Next Article in Journal
Influence of Green Tea Added to Cherry Wine on Phenolic Content, Antioxidant Activity and Alpha-Glucosidase Inhibition during an In Vitro Gastrointestinal Digestion
Next Article in Special Issue
The Characterization of Dry Fermented Sausages under the “Chorizo Zamorano” Quality Label: The Application of an Alternative Statistical Approach
Previous Article in Journal
Metagenomics of Virus Diversities in Solid-State Brewing Process of Traditional Chinese Vinegar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Application of Metagenomics to Study Microbial Communities and Develop Desirable Traits in Fermented Foods

by
Meghana Srinivas
1,2,3,
Orla O’Sullivan
1,2,4,
Paul D. Cotter
1,2,4,
Douwe van Sinderen
2,3 and
John G. Kenny
1,2,4,*
1
Food Biosciences Department, Teagasc Food Research Centre, Moorepark, P61 C996 Cork, Ireland
2
APC Microbiome Ireland, University College Cork, T12 CY82 Cork, Ireland
3
School of Microbiology, University College Cork, T12 CY82 Cork, Ireland
4
VistaMilk SFI Research Centre, Fermoy, P61 C996 Cork, Ireland
*
Author to whom correspondence should be addressed.
Foods 2022, 11(20), 3297; https://doi.org/10.3390/foods11203297
Submission received: 4 September 2022 / Revised: 11 October 2022 / Accepted: 19 October 2022 / Published: 21 October 2022

Abstract

:
The microbial communities present within fermented foods are diverse and dynamic, producing a variety of metabolites responsible for the fermentation processes, imparting characteristic organoleptic qualities and health-promoting traits, and maintaining microbiological safety of fermented foods. In this context, it is crucial to study these microbial communities to characterise fermented foods and the production processes involved. High Throughput Sequencing (HTS)-based methods such as metagenomics enable microbial community studies through amplicon and shotgun sequencing approaches. As the field constantly develops, sequencing technologies are becoming more accessible, affordable and accurate with a further shift from short read to long read sequencing being observed. Metagenomics is enjoying wide-spread application in fermented food studies and in recent years is also being employed in concert with synthetic biology techniques to help tackle problems with the large amounts of waste generated in the food sector. This review presents an introduction to current sequencing technologies and the benefits of their application in fermented foods.

1. Introduction

The use of fermented foods has been recorded for thousands of years and continues to be of global importance to this day. Numerous fermented foods have existed across civilizations with the techniques used being indigenous to the resources available to a region. Traditionally, fermentations were carried out as a method of preservation to improve the microbiological safety while prolonging the shelf-life of food products, or for the exclusion of pathogens, when cold storage methods were not yet invented [1]. Over the years, however, fermented foods have been exploited for their health-promoting activities, and for their appeal to the consumer and industry, leading to their large-scale production [2,3,4]. The extent to which they have been commercialised depends on the region they are from, the techniques used, the availability of starter cultures used to start the fermentation process, along with resources available to research and industrialise the process [1]. A vast number of fermentation processes have been largely amenable to industrialisation where the starter cultures and production techniques have been well-characterised and fine-tuned over decades to produce consistent and high-quality products as in the case of the dairy, bread, meat and brewing industries. Others have remained very traditional, with recipes passed down from one generation to another in a household, or for small-scale production in local cottage industries. These methods depend largely on existing, yet undefined starter cultures that are added to start fermentations through a method known as back-slopping where an amount of a previous batch of fermented product is added to start a new fermentation [5]. However, since the starter cultures involved are still largely uncharacterised and little quality control can be performed while back-slopping, consistency and microbiological safety of the product remains to be a matter of concern, often making it inefficient for industrial fermentations [6]. Another food fermentation method known as spontaneous fermentation, uses naturally occurring microbes that are native to the raw food matrix and surrounding environment to carry out fermentations. Examples of spontaneous fermentations include the production of sour beers, some wines, and vegetable based-fermentations such as sauerkraut and kimchi [7,8,9,10].
Fermentation microbiomes are complex and dynamic with various microbes imparting characteristic flavours, odours, and texture throughout the fermentation process and into the finished product [11,12,13,14]. Profiling and characterisation of starter cultures and autochthonous fermentation microbes provides clarity in understanding underlying fermentation principles and allows optimisation of fermentation processes to improve product organoleptic and microbial safety qualities, and to ensure product consistency. This is where metagenomics plays a major role by allowing microbial characterisation and tracking, while providing insight into their interactions with other members of the fermented food microbial community.
The development of high-throughput sequencing (HTS) has allowed the application of metagenomics in numerous environmental and more recently fermentation microbiome studies [15]. With metagenomics, the entire DNA content of a microbial community can be studied at the same time, unlike culture-dependent methods where single colonies are isolated in order to sequence their whole genomes. Being a culture-independent technique, metagenomics is able to identify and characterise microbial species that are difficult to grow in a lab setting [16]. However, sequencing dead microbial cells confounds metagenomics data analysis. The inadequate detection of microbial populations present at low relative abundance is also problematic to the application of metagenomics to highly diverse and/or low microbial abundance samples. The potential solutions to these issues are described in the following sections.
Metagenomic sequencing can be broadly classified into two methods based on the DNA regions being sequenced, one being targeted or amplicon sequencing, also termed as metabarcoding or metataxonomics [6], where specific regions of the gDNA in a microbiome sample is targeted by PCR amplification and sequencing, and the other being untargeted or shotgun metagenomic sequencing where the entire genetic material in a microbiome sample is sequenced. The sequencing platform or method chosen depends largely on the type and number of samples, budget of the project, and computational resources available to process and analyse the sequencing data [17]. After sample collection and storage, a metagenomics experiment can broadly be broken down into four main steps; (i) extraction of microbial DNA from the sample; (ii) library preparation; (iii) DNA sequencing; (iv) bioinformatic processing and analysis of the generated sequence data [18].

2. Microbial DNA Extraction

The success of sequencing in terms of data quality and output is to a large degree dependent on the quality of DNA extracted from microbial communities. The DNA input requirements, such as concentration and fragment length, vary based on the sequencing method and platform used. Damaged and fragmented DNA can often lead to problems with library preparation, causing inefficient sequencing. DNA isolation can be especially difficult when collecting samples from equipment and food contact surfaces in food processing facilities due to the use of sanitising agents on these surfaces that injure micro-organisms and nick the DNA of the microbes present [19,20]. Unfermented and fermented foods are often rich in lipids and proteins, which can interfere with microbial DNA extractions, and must be removed prior to DNA extraction [21,22]. Pre-DNA extraction processing methods for traditional fermented foods have also been developed for highly viscous and sticky fermented foods rich in microbial polysaccharides and that are otherwise difficult to process [23,24,25]. After pre-processing samples, when required, metagenomic DNA extraction is performed. A number of non-commercially and commercially available DNA extraction kits exist, with each approach having its own advantages and disadvantages depending on the sample type and sequencing method employed [26,27]. Commercial DNA extraction kits can be expensive and may not be applicable to very traditional fermentation setups. However, they have the advantage of being standardised approaches [23].
Long read sequencing can be particularly sensitive to the quality of input DNA as highly fragmented DNA can only produce short reads, thereby failing to realise the advantages of the sequencing platform. Various commercially available DNA extraction kits are recommended by PacBio for long read sequencing [28], and have been used for long read 16S rRNA metabarcoding [29,30,31]. These methods involve mechanical bead beating steps that some consider detrimental to the success of long read sequencing while others consider this a requirement to provide good representations of highly diverse microbial communities [32,33,34,35]. However, mechanical steps involved in extraction procedures can be harsh on the DNA leading to fragmented DNA templates, which may underutilize the potential of long read sequencing platforms [36]. DNA extraction methods specifically suited to long read sequencing, called High Molecular Weight (HMW) DNA extractions, have therefore been developed to avoid bead-beating steps. MetaPolyzyme (a commercial product sold by Sigma Aldrich, Burghausen, Germany), is one such example, where an enzyme cocktail is used instead of mechanical steps to lyse microbial cells [37]. Kits for HMW DNA extractions for metagenomics are also commercially available with a few examples being the DNA extraction kits by Circulomics and “Fire Monkey” by RevoluGen.

