 [2003].

    Google Scholar 

  130. Parducci, L. et al. Glacial survival of boreal trees in northern Scandinavia. Science 335, 1083–1086 [2012].

    ADS  Google Scholar 

  131. Alsos, I. G. et al. The role of sea ice for vascular plant dispersal in the Arctic. Biol. Lett. 12, 20160264 [2016].

    Google Scholar 

  132. Crump, S. E. et al. Arctic shrub colonization lagged peak postglacial warmth: molecular evidence in lake sediment from Arctic Canada. Glob. Chang. Biol. 25, 4244–4256 [2019].

    ADS  Google Scholar 

  133. Ziesemer, K. A. et al. Intrinsic challenges in ancient microbiome reconstruction using 16S rRNA gene amplification. Sci. Rep. 5, 16498 [2015].

    ADS  Google Scholar 

  134. Smith, O. et al. Archaeology. Sedimentary DNA from a submerged site reveals wheat in the British Isles 8000 years ago. Science 347, 998–1001 [2015].

    ADS  Google Scholar 

  135. Parducci, L. et al. Shotgun environmental DNA, pollen, and macrofossil analysis of lateglacial lake sediments from southern Sweden. Front. Ecol. Evol. 7, 189 [2019].

    Google Scholar 

  136. Gaffney, V. et al. Multi-proxy characterisation of the Storegga tsunami and its impact on the early holocene landscapes of the southern North Sea. Geosciences 10, 270 [2020].

    ADS  Google Scholar 

  137. Cribdon, B., Ware, R., Smith, O., Gaffney, V. & Allaby, R. G. PIA: more accurate taxonomic assignment of metagenomic data demonstrated on sedaDNA from the North Sea. Front. Ecol. Evol. 8, 84 [2020].

    Google Scholar 

  138. Noonan, J. P. et al. Genomic sequencing of Pleistocene cave bears. Science 309, 597–599 [2005].

    ADS  Google Scholar 

  139. Noonan, J. P. et al. Sequencing and analysis of Neanderthal genomic DNA. Science 314, 1113–1118 [2006].

    ADS  Google Scholar 

  140. Bentley, D. R. et al. Accurate whole human genome sequencing using reversible terminator chemistry. Nature 456, 53–59 [2008].

    ADS  Google Scholar 

  141. Meyer, M. & Kircher, M. Illumina sequencing library preparation for highly multiplexed target capture and sequencing. Cold Spring Harb Protoc 2010, pdb.prot5448 [2010].

    Google Scholar 

  142. Kircher, M., Sawyer, S. & Meyer, M. Double indexing overcomes inaccuracies in multiplex sequencing on the Illumina platform. Nucleic Acids Res. 40, e3 [2012].

    Google Scholar 

  143. Gansauge, M.-T. & Meyer, M. Single-stranded DNA library preparation for the sequencing of ancient or damaged DNA. Nat. Protoc. 8, 737–748 [2013].

    Google Scholar 

  144. Gansauge, M.-T., Aximu-Petri, A., Nagel, S. & Meyer, M. Manual and automated preparation of single-stranded DNA libraries for the sequencing of DNA from ancient biological remains and other sources of highly degraded DNA. Nat. Protoc. 15, 2279–2300 [2020].

    Google Scholar 

  145. Gansauge, M.-T. et al. Single-stranded DNA library preparation from highly degraded DNA using T4 DNA ligase. Nucleic Acids Res. 45, e79 [2017].

    Google Scholar 

  146. Harkins, K. M. et al. A novel NGS library preparation method to characterize native termini of fragmented DNA. Nucleic Acids Res. 48, e47 [2020].

    Google Scholar 

  147. Carøe, C. et al. Single-tube library preparation for degraded DNA. Methods Ecol. Evol. 9, 410–419 [2018].

    Google Scholar 

  148. Fages, A. et al. Tracking five millennia of horse management with extensive ancient genome time series. Cell 177, 1419–1435.e31 [2019].

    Google Scholar 

  149. van der Valk, T., Vezzi, F., Ormestad, M., Dalén, L. & Guschanski, K. Index hopping on the Illumina HiseqX platform and its consequences for ancient DNA studies. Mol. Ecol. Resour. 20, 1171–1181 [2019].

    Google Scholar 

  150. Dabney, J. & Meyer, M. Length and GC-biases during sequencing library amplification: a comparison of various polymerase-buffer systems with ancient and modern DNA sequencing libraries. BioTechniques 52, 87–94 [2012].

    Google Scholar 

  151. Seguin-Orlando, A. et al. Amplification of TruSeq ancient DNA libraries with AccuPrime Pfx: consequences on nucleotide misincorporation and methylation patterns. STAR: Sci. Technol. Archaeol. Res. 1, 1–9 [2015].

    Google Scholar 

  152. Meyer, M. et al. From micrograms to picograms: quantitative PCR reduces the material demands of high-throughput sequencing. Nucleic Acids Res. 36, e5 [2008].

    Google Scholar 

  153. Thompson, J. R., Marcelino, L. A. & Polz, M. F. Heteroduplexes in mixed-template amplifications: formation, consequence and elimination by ‘reconditioning PCR’. Nucleic Acids Res. 30, 2083–2088 [2002].

    Google Scholar 

  154. Max Planck Institute for the Science of Human History. MPI-SHH archaeogenetics. protocols.io //www.protocols.io/workspaces/mpishh-archaeogenetics [2020].

  155. Burbano, H. A. et al. Targeted investigation of the Neandertal genome by array-based sequence capture. Science 328, 723–725 [2010].

    ADS  Google Scholar 

  156. Devault, A. M. et al. Second-pandemic strain of Vibrio cholerae from the Philadelphia cholera outbreak of 1849. N. Engl. J. Med. 370, 334–340 [2014].

    Google Scholar 

  157. Enk, J. M. et al. Ancient whole genome enrichment using baits built from modern DNA. Mol. Biol. Evol. 31, 1292–1294 [2014].

    Google Scholar 

  158. Haak, W. et al. Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 522, 207–211 [2015].

    ADS  Google Scholar 

  159. Cruz-Dávalos, D. I. et al. Experimental conditions improving in-solution target enrichment for ancient DNA. Mol. Ecol. Resour. 17, 508–522 [2017].

    Google Scholar 

  160. Mathieson, I. et al. Genome-wide patterns of selection in 230 ancient Eurasians. Nature 528, 499–503 [2015].

    ADS  Google Scholar 

  161. Cruz-Dávalos, D. I. et al. In-solution Y-chromosome capture-enrichment on ancient DNA libraries. BMC Genomics 19, 608 [2018].

    Google Scholar 

  162. Günther, T. & Nettelblad, C. The presence and impact of reference bias on population genomic studies of prehistoric human populations. PLoS Genet. 15, e1008302 [2019].