3. Host Depletion

When assessing the quality of DNA extracted for metagenomic sequencing, contamination from non-microbial or host DNA, usually animal or human, should also be considered. More than 90% of the DNA fragments from samples such as blood, saliva and milk can come from the host genomes [38,39]. In shotgun metagenomics, since all the genetic material including host DNA is sequenced, a large amount of sequencing output is wasted on such contaminating host DNA [40,41]. This can lead to a high number of microbial species being unclassified, incorrectly classified and/or under-represented, thus causing serious inaccuracies in sample microbial community profiling [42,43]. This is especially problematic when applying shotgun sequencing to low microbial abundance samples such as saliva, skin and milk [44]. Therefore, host DNA depletion is often advantageous when preparing gDNA for shotgun sequencing. Host depletion in amplicon-based approaches is not required as the PCR step is selective and amplifies only target microbial DNA regions [41].
Host DNA depletion can be performed in two ways depending on whether they are carried out pre- or post-extraction. Pre-extraction methods use chaotropic agents to lyse mammalian host cells while allowing microbial cells to remain intact. The released host DNA is then degraded by nucleases such as DNase I or Benzonase. The latter is increasingly employed due to its wide range of operating conditions compared to DNase I. Once host DNA is degraded, microbial gDNA extraction is performed [41]. Commercial kits such as MolYsis (Molzym, Bremen, Germany) use this pre-extraction method with a proprietary DNase called MolDNase B, while the QIAmp DNA Microbiome kit (Qiagen, Hilden, Germany) performs host depletion using Benzonase [45]. In food metagenomics, the MolYsis kits were observed to be well suited for milk metagenome studies [21,44]. The Host ZERO microbial DNA kit (Zymo, Irvine, CA, USA) also uses the pre-extraction method with bead beating using two different bead sizes for host depletion [45]. Propidium Monoazide (PMA)-based methods for host depletion are also available that are performed prior to extraction [46]. The drawbacks observed for pre-extraction methods are: (i) the likely destruction and consequent under-representation of sensitive microbes such as Mycoplasma spp. and parasites during selective lysis, and (ii) for PMA-based methods, biasing towards Gram-positive bacteria due their increased susceptibility to PMA treatment compared to Gram-negative bacteria [34,41,47,48]. The second approach to host DNA depletion takes place post-DNA extraction, and uses differences in methylation characteristics between microbial and host or eukaryotic genomes. The NEBNext Microbiome DNA Enrichment kit (NEB, Northborough, MA, USA) uses magnetic beads to selectively bind and remove CpG methylated host DNA. However, with this post-extraction method, problems have been identified for AT rich genomes and differentiating between eukaryotic microbial and host DNA such as fungi, algae and protozoa that have similar methylation patterns [41]. In general, the method used for host depletion can vary between sample types with each having their own drawbacks and so should be decided accordingly [49]. Both pre- and post-extraction methods involve a number of washing and spinning steps that can reduce microbial (DNA) abundances in samples [45]. This is a major problem in low biomass samples, sometimes requiring PCR steps to obtain sufficient DNA concentrations for library preparation [50].

4. Differentiating between Live and Dead Bacteria

Differentiating between viable and non-viable microbes in a sample community can provide useful information and is performed through a process known as viability testing. In a microbial community, microbial populations can shift over time with various microbial species dominating and dying out. This shift in microbial populations can affect the type and quantity of metabolites produced which can affect neighbouring microbes and the surrounding environment [43,47,51]. While metagenomics provides information on the entire microbial community, by itself it cannot differentiate between live and dead bacteria. For a better understanding of microbial communities at particular time-points, additional methods to differentiate between live and dead bacteria are needed to be applied [52,53]. Propidium Monoazide (PMA) is the most commonly employed for viability testing. PMA is a dye that intercalates with DNA in the absence of a cell membrane. Upon exposure to visible light, PMA undergoes cleavage in its azide group with a C-H insertion reaction leading it to being covalently bound with the DNA. In this way, PMA acts only on free DNA released from dead and/or membrane-damaged microbial cells to prevent their further processing and sequencing [46]. The sequencing data obtained will therefore be representative of the viable microbial cells at a specific time-point. While PMA offers the benefits of viability testing, its activity has been assessed only on a small subset of biological matrices. A number of factors such as sample type, chemical composition, experimental conditions, duration of light exposure, and incubation time can influence PMA’s activity in degrading free DNA. Some cases have been reported where PMA partially or completely fails to remove free DNA, which can skew the results obtained, leading to under and/or mis-representations of the microbial community [54,55]. PMA penetration into dead cells also may be incomplete and may not be permitted in partially membrane-compromised bacterial cells, which can result in overestimations of live cells [56,57]. Therefore, the use of PMA in viability-based metagenomics needs to be further standardised. Live, metabolically active microbes in a sample can also be selected for and characterised using methods such as metatranscriptomics and metaproteomics where only mRNA or actively-expressed proteins, respectively, are sequenced [53,56,58]. Both metatranscriptomics and metaproteomics have been useful in understanding fermentation microbiomes and the interactions within its communities [59,60,61].

5. Sequencing Platforms

Sanger sequencing was among the first generation of sequencing technologies that largely contributed to the development of automated DNA sequencers [62]. Since then, major advances in sequencing technology has led to the rise of Next Generation Sequencers (NGS) that marked the start of many of the short read and metagenomic applications presently seen. Roche 454, Illumina, and Ion Torrent have been the forerunners of NGS with a vast majority of metagenomic projects employing the Illumina suite of sequencers [62,63,64,65]. Illumina platforms use sequencing by synthesis, which occurs on flow cells and uses fluorescently labelled nucleotides which are incorporated by DNA polymerases complementary to the template DNA strand. On incorporation, light of a specific wavelength is emitted and images are taken by a camera in the instrument. The images are then interpreted to DNA sequences one base at a time [65]. The high throughput, relatively low cost per base, and low error rates of 0.1–1% in Illumina sequencers is the reason behind the platform dominating the short read sequencing market [66]. As the demand for improved sequencing methods is increasing, the recent releases of Illumina aim at improving the throughput capacity and cost efficiency, while also reducing error rates. Among the latest Illumina releases is the NovaSeq 6000, which allows industrial-scale sequencing, generating up to 6 Tb and 20 billion reads with the lowest cost per base compared to earlier versions. Reduced error rates have also been recorded on NovaSeq 6000 and HighSeq X Ten, with the latter of the two sequencers being the most inexpensive human genome sequencer [67].
Over recent years there have also been ever greater developments relating to the third generation of sequencers, i.e., long read sequencers. PacBio and Oxford Nanopore Technologies (ONT) have dominated much of the long read sequencing market. The principle used in PacBio sequencing is that DNA fragments of approx. 250 to 25,000 bp are ligated with hairpin adapters forming a circular template, which when introduced to the Single Molecule Real Time (SMRT) cell, settle in the wells of the cell with one circular template taking up one well each. Within the wells, DNA polymerases add nucleotides complementary to the template DNA strand. This process can happen either multiple times in a mode known as Circular Consensus Sequencing (CCS) to generate HiFi data that is of high accuracy, or in a mode wherein longer DNA templates will be sequenced fewer times with more importance given to sequencing the entire length of the DNA fragment generating continuous long read (CLR) data [68]. A mix of the two methods, CCS and CLR, have been applied to sequence long eukaryotic genomes [69,70,71]. ONT uses protein pores, called nanopores, which are embedded into a membrane on a flow cell. During sequencing, an ion current is applied and single stranded DNA moves through the nanopores. As the DNA passes through a nanopore a characteristic disruption in ion current is identified by sensors and recorded. These recorded disruptions are then analysed to determine the corresponding nucleotide sequences. When HMW DNA extraction methods are followed, ONT platforms can even generate reads of 1 Mb in length or even longer [72].
PacBio and ONT have found application in both amplicon and shotgun metagenomics to varying extents. One of the major drawbacks in both the platforms was the historical high raw error rates of about 10–20% [73]. However, recently numerous studies have been dedicated to addressing this issue and has resulted in a number of bioinformatic tools and pipelines available for reducing and correcting error rates in long read platforms [73,74,75]. Significant improvements are also being made by PacBio and ONT with frequent releases of kit chemistries, sequencing instruments and flow cells allowing improved efficiency, accuracy, and data yield making it more amenable to wider application in metagenomic studies. The latest Sequel II and Sequel IIe platforms by PacBio along with the new 8M flow cells can provide accuracy of 99.8%, comparable to that of short read sequencing [68,72]. The recently released kit 12 chemistry and R10.4 cells by ONT supported by 1D2 technology allows consensus sequencing of complementary DNA strands and has an increased sequencing accuracy of more than 99% [76,77].
Apart from the sequencing platforms that currently dominate much of the market, newer competing platforms have recently been introduced that improve the scope of accessibility of sequencing technologies. Examples include Element Biosciences, MGI, and Omniome. All three target improvements in data accuracy and yield, alongside cost reduction, which will hopefully benefit customers/consumers due to increased competition in the short-read sequencer market.

6. Library Preparation and Multiplexing

Library preparation can be divided into the following steps: DNA processing to obtain PCR amplicons or fragments of desired sizes, multiplexing, and in most cases adapter ligation with the exception of amplicon sequencing on Illumina platforms.
For amplicon sequencing, amplicons are generated by targeting gDNA regions through PCR amplification. The amplicon size and PCR conditions depend on the sequencing platform and objective of the study. For shotgun sequencing, DNA fragments of desired size are obtained through a process known as fragmentation, which can be performed using sonication, acoustic cavitation, or enzymatically with DNA nucleases. DNA fragments of less than 450 bp are recommended for short read sequencing platforms, while fragment lengths up to 75 kb are often desirable for long read sequencing [78]. Often this means a need to isolate HMW DNA as described above. Sometimes it is still useful to fragment HMW to smaller fragments of ~20 kb to improve yields of sequencing, or to allow for multi-pass HiFi reads. In such cases, specific mechanical shearing devices such as the Megaruptor system are used as they improve consistency and reproducibility of the fragment lengths [79,80]. Post fragmentation, size selection for the desired fragment lengths is often performed.
Multiplexing, also called indexing or barcoding, is a method in which multiple libraries are pooled together so they can be sequenced on a single run and is used to reduce cost and save time when sequencing a large number of samples. Multiplexing uses specific and distinct nucleotide sequences, called index sequences or barcodes, which are added to the ends of amplicons or DNA fragments. After sequencing, barcoding allows assignment of the sequencing reads to the specific source sample from the pool of libraries [81].
Adapter ligation is the process in which platform-specific nucleotide sequences are added onto the ends of amplicons or DNA fragments, which allow the DNA regions of interest to bind or settle in the flow cells where sequencing occurs. For shotgun sequencing on Illumina, the adapters help bind the template DNA to the flow cell where sequencing cycles take place [82]. Amplicon sequencing on Illumina does not require adapter ligation because the adapter sequences can be incorporated during PCR. In PacBio, the hairpin adapters provide a circular shape to the long DNA fragments before the DNA polymerases initiate sequencing. In ONT, adapters are ligated to double stranded DNA and allows the strands to be captured by the nanopores on the flow cell. The ONT adapters also act as the starting point for a motor enzyme that runs along a DNA strand helping it pass through the nanopore [72]. The specific processes and order in which multiplexing and adapter ligation is carried out during library prep depends on the sequencing platform, kits used and the sequencing method. Figure 1 presents a general overview of their workflow.
While multiplexing is advantageous, there are a few challenges that are yet to be overcome in the technology. Misassignment of reads to indexes, and so their source libraries, is a common problem on various sequencing platforms leading to issues in downstream analysis [83]. It has been identified as a particular problem with Illumina sequencers using patterned flow cells due to the chemistries involved [84]. This problem of “index hopping” has been linked to the presence of free-floating indexing primers present in the pooled libraries introduced onto the flow cell [85,86]. Ineffective clean up and size selection steps, and improper storage of the prepared libraries leading to fragmentation of the template DNA, are sources of these free-floating indexing primers in the pooled libraries [86]. One solution to this issue is the use of unique dual indexing, where the indexing sequences added on either side of the amplicon or DNA fragment is unique to a single library. This means every library has two index sequences, one at each end of the DNA fragments that are unique to it. No index sequence will be shared between any two or more libraries of that pool [87]. However, the need for high numbers of validated indexes, and the associated costs with having so many indexes available can make unique dual indexing challenging when pooling a large number of samples. In these situations nested metabarcoding, where a combination of two indexing primer pairs are incorporated onto the ends of the template DNA through a nested PCR approach, can be used instead. This allows four distinct indexing primers to be incorporated within each library fragment to minimise the effects of index hopping [81,88]. Cross-talk between indexing primers can also occur by other means, including cross-contamination during the synthesis of primers or adapters, sample handling, the generation of chimeras during PCR steps, multiple misread of bases in the index sequences during sequencing, and carry-over of indexing primers or adapters from previous sequencing runs [83,87]. Many of these sources of error can be eliminated by following good laboratory and library prep practices [16]. However, index hopping continues to be an area of concern with newer sequencing companies such as MGI introducing methods claiming to have reduced index hopping on their platforms [89].
The reagents used for extraction and library preparation are another source of bias in metagenomic sequencing. Microorganisms have been found to grow in the buffers and reagents used in DNA extraction and library preparation, such as in the PCR reagents or water. This contaminating microbial DNA is sequenced along with the intended metagenomic samples, biasing the microbial community representations and causing inaccuracies in taxonomic classifications, microbial abundance and diversity calculations [90]. Shotgun metagenomics, especially for low biomass samples, are also very sensitive to this so-called “kitome” contamination. This makes the inclusion of experimental controls such as mock communities and negative control extractions of paramount importance to remove these sources of bias [16,91].