    Google Scholar 

  163. Lachance, J. & Tishkoff, S. A. SNP ascertainment bias in population genetic analyses: why it is important, and how to correct it. Bioessays 35, 780–786 [2013].

    Google Scholar 

  164. Lang, P. L. M. et al. Hybridization ddRAD-sequencing for population genomics of nonmodel plants using highly degraded historical specimen DNA. Mol. Ecol. Resour. 20, 1228–1247 [2020].

    Google Scholar 

  165. Suchan, T. et al. Hybridization capture using RAD probes [hyRAD], a new tool for performing genomic analyses on collection specimens. PLoS ONE 11, e0151651 [2016].

    Google Scholar 

  166. Hernandez-Rodriguez, J. et al. The impact of endogenous content, replicates and pooling on genome capture from faecal samples. Mol. Ecol. Resour. 18, 319–333 [2018].

    Google Scholar 

  167. Smith, O. et al. Genomic methylation patterns in archaeological barley show de-methylation as a time-dependent diagenetic process. Sci. Rep. 4, 5559 [2014].

    Google Scholar 

  168. Keller, A. et al. New insights into the Tyrolean Iceman’s origin and phenotype as inferred by whole-genome sequencing. Nat. Commun. 3, 698 [2012].

    ADS  Google Scholar 

  169. Mak, S. S. T. et al. Comparative performance of the BGISEQ-500 vs Illumina HiSeq2500 sequencing platforms for palaeogenomic sequencing. Gigascience 6, 1–13 [2017].

    ADS  Google Scholar 

  170. Martins, R. F., Kampmann, M.-L. & Förster, D. W. Sequencing library preparation from degraded samples for non-illumina sequencing platforms. Methods Mol. Biol. 1963, 85–92 [2019].

    Google Scholar 

  171. Kircher, M., Stenzel, U. & Kelso, J. Improved base calling for the Illumina Genome Analyzer using machine learning strategies. Genome Biol. 10, R83 [2009].

    Google Scholar 

  172. Renaud, G., Kircher, M., Stenzel, U. & Kelso, J. freeIbis: an efficient basecaller with calibrated quality scores for Illumina sequencers. Bioinformatics 29, 1208–1209 [2013].

    Google Scholar 

  173. Costello, M. et al. Characterization and remediation of sample index swaps by non-redundant dual indexing on massively parallel sequencing platforms. BMC Genomics 19, 332 [2018].

    Google Scholar 

  174. Dolle, D. et al. CASCADE: a custom-made archiving system for the conservation of ancient DNA experimental data. Front. Ecol. Evol. 8, 185 [2020].

    ADS  Google Scholar 

  175. Renaud, G., Stenzel, U. & Kelso, J. leeHom: adaptor trimming and merging for Illumina sequencing reads. Nucleic Acids Res. 42, e141 [2014].

    Google Scholar 

  176. Schubert, M., Lindgreen, S. & Orlando, L. AdapterRemoval v2: rapid adapter trimming, identification, and read merging. BMC Res. Notes 9, 88 [2016].

    Google Scholar 

  177. Schuenemann, V. J. et al. Genome-wide comparison of medieval and modern Mycobacterium leprae. Science 341, 179–183 [2013].

    ADS  Google Scholar 

  178. Li, H. & Durbin, R. Fast and accurate long-read alignment with Burrows–Wheeler transform. Bioinformatics 26, 589–595 [2010].

    Google Scholar 

  179. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 [2012].

    Google Scholar 

  180. Renaud, G., Hanghøj, K., Willerslev, E. & Orlando, L. gargammel: a sequence simulator for ancient DNA. Bioinformatics 33, 577–579 [2017].

    Google Scholar 

  181. Schubert, M. et al. Improving ancient DNA read mapping against modern reference genomes. BMC Genomics 13, 178 [2012].

    Google Scholar 

  182. Shapiro, B. & Hofreiter, M. A paleogenomic perspective on evolution and gene function: new insights from ancient DNA. Science 343, 1236573 [2014].

    Google Scholar 

  183. Cahill, J. A. et al. Genomic evidence of widespread admixture from polar bears into brown bears during the last ice age. Mol. Biol. Evol. 35, 1120–1129 [2018].

    Google Scholar 

  184. Poullet, M. & Orlando, L. Assessing DNA sequence alignment methods for characterizing ancient genomes and methylomes. Front. Ecol. Evolut. 8, 105 [2020].

    Google Scholar 

  185. Seguin-Orlando, A. et al. Ligation bias in Illumina next-generation DNA libraries: implications for sequencing ancient genomes. PLoS ONE 8, e78575 [2013].

    ADS  Google Scholar 

  186. Garrison, E. et al. Variation graph toolkit improves read mapping by representing genetic variation in the reference. Nat. Biotechnol. 36, 875–879 [2018].

    Google Scholar 

  187. Schubert, M. et al. Characterization of ancient and modern genomes by SNP detection and phylogenomic and metagenomic analysis using PALEOMIX. Nat. Protoc. 9, 1056–1082 [2014].

    Google Scholar 

  188. Peltzer, A. et al. EAGER: efficient ancient genome reconstruction. Genome Biol. 17, 60 [2016].

    Google Scholar 

  189. Renaud, G., Slon, V., Duggan, A. T. & Kelso, J. Schmutzi: estimation of contamination and endogenous mitochondrial consensus calling for ancient DNA. Genome Biol. 16, 224 [2015].

    Google Scholar 

  190. Rasmussen, M. et al. An aboriginal Australian genome reveals separate human dispersals into Asia. Science 334, 94–98 [2011].

    ADS  Google Scholar 

  191. Korneliussen, T. S., Moltke, I., Albrechtsen, A. & Nielsen, R. Calculation of Tajima’s D and other neutrality test statistics from low depth next-generation sequencing data. BMC Bioinforma. 14, 289 [2013].

    Google Scholar 

  192. Peyrègne, S. & Peter, B. M. AuthentiCT: a model of ancient DNA damage to estimate the proportion of present-day DNA contamination. Genome Biol. 21, 246 [2020].

    Google Scholar 

  193. Meyer, M. et al. Nuclear DNA sequences from the Middle Pleistocene Sima de los Huesos hominins. Nature 531, 504–507 [2016].

    ADS  Google Scholar 

  194. Skoglund, P. et al. Origins and genetic legacy of Neolithic farmers and hunter-gatherers in Europe. Science 336, 466–469 [2012].

    ADS  Google Scholar 

  195. Nakatsuka, N. et al. A Paleogenomic reconstruction of the deep population history of the Andes. Cell 181, 1131–1145.e21 [2020].

    Google Scholar 

  196. Skoglund, P., Ersmark, E., Palkopoulou, E. & Dalén, L. Ancient wolf genome reveals an early divergence of domestic dog ancestors and admixture into high-latitude breeds. Curr. Biol. 25, 1515–1519 [2015].