7. Sequencing Methods

7.1. Targeted or Amplicon-Based Sequencing

The DNA regions most often targeted in metabarcoding is the 16S rRNA gene in bacteria and the Internal Transcribed Spacer (ITS) region in fungal studies [92,93]. The 16S gene has been chosen for metabarcoding in bacterial genomes as it is largely conserved in almost all bacterial species allowing the use of universal primers, while hypervariable regions permit the identification and taxonomic classification of bacteria. The 16S rRNA gene plays a crucial role in protein synthesis initiation and mRNA translation and is present in every bacterial cell, making it a universal target [94]. Short read sequencing only allows some of the hypervariable regions (designated V1 through to V9) of the 16S rRNA gene to be sequenced. Generally amplicons of up to 450 bp to include regions such as V1–V3, or V3–V4 are targeted by PCR for sequencing. The appropriateness of the hypervariable regions depends on the nature of sample source. Debate remains in this area as hypervariable regions targeted between different studies and for specific bacterial genera differ [19,95,96,97]. Irrespective of the issue relating to the choice of hypervariable regions used, 16S rRNA sequencing has seen massive application in the metagenomics field, specifically for the V3–V4 region coupled with Illumina sequencing [98,99,100,101]. A majority of metabarcoding studies have employed the Illumina MiSeq or HiSeq 2500 platforms, the latter of which is no longer supported.
The relative ease with which bioinformatic processing and analysis of amplicon data can be performed compared to shotgun metagenomic data is another contributing factor to the widespread application of metabarcoding. The processing and analysis usually involves quality control steps of quality trimming, quality filtering and adapter removal from the reads, followed by taxonomic classification, which is usually performed using alignment methods against reference databases. For short reads, taxonomic classifications can be performed either through clustering sequences, often with 97% similarity, into Operational Taxonomic Units (OTUs), or by grouping of identical or exact matching sequences using Amplicon Sequencing Variants (ASVs). QIIME2 [102], mothur [103,104], MG-RAST [105], UPARSE [106], FROGS [107] are examples of OTU based pipelines while, Bioconductor [108], Deblur [109], and DADA2 [110] are examples of ASV-based pipelines. ASV-based methods have been found to provide better resolution than OTU-based methods [111]. A detailed discussion of 16S analysis pipelines is beyond the scope of this review, and for more information we refer to some excellent reviews [112,113].
Metabarcoding using long read sequencing has developed substantially over the recent years with improvements in base calling, reduced error rates, and fine tuning of bioinformatic pipelines [114]. Many fermented food studies have applied full length sequencing of the 16S gene (approx. 1500 bp in size) to determine microbial communities [19,29,30,31,115,116]. Compared to short read sequencing of one or two hypervariable regions, long read sequencing of the entire 16S gene does improve resolution of taxonomic assignments from genus level to species level. This avoids problems associated with the choice of which hypervariable regions to target, but strain level resolution still cannot be obtained. As a solution to this, attempts have been made to use long read amplicon sequencing to target the entire RRN operon (approx. 4300 bp in size) consisting of the 16S rRNA gene, ITS region and 23S rRNA gene [117]. Targeting the combined 16S-ITS-23S regions instead of individual rRNA locus-derived fragments as commonly done in short read metabarcoding, can provide information on 16S and 23S gene sequences from single reads allowing strain level resolution of microbial communities, and improve diversity, divergence and phylogenetic estimations [117,118,119,120,121]. Depending on the primers used, sequencing of the RRN operon also enables identification and classification of Archaea and Bacteria from the same libraries [122]. However, the recent nature of developments means there are new challenges within the field, and the methods are yet to be applied to fermented foods. One such challenge is that long-read sequencing has a higher raw error rate compared to short-read sequencing. Custom-made bioinformatic pipelines are being developed specifically to reduce error rates within RRN operon sequencing [122]. With long PCR products, chimerism can also be problematic for which Unique Molecular Identifiers (UMIs) have been identified that can be useful to generate highly accurate long amplicons [123]. Additionally, the unlinked arrangement of the 16S and 23S genes in the genomes of soil bacteria presents a challenge to the scope of RRN amplicon-based community profiling in environmental samples [124]. Metabarcoding, being a highly database-dependent approach, requires large and regularly maintained databases to accurately perform taxonomical classifications [125]. Therefore, with RRN amplicon sequencing providing improved resolution and taxonomic characterisation of microbial communities, the presence of an RRN database is crucial. Taxonomic classification using RRN long reads have been performed majorly using the rrn database that searches bacterial strains based on the 16S, 23S, 5S, ITS and tRNA copy numbers, or through modified pipelines of existing 16S databases such as NCBI and SILVA to suit RRN application [120,122,126]. A commercially available RRN database named Athena along with the bioinformatic pipelines required to process and analyse long RRN amplicon reads has recently been added to the market by Shoreline in collaboration with PacBio, access to which can be obtained on purchasing their DNA extraction and library preparation kits [127]. To our knowledge, to date only one freely available reference database, named MIrROR, currently exists for RRN operon-based profiling applications [128]. The bioinformatic methods used to process and analyse long amplicon sequencing data also differ from those used for short 16S reads. OTU and ASV-based methods can be inconsistent for long reads leading to uncertainty in microbial classification and abundance calculations [129]. Presently minimap2 and BLAST, a very early aligner, are the most commonly used alignment tools to perform taxonomic assignment of long amplicon data [130,131]. While more tools are being developed, many are yet to be benchmarked preventing long amplicon sequencing from realising its full potential. Wider adoption of long amplicon sequencing will lead to its development and standardisation.