    Google Scholar 

  197. Soraggi, S., Wiuf, C. & Albrechtsen, A. Powerful inference with the D-statistic on low-coverage whole-genome data. G3 8, 551–566 [2018].

    Google Scholar 

  198. Skoglund, P., Storå, J., Götherström, A. & Jakobsson, M. Accurate sex identification of ancient human remains using DNA shotgun sequencing. J. Archaeol. Sci. 40, 4477–4482 [2013].

    Google Scholar 

  199. Mittnik, A., Wang, C.-C., Svoboda, J. & Krause, J. A molecular approach to the sexing of the triple burial at the upper Paleolithic site of Dolní Věstonice. PLoS ONE 11, e0163019 [2016].

    Google Scholar 

  200. Schubert, M. et al. Zonkey: a simple, accurate and sensitive pipeline to genetically identify equine F1-hybrids in archaeological assemblages. J. Archaeol. Sci. 78, 147–157 [2017].

    Google Scholar 

  201. Lipatov, M., Sanjeev, K., Patro, R. & Veeramah, K. Maximum likelihood estimation of biological relatedness from low coverage sequencing data. Preprint at bioRxiv //doi.org/10.1101/023374 [2015].

    Article  Google Scholar 

  202. Kennett, D. J. et al. Archaeogenomic evidence reveals prehistoric matrilineal dynasty. Nat. Commun. 8, 14115 [2017].

    ADS  Google Scholar 

  203. Monroy Kuhn, J. M., Jakobsson, M. & Günther, T. Estimating genetic kin relationships in prehistoric populations. PLoS ONE 13, e0195491 [2018].

    Google Scholar 

  204. Ziesemer, K. A. et al. The efficacy of whole human genome capture on ancient dental calculus and dentin. Am. J. Phys. Anthropol. 168, 496–509 [2019].

    Google Scholar 

  205. Mittnik, A. et al. Kinship-based social inequality in Bronze Age Europe. Science 366, 731–734 [2019].

    ADS  Google Scholar 

  206. Olalde, I. et al. The Beaker phenomenon and the genomic transformation of northwest Europe. Nature 555, 190–196 [2018]; erratum 555, 543 [2018].

    ADS  Google Scholar 

  207. Patterson, N., Price, A. L. & Reich, D. Population structure and eigenanalysis. PLoS Genet. 2, e190 [2006].

    Google Scholar 

  208. McVean, G. A genealogical interpretation of principal components analysis. PLoS Genet. 5, e1000686 [2009].

    Google Scholar 

  209. Meisner, J. & Albrechtsen, A. Inferring population structure and admixture proportions in low-depth NGS data. Genetics 210, 719–731 [2018].

    Google Scholar 

  210. Lazaridis, I. et al. Ancient human genomes suggest three ancestral populations for present-day Europeans. Nature 513, 409–413 [2014].

    ADS  Google Scholar 

  211. Engelhardt, B. E. & Stephens, M. Analysis of population structure: a unifying framework and novel methods based on sparse factor analysis. PLoS Genet. 6, e1001117 [2010].

    Google Scholar 

  212. Skoglund, P., Sjödin, P., Skoglund, T., Lascoux, M. & Jakobsson, M. Investigating population history using temporal genetic differentiation. Mol. Biol. Evol. 31, 2516–2527 [2014].

    Google Scholar 

  213. Malaspinas, A.-S. et al. bammds: a tool for assessing the ancestry of low-depth whole-genome data using multidimensional scaling [MDS]. Bioinformatics 30, 2962–2964 [2014].

    Google Scholar 

  214. Alexander, D. H., Novembre, J. & Lange, K. Fast model-based estimation of ancestry in unrelated individuals. Genome Res. 19, 1655–1664 [2009].

    Google Scholar 

  215. Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155, 945–959 [2000].

    Google Scholar 

  216. Skotte, L., Korneliussen, T. S. & Albrechtsen, A. Estimating individual admixture proportions from next generation sequencing data. Genetics 195, 693–702 [2013].

    Google Scholar 

  217. Joseph, T. A. & Pe’er, I. Inference of population structure from time-series genotype data. Am. J. Hum. Genet. 105, 317–333 [2019].

    Google Scholar 

  218. Reich, D., Thangaraj, K., Patterson, N., Price, A. L. & Singh, L. Reconstructing Indian population history. Nature 461, 489–494 [2009].

    ADS  Google Scholar 

  219. Patterson, N. et al. Ancient admixture in human history. Genetics 192, 1065–1093 [2012]. This article describes the statistical methodology forming the core of most subsequent aDNA studies and aims at investigating population structure and admixture patterns.

    Google Scholar 

  220. Raghavan, M. et al. Upper Palaeolithic Siberian genome reveals dual ancestry of native Americans. Nature 505, 87–91 [2014].

    ADS  Google Scholar 

  221. Reich, D. et al. Denisova admixture and the first modern human dispersals into Southeast Asia and Oceania. Am. J. Hum. Genet. 89, 516–528 [2011].

    Google Scholar 

  222. Cassidy, L. M. et al. A dynastic elite in monumental Neolithic society. Nature 582, 384–388 [2020].

    ADS  Google Scholar 

  223. McCarthy, S. et al. A reference panel of 64,976 haplotypes for genotype imputation. Nat. Genet. 48, 1279–1283 [2016].

    Google Scholar 

  224. Antonio, M. L. et al. Ancient Rome: a genetic crossroads of Europe and the Mediterranean. Science 366, 708–714 [2019].

    ADS  Google Scholar 

  225. Sirugo, G., Williams, S. M. & Tishkoff, S. A. The missing diversity in human genetic studies. Cell 177, 26–31 [2019].

    Google Scholar 

  226. Allentoft, M. E. et al. Population genomics of Bronze Age Eurasia. Nature 522, 167–172 [2015].

    ADS  Google Scholar 

  227. Scheib, C. L. et al. East Anglian early Neolithic monument burial linked to contemporary Megaliths. Ann. Hum. Biol. 46, 145–149 [2019].

    Google Scholar 

  228. Ringbauer, H., Novembre, J. & Steinrücken, M. Detecting runs of homozygosity from low-coverage ancient DNA. Preprint at bioRxiv //doi.org/10.1101/2020.05.31.126912 [2020].

    Article  Google Scholar 

  229. Lawson, D. J., Hellenthal, G., Myers, S. & Falush, D. Inference of population structure using dense haplotype data. PLoS Genet. 8, e1002453 [2012].

    Google Scholar 

  230. Burger, J. et al. Low prevalence of lactase persistence in Bronze Age Europe indicates ongoing strong selection over the last 3,000 years. Curr. Biol. 30, 1–9 [2020].

    Google Scholar 

  231. Maixner, F. et al. The 5,300-year-old Helicobacter pylori genome of the Iceman. Science 351, 162–165 [2016].

    ADS  Google Scholar 

  232. Lugli, G. A. et al. Ancient bacteria of the Ötzi’s microbiome: a genomic tale from the Copper Age. Microbiome 5, 5 [2017].

    Google Scholar 

  233. Borry, M. et al. CoproID predicts the source of coprolites and paleofeces using microbiome composition and host DNA content. PeerJ 8, e9001 [2020].

    Google Scholar 

  234. Rifkin, R. F. et al. Multi-proxy analyses of a mid-15th century Middle Iron Age Bantu-speaker palaeo-faecal specimen elucidates the configuration of the ‘ancestral’ sub-Saharan African intestinal microbiome. Microbiome 8, 62 [2020].

    Google Scholar 

  235. Tett, A. et al. The Prevotella copri complex comprises four distinct clades underrepresented in westernized populations. Cell Host Microbe 26, 666–679.e7 [2019].

    Google Scholar 

  236. Vågene, Å. J. et al. Salmonella enterica genomes from victims of a major sixteenth-century epidemic in Mexico. Nat. Ecol. Evol. 2, 520–528 [2018].

    Google Scholar 

  237. Wood, D. E. & Salzberg, S. L. Kraken: ultrafast metagenomic sequence classification using exact alignments. Genome Biol. 15, R46 [2014].

    Google Scholar 

  238. Velsko, I. M., Frantz, L. A. F., Herbig, A., Larson, G. & Warinner, C. Selection of appropriate metagenome taxonomic classifiers for ancient microbiome research. mSystems 3, e00080-18 [2018].

    Google Scholar 

  239. Eisenhofer, R. & Weyrich, L. S. Assessing alignment-based taxonomic classification of ancient microbial DNA. PeerJ 7, e6594 [2019].

    Google Scholar 

  240. Segata, N. et al. Metagenomic microbial community profiling using unique clade-specific marker genes. Nat. Methods 9, 811–814 [2012].

    Google Scholar 

  241. Hübler, R. et al. HOPS: automated detection and authentication of pathogen DNA in archaeological remains. Genome Biol. 20, 280 [2019].

    Google Scholar 

  242. Louvel, G., Der Sarkissian, C., Hanghøj, K. & Orlando, L. metaBIT, an integrative and automated metagenomic pipeline for analysing microbial profiles from high-throughput sequencing shotgun data. Mol. Ecol. Resour. 16, 1415–1427 [2016].

    Google Scholar 

  243. Knights, D. et al. Bayesian community-wide culture-independent microbial source tracking. Nat. Methods 8, 761–763 [2011].

    Google Scholar 

  244. Llamas, B. et al. High-resolution analysis of cytosine methylation in ancient DNA. PLoS ONE 7, e30226 [2012].

    ADS  Google Scholar 

  245. Smith, R. W. A., Monroe, C. & Bolnick, D. A. Detection of cytosine methylation in ancient DNA from five Native American populations using bisulfite sequencing. PLoS ONE 10, e0125344 [2015].

    Google Scholar 

  246. Wagner, S., Plomion, C. & Orlando, L. Uncovering signatures of DNA methylation in ancient plant remains from patterns of post-mortem DNA damage. Front. Ecol. Evol. 8, 11 [2020].

    Google Scholar 

  247. Hanghøj, K. et al. Fast, accurate and automatic ancient nucleosome and methylation maps with epiPALEOMIX. Mol. Biol. Evol. 33, 3284–3298 [2016].

    Google Scholar 

  248. Hanghøj, K., Renaud, G., Albrechtsen, A. & Orlando, L. DamMet: ancient methylome mapping accounting for errors, true variants, and post-mortem DNA damage. Gigascience 8, giz025 [2019].

    Google Scholar 

  249. Pääbo, S. et al. Genetic analyses from ancient DNA. Annu. Rev. Genet. 38, 645–679 [2004].

    Google Scholar 

  250. Racimo, F., Marnetto, D. & Huerta-Sánchez, E. Signatures of archaic adaptive introgression in present-day human populations. Mol. Biol. Evol. 34, 296–317 [2017].

    Google Scholar 

  251. de Barros Damgaard, P. et al. The first horse herders and the impact of early Bronze Age steppe expansions into Asia. Science 360, eaar7711 [2018].

    Google Scholar 

  252. Damgaard, PdeB. et al. 137 ancient human genomes from across the Eurasian steppes. Nature 557, 369–374 [2018]; author correction 563, E16 [2018].

    ADS  Google Scholar 

  253. Narasimhan, V. M. et al. The formation of human populations in South and Central Asia. Science 365, eaat7487 [2019].

    Google Scholar 

  254. Seguin-Orlando, A. et al. Paleogenomics. Genomic structure in Europeans dating back at least 36,200 years. Science 346, 1113–1118 [2014].

    ADS  Google Scholar 

  255. Sikora, M. et al. Ancient genomes show social and reproductive behavior of early Upper Paleolithic foragers. Science 358, 659–662 [2017].

    ADS  Google Scholar 

  256. Fu, Q. et al. Genome sequence of a 45,000-year-old modern human from western Siberia. Nature 514, 445–449 [2014].

    ADS  Google Scholar 

  257. Yang, M. A. et al. 40,000-year-old individual from Asia provides insight into early population structure in Eurasia. Curr. Biol. 27, 3202–3208.e9 [2017].

    Google Scholar 

  258. Olalde, I. & Posth, C. Latest trends in archaeogenetic research of west Eurasians. Curr. Opin. Genet. Dev. 62, 36–43 [2020].

    Google Scholar 

  259. Yang, M. A. et al. Ancient DNA indicates human population shifts and admixture in northern and southern China. Science 369, 282–288 [2020].

    ADS  Google Scholar 

  260. Lipson, M. et al. Population turnover in remote Oceania shortly after initial settlement. Curr. Biol. 28, 1157–1165.e7 [2018].

    Google Scholar 

  261. Lipson, M. et al. Ancient genomes document multiple waves of migration in Southeast Asian prehistory. Science 361, 92–95 [2018].

    ADS  Google Scholar 

  262. McColl, H. et al. The prehistoric peopling of Southeast Asia. Science 361, 88–92 [2018].

    ADS  Google Scholar 

  263. Lazaridis, I. et al. Genetic origins of the Minoans and Mycenaeans. Nature 548, 214–218 [2017].

    ADS  Google Scholar 

  264. Wang, K. et al. Ancient genomes reveal complex patterns of population movement, interaction, and replacement in sub-Saharan Africa. Sci. Adv. 6, eaaz0183 [2020].

    ADS  Google Scholar 

  265. Sánchez-Quinto, F. et al. Genomic affinities of two 7,000-year-old Iberian hunter-gatherers. Curr. Biol. 22, 1494–1499 [2012].