7.2. Untargeted or Shotgun Metagenomic Sequencing

Unlike metabarcoding methods, shotgun metagenomics approaches provide sequence data on all of the DNA content of a given sample allowing a number of genes and genome characteristics to be identified that can otherwise be complex to profile using amplicon-based methods. While tools such as PICURSt2 [132] and Tax4Fun [133] exist to functionally profile microbes using 16S sequencing data, it can be difficult to obtain strain level resolution and account for mobile genetic elements such as Horizontal Gene Transfers (HGT) using these tools [97]. Therefore, functional profiles obtained from shotgun metagenomics are superior to metabarcoding and can be useful in identifying secondary metabolites, bacteriocin gene clusters, complex metabolic pathways and interactions between pathways in microbial communities [19]. While the large amounts of sequencing data generated by shotgun metagenomics is beneficial as mentioned above, it is also more complex to process and analyse making the method computationally heavy and expensive [134]. The advantages and disadvantages of shotgun metagenomics when compared to metabarcoding are highlighted in Table 1.
Following sequencing, the raw data generated from shotgun sequencing is first passed through quality control steps. Tools such as TrimGalore, KneadData and Bowtie 2 are commonly used for adapter removal, quality trimming and host DNA removal for shotgun data generated on an Illumina platform [16,135]. Taxonomic and functional profiling can then be carried out in two ways on shotgun data, one through direct or assembly-free methods, such as Kaiju [136], Kraken [137,138] and Metaphlan [139] that assign reads using either amino acid sequence similarity, lowest common ancestor (LCA) along with k-mer matching, or clade specific markers, respectively [16,17]. Each pipeline used for assembly-free analysis has its own advantages and disadvantages with variations in results obtained based on the type of classifier and data used [21,140,141]. The pipeline chosen depends on the computation resources available, ease of use, along with the specific requirements of each pipeline [141]. Assembly-free methods work well if reference databases are constantly added to and maintained to include a diverse range of high-quality microbial genomes from across multiple sample types. However, as databases expand with a high number of metagenomic studies being conducted currently, assembly-free methods will need to be redesigned to enable their application with such large datasets [141,142].
Another method in shotgun sequencing is the assembly of reads to generate individual genomes of various microbial species/strains originating from metagenomic samples, called Metagenome Assembled Genomes (MAGs). MAGs can provide better microbiome resolution and can improve microbial characterisation and identification at species and/or strain level. MAG assembly for short reads uses overlapping reads to form contigs which are then sown together to form assemblies. While MAGs can be extremely informative about microbial populations, difficulties are still faced during the process of assembly [143]. Differing abundances of strains results in different levels (also known as depth, or coverage), of sequencing for the various genomes in a community. This variation in coverage, as well as variations in GC content are challenges to perform accurate genome assembly [144]. One method of improving MAGs generated from short reads is the process of binning, wherein similar reads are grouped together into bins before assembly. It can be carried out in two ways, supervised, where the reads are aligned against reference genomes, or unsupervised where genome characteristics such as k-mers can be used to construct assemblies which is especially useful in de novo assembly and characterisation [144]. metaSPAdes [145], Meta-IDBA [146], MetaBAT [147], CONCOCT [148], MEGAHIT [149], and MaxBin [150] are commonly employed assembly software programmes [16]. Most tools currently take GC content and coverage into account while binning. However, repetitive and mobile genetic elements continue to be problematic to MAG generation even when binning techniques are employed [151,152].
Long read sequencing helps to overcome problems associated with repetitive genome elements by producing reads that are long enough to span these sequences. High quality MAGs generated from long read metagenomic data can provide improved microbial community resolution down to the strain level and allow identification and taxonomic characterisation of rare microbial strains [153,154]. Long read shotgun metagenomic methods and bioinformatics pipelines are still being developed with frequent testing against mock communities, to reduce error rates, generate better quality MAGs, and improve the overall accuracy of the method [155,156]. The constantly improving nature of library preparation methods and the sequencing chemistries means that computational “gold standards” remain to be established. The steps involved in long read bioinformatic pipelines usually include additional error rate reduction and polishing steps besides the usual quality control and classification steps. Long read pipelines are therefore complex, using a combination of tools which are beyond the scope of this review, but more detailed information is available in the following reviews [73,74,130,157,158]. The potential of long read sequencing is expected to see extensive growth in the near future as technological developments continue.
Table 1. Metabarcoding vs. shotgun metagenomics advantages and disadvantages.
Table 1. Metabarcoding vs. shotgun metagenomics advantages and disadvantages.
FactorsAmplicon SequencingShotgun SequencingReferences
Cost and speed of analysisAdvantages:
(1) Requires less sequencing per sample
(2) Faster and financially feasible when many samples are to be analysed or when only taxonomic profiling is required
(3) Bioinformatic analysis is relatively easier with many GUI-based software freely available, thereby reducing computational costs
Disadvantages:
Less data/information obtained on microbial communities
Advantages:
Untargeted sequencing of metagenomic samples generates large amounts of data useful for functional profiling
Disadvantages:
Analysis methods involved can be time consuming and computationally heavy often requiring complex and expensive network infrastructures
[133,143]
Library prepAdvantages:
(1) PCR-involving library preparation steps can increase template DNA numbers for low microbial populations, thereby improving their representation in the sequencing data generated
(2) Improves microbial sequencing from host-derived samples
Disadvantages:
(1) PCR related biases apply such as differences in:
(i) ease or rate of amplification
(ii) variation in GC content
(iii) copy number of 16S gene
(iv) sequence variation between 16S copies within a bacterial genome
(v) selection of targeted region
(2) More susceptible to biasing microbial community representations in the presence of contaminating microbial strains such as those introduced into libraries from kit reagents used
Advantages:
(1) PCR related biases also apply, but can be reduced using PCR-free library prep methods
(2) Less susceptible to biasing microbial community representations in the presence of kitome contaminants
Disadvantages:
Host-derived samples need to be depleted for host DNA before sequencing, if not sequencing resources will be wasted on sequencing large proportions of host DNA and can lead to under/mis-representations of microbial communities
[16,72,82,118]
Microbial community profilingAdvantages:
(1) Taxonomic classification possible for which computational processing and analysis is relatively simple and quick
(2) For functional classification tools such as PICURSt2 and Tax4Fun exist that functionally assign species detected in a community through metabarcoding to predict microbial functional abilities
Disadvantages:
Functional profiles can only be predicted from amplicon data but is difficult for highly diverse and complex samples. The resulting profiles are often of low resolution and do not account for mobile genetic elements such as Horizontal Gene Transfers (HGT) and pathogenicity islands
Advantages:
(1) The large amounts of sequencing data generated through shotgun metagenomics allows better functional profiling than metabarcoding
(2) Better resolution of microbial community, even at strain level, can be obtained
Disadvantages:
(1) The extent and quality of the functional profiles obtained depend on the complexity of the sample community and the sequencing depth
(2) Computational analysis is time consuming and requires complex network infrastructure to be set up and maintained which is expensive
[19,97,133,159,160]
Detection and classification of previously unidentified or uncharacterised genomes in a communityDisadvantages:
Dependent on existing databases, making classification of new species and strains difficult
Advantages:
Performance of de novo assembly allows characterisation of new species and strains and their addition to databases
Disadvantages:
MAG assembly for new species and strains can be very difficult for low abundance microbial populations and highly diverse microbial communities
[125,144]
Fungal or viral profiling Advantages:
(1) ITS-based fungal metabarcoding is relatively well characterized
(2) PCR-based library prep can improve sequencing of low abundance viral microbial community members
Disadvantages:
(1) Requires different primers for fungal and viral community members and cannot be identified from a single library
(2) PCR-based approaches for viral sequencing is restricted to similar or closely related viral families and can fail to detect new viral families
Advantages:
Bacterial, fungal and viral sequences can be identified from a single library
Disadvantages:
(1) Fungal sub-populations or secondary symbionts are difficult to sequence
(2) Only DNA-encoded viruses can be identified
[161,162,163,164]
Extra-chromosomal DNA profilingDisadvantages:
Plasmidome study is not possible
Advantages:
Plasmidomes can be characterised along with gDNA
Disadvantages:
It is difficult to extract plasmid and genomic DNA together, and to computationally process and assemble reads. However, Hi-C approaches developed are allowing the linkage of plasmids to their carriage strains
[165,166,167]

8. New Technologies

Despite the advantages, barriers to long read sequencing still exist causing short read platforms to have a continued dominance of much of the metagenomics sequencing market. This has led to the rise of technologies such as synthetic long reads and Hi-C that use alternative library preparation methods and short read sequencers as alternatives to long read sequencing.