    Google Scholar 

  266. Lipson, M. et al. Parallel palaeogenomic transects reveal complex genetic history of early European farmers. Nature 551, 368–372 [2017].

    ADS  Google Scholar 

  267. Jeong, C. et al. The genetic history of admixture across inner Eurasia. Nat. Ecol. Evol. 3, 966–976 [2019].

    Google Scholar 

  268. Moreno-Mayar, J. V. et al. Terminal Pleistocene Alaskan genome reveals first founding population of Native Americans. Nature 553, 203–207 [2018].

    ADS  Google Scholar 

  269. Moreno-Mayar, J. V. et al. Early human dispersals within the Americas. Science 362, eaav2621 [2018].

    ADS  Google Scholar 

  270. Flegontov, P. et al. Palaeo-Eskimo genetic ancestry and the peopling of Chukotka and North America. Nature 570, 236–240 [2019].

    ADS  Google Scholar 

  271. Nieves-Colón, M. A. et al. Ancient DNA reconstructs the genetic legacies of precontact Puerto Rico communities. Mol. Biol. Evol. 37, 611–626 [2020].

    Google Scholar 

  272. Nägele, K. et al. Genomic insights into the early peopling of the Caribbean. Science 369, 456–460 [2020].

    ADS  Google Scholar 

  273. Schroeder, H. et al. Genome-wide ancestry of 17th-century enslaved Africans from the Caribbean. Proc. Natl Acad. Sci. USA 112, 3669–3673 [2015].

    ADS  Google Scholar 

  274. Barquera, R. et al. Origin and health status of first-generation Africans from early Colonial Mexico. Curr. Biol. 30, 2078–2091.e11 [2020].

    Google Scholar 

  275. Larson, G. et al. Current perspectives and the future of domestication studies. Proc. Natl Acad. Sci. USA 111, 6139–6146 [2014].

    ADS  Google Scholar 

  276. Larson, G. et al. Phylogeny and ancient DNA of Sus provides insights into neolithic expansion in island Southeast Asia and Oceania. Proc. Natl Acad. Sci. USA 104, 4834–4839 [2007].

    ADS  Google Scholar 

  277. Kistler, L. et al. Ancient plant genomics in archaeology, herbaria, and the environment. Annu. Rev. Plant. Biol. 71, 605–629 [2020].

    Google Scholar 

  278. Ní Leathlobhair, M. et al. The evolutionary history of dogs in the Americas. Science 361, 81–85 [2018].

    ADS  Google Scholar 

  279. Frantz, L. A. F. et al. Genomic and archaeological evidence suggest a dual origin of domestic dogs. Science 352, 1228–1231 [2016].

    ADS  Google Scholar 

  280. Sinding, M.-H. S. et al. Arctic-adapted dogs emerged at the Pleistocene–Holocene transition. Science 368, 1495–1499 [2020].

    ADS  Google Scholar 

  281. Gaunitz, C. et al. Ancient genomes revisit the ancestry of domestic and Przewalski’s horses. Science 360, 111–114 [2018].

    ADS  Google Scholar 

  282. Frantz, L. A. F. et al. Ancient pigs reveal a near-complete genomic turnover following their introduction to Europe. Proc. Natl Acad. Sci. USA 116, 17231–17238 [2019].

    Google Scholar 

  283. Gutaker, R. M. et al. The origins and adaptation of European potatoes reconstructed from historical genomes. Nat. Ecol. Evol. 3, 1093–1101 [2019].

    Google Scholar 

  284. Ameen, C. et al. Specialized sledge dogs accompanied Inuit dispersal across the North American Arctic. Proc. R. Soc. Lond. B Biol. Sci. 286, 20191929 [2019].

    Google Scholar 

  285. da Fonseca, R. R. et al. The origin and evolution of maize in the Southwestern United States. Nat. Plants 1, 14003 [2015].

    MathSciNet  Google Scholar 

  286. Gutaker, R. M. et al. Flax latitudinal adaptation at LuTFL1 altered architecture and promoted fiber production. Sci. Rep. 9, 976 [2019].

    ADS  Google Scholar 

  287. Park, S. D. E. et al. Genome sequencing of the extinct Eurasian wild aurochs, Bos primigenius, illuminates the phylogeography and evolution of cattle. Genome Biol. 16, 234 [2015].

    Google Scholar 

  288. Verdugo, M. P. et al. Ancient cattle genomics, origins, and rapid turnover in the Fertile Crescent. Science 365, 173–176 [2019].

    ADS  Google Scholar 

  289. Daly, K. G. et al. Ancient goat genomes reveal mosaic domestication in the Fertile Crescent. Science 361, 85–88 [2018].

    ADS  Google Scholar 

  290. Kistler, L. et al. Multiproxy evidence highlights a complex evolutionary legacy of maize in South America. Science 362, 1309–1313 [2018].

    ADS  Google Scholar 

  291. Allaby, R. G., Ware, R. L. & Kistler, L. A re-evaluation of the domestication bottleneck from archaeogenomic evidence. Evol. Appl. 12, 29–37 [2019].

    Google Scholar 

  292. Scott, M. F. et al. A 3,000-year-old Egyptian emmer wheat genome reveals dispersal and domestication history. Nat. Plants 5, 1120–1128 [2019].

    Google Scholar 

  293. Wales, N. et al. Ancient DNA reveals the timing and persistence of organellar genetic bottlenecks over 3,000 years of sunflower domestication and improvement. Evol. Appl. 12, 38–53 [2018].

    Google Scholar 

  294. Smith, O. et al. A domestication history of dynamic adaptation and genomic deterioration in sorghum. Nat. Plants 5, 369–379 [2019].

    Google Scholar 

  295. Swarts, K. et al. Genomic estimation of complex traits reveals ancient maize adaptation to temperate North America. Science 357, 512–515 [2017].

    ADS  Google Scholar 

  296. Girdland Flink, L. et al. Establishing the validity of domestication genes using DNA from ancient chickens. Proc. Natl Acad. Sci. USA 111, 6184–6189 [2014].

    ADS  Google Scholar 

  297. Loog, L. et al. Inferring allele frequency trajectories from ancient DNA indicates that selection on a chicken gene coincided with changes in medieval husbandry practices. Mol. Biol. Evol. 34, 1981–1990 [2017].

    Google Scholar 

  298. Librado, P. et al. Ancient genomic changes associated with domestication of the horse. Science 356, 442–445 [2017].

    ADS  Google Scholar 

  299. Salo, W. L., Aufderheide, A. C., Buikstra, J. & Holcomb, T. A. Identification of Mycobacterium tuberculosis DNA in a pre-Columbian Peruvian mummy. Proc. Natl Acad. Sci. USA 91, 2091–2094 [1994].

    ADS  Google Scholar 

  300. Taubenberger, J. K. et al. Characterization of the 1918 influenza virus polymerase genes. Nature 437, 889–893 [2005].