8.1. Synthetic Long Read (SLR) Sequencing

This method uses synthetic, artificial or virtual long reads generated from short read data. Loop Genomics, TELL Seq, and Illumina TruSeq Synthetic Long-Read are major contributors to the field of SLR sequencing. The three technologies use barcoding of short read sequences, during library prep, which can be virtually linked post sequencing to generate long reads [154,168]. Illumina’s latest SLR technology, Infinity, which is still in its developmental stage claims to generate 10 kb contiguous reads with reduced input requirements compared to long read sequencing platforms. Longas is another contributor to the SLR field, which uses MorphoSeq technology, wherein uniform random mutagenesis is performed. Tracking of these mutations allows linkage of the short reads informatically to generate long reads. SLR sequencing leverages the cost, quality, and accessibility benefits of short read sequencing while improving genome assembly and gap finishing abilities. This further contributes to the increase in the number of finished genomes added to public databases [169,170]. SLR has also found application in amplicon sequencing to improve microbial resolution [168].

8.2. Hi-C

Another approach to improve genome assembly is using Hi-C. This method takes advantage of linking co-located DNA during library preparation. It was originally used to improve genome assembly for larger genomes, but more recently has been applied to metagenomics [171,172,173,174]. During library preparation of metagenomic samples, DNA within the bacterial cell is cross-linked by binding to surrounding proteins, following which it is cut using restriction enzymes, and ligated. This allows DNA fragments from within the same cell to stick together [175]. After sequencing, the reads are then informatically assigned to the same cell, helping improve MAG generation, and linking of plasmid and phage DNA to specific host strains. Commercial options for kits and analysis pipelines are available, with Phase Genomics being a major contributor to the field.

9. Applications of Metagenomics in the Fermented Food Industry

As sequencing technologies are becoming more reliable, accessible, with higher throughputs and reduced costs, many food companies and regulatory bodies have moved away from culture-based and classical sequencing methods such as single nucleotide polymorphism (SNP) and multilocus sequence typing (MLST), and have generally adopted NGS alternatives [176]. The rapid analysis speeds further supported by real-time base calling and identification of microbial species, offered by third generation sequencing technologies such as ONT, allow food industries and regulatory bodies to make quick, informed decisions that are crucial to preventing and/or limiting foodborne outbreaks and bacteriophage invasions within the processing facilities [177,178,179]. Recently developed technologies such as “Read Until” in ONT platforms allow selective sequencing through the classification of the short prefix sequence of a DNA or RNA strand entering a nanopore into a target or non-target sequence. If classified as belonging to a set of target sequences, the entire strand is then base-called and analysed, and if not, the non-target strand is then rejected from the nanopore making it available to other strands [50,180,181]. This technology can further improve analysis speed while extending flow cell life-span and reducing sequencing costs. The “Read Until” technology has potential application in the fermented food industry specifically in screening for industrial and health-related harmful and beneficial traits.
A number of metagenomic studies have linked the presence of various genes and the metabolic pathways involved to harmful or beneficial traits possessed by microbial populations. Antibiotic resistance genes (ARGs), are examples of harmful trait-associated genes, which have has been flagged by the European Food Safety Authority (EFSA) as being linked to harms or concerns associated with foods [182,183,184,185,186,187]. Specific databases, such as CARD [188] and ResFinder [189], are available that screen for ARGs in sequencing data. Genes associated with flavor development and health promotion are examples of beneficial trait-associated genes. Various metabolites produce characteristic flavours and/or textures, the composition of which largely depends on the microbial community, the succession patterns and interactions within the community. Genes associated with acid and ethanol production, amino acid and sugar metabolism, lipid and protein lysis are often screened for when studying flavour development during the different stages of fermentations [3,13,183,190,191,192]. The identification of sugar, specifically lactose, metabolism-associated genes can further aid in determining the health promoting traits of a fermented food as the microbial breakdown of lactose to lactate during yoghurt fermentation helps alleviate problems linked with lactose consumption in lactose intolerant individuals [193]. Fermented food microbial communities are also suggested to promote health through immuno-modulation, improving gut barrier functions, preventing pathogen colonization of the gut, neutralizing microbial toxins, and producing antimicrobials such as bacteriocins [3,183,187,191,194,195]. The genes associated with these health promoting functions are commonly screened for when understanding the health benefits of consuming fermented foods. Genes associated with prebiotic functions, linked to the breakdown of complex nutrients to reduce inflammation and irritation in the gut, along with producing health promoting metabolites such as short chain fatty acids (SCFAs) have also been identified [196,197,198,199]. The health promoting abilities of fermentation microbes have been associated with survival in the gut. Genes associated with these strains include exopolysaccharide production (EPS), urease, bile salt hydrolase and mucin-binding protein synthesis [192,194,200]. The successful linking of specific trait-associated genes with certain harmful and beneficial properties in metagenomic projects is supported by the accurate collection of metadata, such as sample collection location, host health, fermentation conditions, and fermentation batches, which allows researchers to better characterise microbial communities and their associations with various sample types [201]. The applications of metagenomics are further expanded by its combination with other methods such as viability-based approaches mentioned above and with other meta-omics methods such as metatranscriptomics or metabolomics to characterise only viable microbes that are actively producing metabolites of interest [187,191,202].
NGS in combination with metagenomics allows the benefits of rapidly developing sequencing technologies to be applied to microbial population studies. Metagenomic NGS is valuable to the study of fermented foods because the microbiomes involved, either in the form of starter cultures consisting of a few selected strains, or as a large microbiome native to the raw materials used for example in spontaneous fermentations, are spatiotemporally dynamic within the food matrix. The strains are often involved in complex interactions such as cross-feeding of metabolites produced by one species to another, and/or in competitive or co-operative relationships with one other [203,204,205,206]. These interactions are often the cause of desirable organoleptic or health-promoting traits being imparted to the fermented food. Without these complex interactions the same desirable metabolites might not be produced leading to inconsistencies, as well as reduced organoleptic characteristics and microbial safety of the final fermentation end product. For this reason, entire microbiomes involved in fermentations need to be studied together and not as individual isolates, unlike in earlier single isolate WGS methods, wherein certain key pathways may not be expressed without the influence of neighbouring microbial community members and surrounding food matrix conditions [207]. The high throughput abilities and technological advances of NGS have made metagenomics feasible and affordable allowing its application in studying the influence of a variety of factors, such as geographic location and food facility conditions, on the fermentation microbiome and the effects they have on the fermentation process and the end products. Applying metagenomics in this manner contributes to stream-lining food processing pipelines, ensuring consistency and microbial safety, while protecting food and microbe-associated IP rights, preventing food fraud and unauthorized use of microbial strains. NGS has found widespread application in the food sector with rapid developments seen in the field and an extensive array of publications within the area, a few examples of which are listed in Table 2.
Metagenomics has shed light on the viromes present in fermented foods whereas culture-based methods allow the study of only singular phages causing fermentation flaws or singular foodborne viruses at a time [163]. Virome studies are scarce in fermented foods but should not be neglected. Fermented foods can contain numerous phages that can have a substantial effect on the fermentation process and can lead to low quality or failed fermentations and fermented end products. Similarly, virome studies have significant potential in improving fermented food safety through the detection of foodborne viruses [208,209]. However, the sequencing of viruses in fermented foods can be problematic due to their low abundance and smaller genome size compared to bacterial and fungal populations present in the food. This is especially true for foodborne viruses that do not multiply in food substrates [163]. Virus genomes can be DNA or RNA-encoded and only small percentages of viromes have been taxonomically assigned [210,211]. The direct sequencing of RNA, without first converting to cDNA, is a developing field with a few platforms, such as ONT and TERA-Seq, introducing native RNA sequencing [212,213]. However, library preparation involving RNA to cDNA conversion coupled with targeted amplification can improve representation of low abundance viral RNA [163].
The method and/or platform selected to sequence metagenomic samples plays an important role in determining the type and quality of sequencing data obtained. Sequencing platforms are selected based on the objective of a study, no. of samples, funding, and computational infrastructure available. When monitoring food safety in terms of screening food for pathogens, a large number of samples may be involved especially at the scale of food industries [214,215]. The rapid analysis timelines offered by HTS compared to culture-based methods promotes the application of metagenomics in food quality control enabling quick and informed decisions on product recall. The low sequencing and computational costs of targeted amplicon sequencing compared to shotgun-based approaches makes it the more cost effective choice when sequencing a large no. of samples [216,217]. The free availability of many Graphic User Interface (GUI)-based computational resources in targeted amplicon sequencing further reduces computational costs, circumvents the need for specialist intervention, and makes the analysis process more open to standardisation [112,218]. Where more in-depth microbiome studies are required, such as screening for bacteriocin genes, antimicrobial resistance genes (ARGs), and functionally characterising microbial communities for health promoting or organoleptic qualities, amplicon sequencing cannot provide sufficient information. Shotgun sequencing is required for these objectives [216,217,219]. Although shotgun sequencing is more expensive, there is a trade-off between the cost and information obtained [220,221]. The sequencing approach used to study fermented food authenticity and the influence of various factors on the fermentation microbiome can depend on the objective of the study and the amount of information required.
While metagenomics is proving to be beneficial, the technologies involved may not be presently accessible or affordable to every fermentation process. However, with the market for sequencing technologies expanding and sequencing costs reducing, along with workshops on metagenomics being organised in rural, developing areas, the scope for metagenomic applications in traditional fermented foods is steadily increasing [23].
Table 2. Applications of metagenomics and NGS in fermented foods.
Table 2. Applications of metagenomics and NGS in fermented foods.
AreaApplicationReferences
Health promotionScreening for health promoting bacteria
Understanding the gut-brain axis
Identifying prebiotics and their effect on host gut microbiota and health
[24,187,195]
[222,223]
[224,225,226,227]
Characterising fermentationsOrganoleptic quality assessment through fermentation microbiome and volatile profiling
Bacteriophage:
(1)
Detection and characterisation
(2)
Screening for phage resistant microbial strains
[6,162,228,229,230,231,232]
[233,234,235]
[5,236]
Food safetyDetection and prediction of foodborne pathogens and spoilage microbes
Screening for bacteriocin gene clusters
Checking for the presence of antibiotic resistance genes (ARGs)
[176,237,238]
[239,240,241]
[185,242,243]
Food fraudFingerprinting plant, animal and microbial components of food, determining food authenticity, and detection of contaminants and adulterants[244,245,246,247]
Production analysisAccessing the effect of the following factors on fermentations:
(1)
Raw materials and fermentation facility conditions
(2)
Variation between batches
(3)
Geographical location
[101,248,249,250,251,252]
[230]
[204,230,253]