    ADS  Google Scholar 

  301. Wilbur, A. K. et al. Deficiencies and challenges in the study of ancient tuberculosis DNA. J. Archaeol. Sci. 36, 1990–1997 [2009].

    Google Scholar 

  302. Wagner, D. M. et al. Yersinia pestis and the plague of Justinian 541–543 AD: a genomic analysis. Lancet Infect. Dis. 14, 319–326 [2014].

    Google Scholar 

  303. Spyrou, M. A. et al. Phylogeography of the second plague pandemic revealed through analysis of historical Yersinia pestis genomes. Nat. Commun. 10, 4470 [2019].

    ADS  Google Scholar 

  304. Keller, M. et al. Ancient Yersinia pestis genomes from across Western Europe reveal early diversification during the first Pandemic [541–750]. Proc. Natl Acad. Sci. USA 116, 12363–12372 [2019].

    Google Scholar 

  305. Susat, J. et al. Yersinia pestis strains from Latvia show depletion of the pla virulence gene at the end of the second plague pandemic. Sci. Rep. 10, 14628 [2020].

    ADS  Google Scholar 

  306. Bos, K. I. et al. Pre-Columbian mycobacterial genomes reveal seals as a source of New World human tuberculosis. Nature 514, 494–497 [2014].

    ADS  Google Scholar 

  307. Sabin, S. et al. A seventeenth-century Mycobacterium tuberculosis genome supports a Neolithic emergence of the Mycobacterium tuberculosis complex. Genome Biol. 21, 201 [2020].

    Google Scholar 

  308. Shillito, L.-M., Blong, J. C., Green, E. J. & van Asperen, E. N. The what, how and why of archaeological coprolite analysis. Earth Sci. Rev. 207, 103196 [2020].

    Google Scholar 

  309. Warinner, C., Speller, C. & Collins, M. J. A new era in palaeomicrobiology: prospects for ancient dental calculus as a long-term record of the human oral microbiome. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 20130376 [2015].

    Google Scholar 

  310. Velsko, I. M. et al. Microbial differences between dental plaque and historic dental calculus are related to oral biofilm maturation stage. Microbiome 7, 102 [2019].

    Google Scholar 

  311. Farrer, A. G. et al. Biological and cultural drivers of oral microbiota in Medieval and Post-Medieval London, UK. Preprint at bioRxiv //doi.org/10.1101/343889 [2018].

    Article  Google Scholar 

  312. Ho, S. Y. W. & Shapiro, B. Skyline-plot methods for estimating demographic history from nucleotide sequences. Mol. Ecol. Resour. 11, 423–434 [2011].

    Google Scholar 

  313. Allentoft, M. E. et al. Extinct New Zealand megafauna were not in decline before human colonization. Proc. Natl Acad. Sci. USA 111, 4922–4927 [2014].

    ADS  Google Scholar 

  314. Thomas, J. E. et al. Demographic reconstruction from ancient DNA supports rapid extinction of the great auk. eLife 8, e47509 [2019].

    Google Scholar 

  315. Dussex, N. et al. Complete genomes of two extinct New Zealand passerines show responses to climate fluctuations but no evidence for genomic erosion prior to extinction. Biol. Lett. 15, 20190491 [2019].

    Google Scholar 

  316. Li, H. & Durbin, R. Inference of human population history from whole genome sequence of a single individual. Nature 475, 493–496 [2011].

    Google Scholar 

  317. Schiffels, S. & Durbin, R. Inferring human population size and separation history from multiple genome sequences. Nat. Genet. 46, 919–925 [2014].

    Google Scholar 

  318. Murray, G. G. R. et al. Natural selection shaped the rise and fall of passenger pigeon genomic diversity. Science 358, 951–954 [2017].

    ADS  Google Scholar 

  319. Díez-Del-Molino, D., Sánchez-Barreiro, F., Barnes, I., Gilbert, M. T. P. & Dalén, L. Quantifying temporal genomic erosion in endangered species. Trends Ecol. Evol. 33, 176–185 [2018].

    Google Scholar 

  320. Der Sarkissian, C. et al. Evolutionary genomics and conservation of the endangered Przewalski’s horse. Curr. Biol. 25, 2577–2583 [2015].

    Google Scholar 

  321. Norén, K., Godoy, E., Dalén, L., Meijer, T. & Angerbjörn, A. Inbreeding depression in a critically endangered carnivore. Mol. Ecol. 25, 3309–3318 [2016].

    Google Scholar 

  322. Pedersen, M. W. et al. Ancient and modern environmental DNA. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 20130383 [2015].

    Google Scholar 

  323. Haile, J. et al. Ancient DNA chronology within sediment deposits: are paleobiological reconstructions possible and is DNA leaching a factor? Mol. Biol. Evol. 24, 982–989 [2007].

    Google Scholar 

  324. Binladen, J. et al. The use of coded PCR primers enables high-throughput sequencing of multiple homolog amplification products by 454 parallel sequencing. PLoS ONE 2, e197 [2007].

    ADS  Google Scholar 

  325. Murchie, T. J. et al. Optimizing extraction and targeted capture of ancient environmental DNA for reconstructing past environments using the PalaeoChip Arctic-1.0 bait-set. Quat. Res. //doi.org/10.1017/qua.2020.59 [2020].

  326. Haile, J. et al. Ancient DNA reveals late survival of mammoth and horse in interior Alaska. Proc. Natl Acad. Sci. USA 106, 22352–22357 [2009].

    ADS  Google Scholar 

  327. Seersholm, F. V. et al. Rapid range shifts and megafaunal extinctions associated with late Pleistocene climate change. Nat. Commun. 11, 2770 [2020].

    ADS  Google Scholar 

  328. Key, F. M., Posth, C., Krause, J., Herbig, A. & Bos, K. I. Mining metagenomic data sets for ancient DNA: recommended protocols for authentication. Trends Genet. 33, 508–520 [2017].

    Google Scholar 

  329. Günther, T. & Jakobsson, M. in Handbook of Statistical Genomics [eds Balding, D., Moltke, I. & Marioni, J.] 295–340 [Wiley, 2019].

  330. Skourtanioti, E. et al. Genomic history of Neolithic to bronze age Anatolia, Northern Levant, and Southern Caucasus. Cell 181, 1158–1175.e28 [2020].

    Google Scholar 

  331. Burmeister, S. Archaeological research on migration as a multidisciplinary challenge. medieval worlds 4, 42–64 [2016].

    Google Scholar 

  332. Knipper, C. et al. Female exogamy and gene pool diversification at the transition from the final Neolithic to the early bronze age in central Europe. Proc. Natl Acad. Sci. USA 114, 10083–10088 [2017].

    Google Scholar 

  333. Zarrillo, S. et al. The use and domestication of Theobroma cacao during the mid-Holocene in the upper Amazon. Nat. Ecol. Evol. 2, 1879–1888 [2018].