10. Synthetic Biology

Metagenomics and metaproteomics together have improved the scientific community’s understanding of microbial species and aided in comprehending the vast varieties of metabolic functions they can perform. A large number of proteins, genes and metabolic pathways that were previously unidentified and/or unclassified are now (being) characterised. Constant development in the field of biotechnology, and more recently synthetic biology, has allowed genetic manipulations of microbial species at large-scale to produce desirable end products such as fuels, enzymes, growth hormones, insulin, and monoclonal antibodies [254]. The addition of CRISPR-cas9 methods to microbial genome editing options when compared to more traditional promoter and terminator, or plasmid-based genetic manipulations improves the robustness and scalability of synthetic biology [254,255]. The relative ease with which microbial cells can be handled, propagated and cultured, and the whole production process scaled up further contributes to microbes and/or microbial-derived products to be successfully applied to solve problems currently marring the food sector [256,257].

11. Food Waste Valorisation

A significant area of concern in the food sector is food waste. About 1.3 billion tonnes of food waste is generated along the food supply chain from farms to final consumption [258]. A substantial portion of this waste is produced by food processing facilities [259]. The waste generated is often rich in lipids, proteins and carbohydrates, the direct disposal of which can be harmful to the environment [260,261,262,263]. Many current food production methods are not sustainable and are proving to be detrimental to the environment. In order to meet the growing demand for food, current farming, agriculture and industrial food processing strategies need to be re-evaluated [264,265]. Metagenomics has the potential to help resolve these difficulties. Farm hygiene conditions, animal health and soil fertility are important factors that contribute to food safety and quality and can be linked to the microbial communities present in these environments. Metagenomics has allowed the study of these microbial communities enabling researchers to identify solutions to improving food production techniques and possibly predict and control food loses caused due to disease conditions or unnatural-disease states linked to microbial communities [266,267]. In this way metagenomics can help to prevent and reduce food waste at the farm level. The food waste streams produced by processing facilities is another point where current molecular techniques can reduce food waste [268]. Food waste streams can be used as media to culture useful microbial strains to produce value-added compounds. For this, the technologies of metagenomics, synthetic biology and microbial biotransformation can be employed. Metagenomics allows researchers to first identify microbial genes linked to the production of useful enzymes or value-added compounds [269,270,271,272]. Synthetic biology techniques would then enable the commercial application of these pathways by improving efficiency and allowing upscaling [269,270,271,272]. This way, food waste streams can be microbially-biotransformed to value-added products, paving the way for the development of circular bioeconomies (Figure 2) [273,274,275].

12. Future of Molecular Biology in Fermented Foods

The increased commercial interest in sequencing is leading to rapid developments within the metagenomics field. These include the development of existing and new sequencing platforms such as Element Biosciences, Singular Genomics, Omniome, Genapsys, and Ultima Genomics. These platforms can be coupled with major advancements in accompanying technologies such as library reagents, spatial profiling, single cell-technologies, and analysis pipelines. Past performance indicates that improving the efficacy and reducing the financial burden of sequencing will continue to make the technology increasingly accessible to routine applications in the food sector, leading to more widespread adoption.

Author Contributions

Conceptualization, M.S. and J.G.K.; writing—original draft preparation, M.S.; writing—review and editing, O.O., P.D.C., D.v.S. and J.G.K.; supervision, O.O., P.D.C., D.v.S. and J.G.K.; project administration, J.G.K.; funding acquisition, J.G.K. and D.v.S. All authors have read and agreed to the published version of the manuscript.