    Google Scholar 

  334. Hendy, J. et al. Proteomic evidence of dietary sources in ancient dental calculus. Proc. Biol. Sci. 285, 20180977 [2018].

    Google Scholar 

  335. Roberts, P. et al. Calling all archaeologists: guidelines for terminology, methodology, data handling, and reporting when undertaking and reviewing stable isotope applications in archaeology. Rapid Commun. Mass. Spectrom. 32, 361–372 [2018].

    ADS  Google Scholar 

  336. Radini, A. et al. Medieval women’s early involvement in manuscript production suggested by lapis lazuli identification in dental calculus. Sci. Adv. 5, eaau7126 [2019].

    ADS  Google Scholar 

  337. Ren, M. et al. The origins of cannabis smoking: chemical residue evidence from the first millennium BCE in the Pamirs. Sci. Adv. 5, eaaw1391 [2019].

    ADS  Google Scholar 

  338. Reitsema, L. J. & Holder, S. Stable isotope analysis and the study of human stress, disease, and nutrition. Bioarc. Int. 2, 63–74 [2018].

    Google Scholar 

  339. Maixner, F. et al. The Iceman’s last meal consisted of fat, wild meat, and cereals. Curr. Biol. 28, 2348–2355.e9 [2018].

    Google Scholar 

  340. Lejzerowicz, F. et al. Ancient DNA complements microfossil record in deep-sea subsurface sediments. Biol. Lett. 9, 20130283 [2013].

    Google Scholar 

  341. van Dijk, E. L., Jaszczyszyn, Y., Naquin, D. & Thermes, C. The third revolution in sequencing technology. Trends Genet. 34, 666–681 [2018].

    Google Scholar 

  342. Komor, A. C., Badran, A. H. & Liu, D. R. CRISPR-based technologies for the manipulation of eukaryotic genomes. Cell 168, 20–36 [2017].

    Google Scholar 

  343. Li, H. L., Gee, P., Ishida, K. & Hotta, A. Efficient genomic correction methods in human iPS cells using CRISPR–Cas9 system. Methods 101, 27–35 [2016].

    Google Scholar 

  344. Muchnik, S. K., Lorente-Galdos, B., Santpere, G. & Sestan, N. Modeling the evolution of human brain development using organoids. Cell 179, 1250–1253 [2019].

    Google Scholar 

  345. Kanton, S. et al. Organoid single-cell genomic atlas uncovers human-specific features of brain development. Nature 574, 418–422 [2019].

    ADS  Google Scholar 

  346. Racimo, F., Renaud, G. & Slatkin, M. Joint estimation of contamination, error and demography for nuclear DNA from ancient humans. PLoS Genet. 12, e1005972 [2016].

    Google Scholar 

  347. Jun, G. et al. Detecting and estimating contamination of human DNA samples in sequencing and array-based genotype data. Am. J. Hum. Genet. 91, 839–848 [2012].

    Google Scholar 

  348. Link, V. et al. ATLAS: analysis tools for low-depth and ancient samples. Preprint at bioRxiv //doi.org/10.1101/105346 [2017].

    Article  Google Scholar 

  349. Librado, P. & Orlando, L. Detecting signatures of positive selection along defined branches of a population tree using LSD. Mol. Biol. Evol. 35, 1520–1535 [2018].

    Google Scholar 

  350. Refoyo-Martínez, A. et al. Identifying loci under positive selection in complex population histories. Genome Res. 29, 1506–1520 [2019].

    Google Scholar 

  351. Renaud, G., Hanghøj, K., Korneliussen, T. S., Willerslev, E. & Orlando, L. Joint estimates of heterozygosity and runs of homozygosity for modern and ancient samples. Genetics 212, 587–614 [2019].

    Google Scholar 

  352. Brace, S. et al. Ancient genomes indicate population replacement in early neolithic britain. Nat. Ecol. Evol. 3, 765–771 [2019].

    Google Scholar 

  353. Yates, J. A. F. et al. Community-curated and standardised metadata of published ancient metagenomic samples with AncientMetagenomeDir. Preprint at bioRxiv //doi.org/10.1101/2020.09.02.279570 [2020].

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank A. Hübner for assistance with figure 4c. L.O., P.S., P.W.S. and C.W. received funding from the European Research Council [ERC] under the European Union’s Horizon 2020 research and innovation programme [grant agreements ERC-2015-CoG 681605-PEGASUS, ERC-2018-StG 852558-AGRICON, ERC-2015-StG 678901-FoodTransforms and ERC-2017-StG 804844-DAIRYCULTURES, respectively]. L.O. was also supported by ANR [LifeChange] and the Simone et Cino Del Duca Foundation [HealthTimeTravel]. P.S. was also supported by the Francis Crick Institute core funding [FC001595] from Cancer Research UK, the UK Medical Research Council and the Wellcome Trust, a Wellcome Trust Investigator award [217223/Z/19/Z] and the Vallee Foundation. C.W. also received funding from the Max Planck Society, the Deutsche Forschungsgemeinschaft [EXC 2051 #390713860] and the Siemens Foundation [Paleobiochemistry].

Author information

Authors and Affiliations

  1. Centre d’Anthropobiologie et de Génomique de Toulouse, Université Paul Sabatier Toulouse III, Faculté de Médecine Purpan, Toulouse, France

    Ludovic Orlando & Clio Der Sarkissian

  2. School of Life Sciences, University of Warwick, Coventry, UK

    Robin Allaby

  3. Ancient Genomics Laboratory, Francis Crick Institute, London, UK

    Pontus Skoglund

  4. Institute for Pre- and Protohistoric Archaeology and Archaeology of the Roman Provinces, Ludwig Maximilian University, Munich, Germany

    Philipp W. Stockhammer

  5. Department of Archaeogenetics, Max Planck Institute for the Science of Human History, Jena, Germany

    Philipp W. Stockhammer, Johannes Krause & Christina Warinner

  6. International Laboratory for Human Genome Research, National Autonomous University of Mexico, Queretaro, Mexico

    María C. Ávila-Arcos

  7. Key Laboratory of Vertebrate Evolution and Human Origins, Institute of Vertebrate Paleontology and Paleoanthropology, Center for Excellence in Life and Paleoenvironment, Chinese Academy of Sciences, Beijing, China

    Qiaomei Fu

  8. Lundbeck Foundation GeoGenetics Center, GLOBE Institute, University of Copenhagen, Copenhagen, Denmark

    Eske Willerslev

  9. Welcome Trust, Sanger Institute, Hinxton, UK

    Eske Willerslev

  10. The Danish Institute for Advanced Study, University of Southern Denmark, Odense, Denmark

    Eske Willerslev

  11. Department of Zoology, University of Cambridge, Cambridge, UK

    Eske Willerslev

  12. School of Human Evolution and Social Change, Arizona State University, Tempe, AZ, USA

    Anne C. Stone

  13. Department of Anthropology, Harvard University, Cambridge, MA, USA

    Christina Warinner

Authors

  1. Ludovic Orlando

    View author publications

    You can also search for this author in PubMed Google Scholar

  2. Robin Allaby

    View author publications

    You can also search for this author in PubMed Google Scholar

  3. Pontus Skoglund

    View author publications

    You can also search for this author in PubMed Google Scholar

  4. Clio Der Sarkissian

    View author publications

    You can also search for this author in PubMed Google Scholar

  5. Philipp W. Stockhammer

    View author publications

    You can also search for this author in PubMed Google Scholar

  6. María C. Ávila-Arcos

    View author publications

    You can also search for this author in PubMed Google Scholar

  7. Qiaomei Fu

    View author publications

    You can also search for this author in PubMed Google Scholar

  8. Johannes Krause

    View author publications

    You can also search for this author in PubMed Google Scholar

  9. Eske Willerslev

    View author publications

    You can also search for this author in PubMed Google Scholar

  10. Anne C. Stone

    View author publications

    You can also search for this author in PubMed Google Scholar

  11. Christina Warinner

    View author publications

    You can also search for this author in PubMed Google Scholar

Contributions

Introduction [L.O., A.C.S. and C.W.]; Experimentation [L.O., R.A., C.D.S., P.W.S., A.C.S. and C.W.]; Results [L.O., P.S., R.A., P.W.S., M.C.A.-A. and C.W.]; Applications [L.O., R.A., P.W.S., C.D.S., M.C.A.-A., Q.F., J.K., E.W., A.C.S. and C.W.]; Reproducibility and data deposition [L.O. and M.C.A.-A.]; Limitations and optimizations [L.O.]; Outlook [L.O., P.W.S., A.C.S. and C.W.]; Overview of the Primer [L.O. and C.W.].

Corresponding author

Correspondence to Ludovic Orlando.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information

Nature Reviews Methods Primers thanks T. Günther, L. Matisoo-Smith, R. Pinhasi, N. Rawlence, A. Zink and the other, anonymous, reviewer[s] for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Related links

Ancient Human DNA uMap: //umap.openstreetmap.fr/en/map/ancient-human-dna_41837#6/51.000/2.000

Bitbucket: //bitbucket.org/product

European Nucleotide Archive [ENA]: //www.ebi.ac.uk/ena

GitHub: //github.com

International Symposium for Biomolecular Archaeology [ISBA]: //isba9.sciencesconf.org

Max Planck Harvard Research Center for the Archaeoscience of the Ancient Mediterranean: //www.archaeoscience.org/

Protocols.io: //www.protocols.io

Reich laboratory: //reich.hms.harvard.edu/datasets

Sequence Read Archive [SRA]: //www.ncbi.nlm.nih.gov/sra

Standards and Precautions and Advances in Ancient Metagenomics [SPAAM]: //github.com/SPAAM-workshop

Glossary

Ancient DNA

[aDNA]. Ultrashort and degraded DNA fragments that are preserved in subfossil material, including hard tissues, such as bones, teeth and shells, and soft tissues, such as mummified skin and hair, as well as sediments.

Holobiomes

The total sum of the DNA fragments making up the genome of a host organism and all of its microbiota.

DNA library

A molecular construction in which DNA fragments are ligated to DNA adapters of known sequences in order to be amplified and optionally captured prior to sequencing; different sequencing platforms require different library constructs.

DNA barcoding

The taxonomic assignment of metagenomic DNA content on the basis of DNA fragments that show limited intra-specific sequence diversity but large inter-specific sequence diversity.

Shotgun sequencing

Non-targeted sequencing of DNA library content.

DNA ligases

A class of enzymes that are capable of stitching together different DNA fragments.

Ascertainment bias

Statistical bias resulting from the collection of genetic data at a subset of loci that do not reflect the overall genetic diversity present at the whole-genome scale.

Demultiplexing

A process by which pools of sequences originating from different DNA libraries are assigned back to their original samples on the basis of short synthetic sequences added during library indexing.

Outgroup

An individual, a population or a group of populations and/or species that are genetically close but different from those under study.

Identity by descent

DNA segments between two or more individuals are identical by descent when they are inherited from a common ancestor in the absence of recombination.

Procrustes analysis

Also known as Procrustes superimposition. A statistical method allowing the translation, rotation and scaling of multidimensional objects within a single analytical space where they can be compared.

16S meta-barcodes

Selected variable regions of the 16S ribosomal RNA gene whose sequence provides taxonomic resolution amongst bacteria and archaea.

DNA methylation

A biological process by which the activity of a DNA segment is modified without changing the underlying sequence but by adding methyl groups to the DNA molecule.

Bisulfite conversion

A chemical reaction using sodium bisulfite that converts unmethylated CpG dinucleotides into UpGs but leaves methylated CpGs intact, thereby allowing the detection of DNA methylation by sequencing.

Immunoprecipitation

A molecular laboratory technique by which specific molecules are purified on the basis of their chemical affinities for particular protein groups, such as antibodies.

Population replacement

A population process by which the gene pool of one local population is at least partially replaced by that coming from another, genetically distinct, population.

Environmental DNA

[eDNA]. Fragments of DNA that are preserved within sediments and water that can be used for a fast, cost-effective monitoring of the ecology of a given region.

Stratigraphic leaching

The migration of DNA across strata in sediments caused by water movement, microorganism growth or bioturbation and compromising the reliability of the stratigraphy, that is, the order, position and age of the geological layers formed by the different piles of sediments.

Rights and permissions

Reprints and Permissions

About this article

Cite this article

Orlando, L., Allaby, R., Skoglund, P. et al. Ancient DNA analysis. Nat Rev Methods Primers 1, 14 [2021]. //doi.org/10.1038/s43586-020-00011-0

Is an act that takes advantage of a vulnerability to compromise a controlled?

A computer exploit, or exploit, is an attack on a computer system, especially one that takes advantage of a particular vulnerability the system offers to intruders. Used as a verb, exploit refers to the act of successfully making such an attack.

Is the process of using social skills to convince people to reveal access credentials?

Social Engineering: social engineering is the process of using social skills to convince people to reveal access credentials or other valuable information to the attacker.

What of information is the quality or state of being genuine or original?

Authenticity: Authenticity of information is the quality or state of being genuine or original, rather than a reproduction or fabrication. Information is authentic when it is in the same state in which it was created, placed, stored, or transferred.

What is Information Security's primary mission?

Three primary goals of information security are preventing the loss of availability, the loss of integrity, and the loss of confidentiality for systems and data. Most security practices and controls can be traced back to preventing losses in one or more of these areas.

Bài Viết Liên Quan

Chủ Đề