Funding

M.S. is funded by the Teagasc Walsh Scholar Scheme (ref no. 2020018). This publication also received financial support of Science Foundation Ireland under grant number 12/RC/2273. D.v.S. and P.D.C. are members of The Alimentary Pharmabiotic Centre, which is a Centre for Science and Technology (CSET) funded by the Science Foundation Ireland (SFI), through the Irish Government’s National Development Plan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tamang, J.P.; Cotter, P.D.; Endo, A.; Han, N.S.; Kort, R.; Liu, S.Q.; Mayo, B.; Westerik, N.; Hutkins, R. Fermented Foods in a Global Age: East Meets West. Compr. Rev. Food Sci. Food Saf. 2020, 19, 184–217. [Google Scholar] [CrossRef] [Green Version]
  2. Morais, L.H.; Schreiber, H.L.; Mazmanian, S.K. The Gut Microbiota–Brain Axis in Behaviour and Brain Disorders. Nat. Rev. Microbiol. 2021, 19, 241–255. [Google Scholar] [CrossRef] [PubMed]
  3. Obafemi, Y.D.; Oranusi, S.U.; Ajanaku, K.O.; Akinduti, P.A.; Leech, J.; Cotter, P.D. African Fermented Foods: Overview, Emerging Benefits, and Novel Approaches to Microbiome Profiling. Npj Sci. Food 2022, 6, 15. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, M.; Zhang, D.; Su, X.; Duan, S.; Wan, J.; Yuan, W.; Liu, B.; Ma, Y.; Pan, Y. An Integrated Metagenomics/Metaproteomics Investigation of the Microbial Communities and Enzymes in Solid-State Fermentation of Pu-Erh Tea. Sci. Rep. 2015, 5, 10117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Somerville, V.; Berthoud, H.; Schmidt, R.S.; Bachmann, H.-P.; Meng, Y.H.; Fuchsmann, P.; von Ah, U.; Engel, P. Functional Strain Redundancy and Persistent Phage Infection in Swiss Hard Cheese Starter Cultures. ISME J. 2022, 16, 388–399. [Google Scholar] [CrossRef]
  6. Ferrocino, I.; Rantsiou, K.; Cocolin, L. Investigating Dairy Microbiome: An Opportunity to Ensure Quality, Safety and Typicity. Curr. Opin. Biotechnol. 2022, 73, 164–170. [Google Scholar] [CrossRef] [PubMed]
  7. Dysvik, A.; La Rosa, S.L.; De Rouck, G.; Rukke, E.-O.; Westereng, B.; Wicklund, T. Microbial Dynamics in Traditional and Modern Sour Beer Production. Appl. Environ. Microbiol. 2020, 86, e00566-20. [Google Scholar] [CrossRef] [PubMed]
  8. Mateus, D.; Sousa, S.; Coimbra, C.; Rogerson, S.F.; Simões, J. Identification and Characterization of Non-Saccharomyces Species Isolated from Port Wine Spontaneous Fermentations. Foods 2020, 9, 120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Song, H.S.; Lee, S.H.; Ahn, S.W.; Kim, J.Y.; Rhee, J.-K.; Roh, S.W. Effects of the Main Ingredients of the Fermented Food, Kimchi, on Bacterial Composition and Metabolite Profile. Food Res. Int. 2021, 149, 110668. [Google Scholar] [CrossRef] [PubMed]
  10. Tlais, A.Z.A.; Lemos Junior, W.J.F.; Filannino, P.; Campanaro, S.; Gobbetti, M.; Di Cagno, R. How Microbiome Composition Correlates with Biochemical Changes during Sauerkraut Fermentation: A Focus on Neglected Bacterial Players and Functionalities. Microbiol. Spectr. 2022, 10, e00168-22. [Google Scholar] [CrossRef] [PubMed]
  11. Das, S.; Tamang, J.P. Changes in Microbial Communities and Their Predictive Functionalities during Fermentation of Toddy, an Alcoholic Beverage of India. Microbiol. Res. 2021, 248, 126769. [Google Scholar] [CrossRef] [PubMed]
  12. Ashaolu, T.J.; Khalifa, I.; Mesak, M.A.; Lorenzo, J.M.; Farag, M.A. A Comprehensive Review of the Role of Microorganisms on Texture Change, Flavor and Biogenic Amines Formation in Fermented Meat with Their Action Mechanisms and Safety. Crit Rev. Food Sci. Nutr. 2021, 1–18. [Google Scholar] [CrossRef]
  13. Hu, Y.; Zhang, L.; Wen, R.; Chen, Q.; Kong, B. Role of Lactic Acid Bacteria in Flavor Development in Traditional Chinese Fermented Foods: A Review. Crit. Rev. Food Sci. Nutr. 2022, 62, 2741–2755. [Google Scholar] [CrossRef] [PubMed]
  14. Jiang, N.; Wu, R.; Wu, C.; Wang, R.; Wu, J.; Shi, H. Multi-Omics Approaches to Elucidate the Role of Interactions between Microbial Communities in Cheese Flavor and Quality. Food Rev. Int. 2022, 1–13. [Google Scholar] [CrossRef]
  15. Techtmann, S.M.; Hazen, T.C. Metagenomic Applications in Environmental Monitoring and Bioremediation. J. Ind. Microbiol. Biotechnol. 2016, 43, 1345–1354. [Google Scholar] [CrossRef] [Green Version]
  16. Quince, C.; Walker, A.W.; Simpson, J.T.; Loman, N.J.; Segata, N. Shotgun Metagenomics, from Sampling to Analysis. Nat. Biotechnol. 2017, 35, 833–844. [Google Scholar] [CrossRef] [Green Version]
  17. Breitwieser, F.P.; Lu, J.; Salzberg, S.L. A Review of Methods and Databases for Metagenomic Classification and Assembly. Brief Bioinform. 2019, 20, 1125–1136. [Google Scholar] [CrossRef] [PubMed]
  18. Prayogo, F.A.; Budiharjo, A.; Kusumaningrum, H.P.; Wijanarka, W.; Suprihadi, A.; Nurhayati, N. Metagenomic Applications in Exploration and Development of Novel Enzymes from Nature: A Review. J. Genet. Eng. Biotechnol. 2020, 18, 39. [Google Scholar] [CrossRef] [PubMed]
  19. Cao, Y.; Fanning, S.; Proos, S.; Jordan, K.; Srikumar, S. A Review on the Applications of Next Generation Sequencing Technologies as Applied to Food-Related Microbiome Studies. Front. Microbiol. 2017, 8, 1829. [Google Scholar]
  20. De Filippis, F.; Valentino, V.; Alvarez-Ordóñez, A.; Cotter, P.D.; Ercolini, D. Environmental Microbiome Mapping as a Strategy to Improve Quality and Safety in the Food Industry. Curr. Opin. Food Sci. 2021, 38, 168–176. [Google Scholar] [CrossRef]
  21. Yap, M.; Feehily, C.; Walsh, C.J.; Fenelon, M.; Murphy, E.F.; McAuliffe, F.M.; van Sinderen, D.; O’Toole, P.W.; O’Sullivan, O.; Cotter, P.D. Evaluation of Methods for the Reduction of Contaminating Host Reads When Performing Shotgun Metagenomic Sequencing of the Milk Microbiome. Sci. Rep. 2020, 10, 21665. [Google Scholar] [CrossRef] [PubMed]
  22. Escobar-Zepeda, A.; Sanchez-Flores, A.; Quirasco Baruch, M. Metagenomic Analysis of a Mexican Ripened Cheese Reveals a Unique Complex Microbiota. Food Microbiol. 2016, 57, 116–127. [Google Scholar] [CrossRef]
  23. Diaz, M.; Kellingray, L.; Akinyemi, N.; Adefiranye, O.O.; Olaonipekun, A.B.; Bayili, G.R.; Ibezim, J.; du Plessis, A.S.; Houngbédji, M.; Kamya, D.; et al. Comparison of the Microbial Composition of African Fermented Foods Using Amplicon Sequencing. Sci. Rep. 2019, 9, 13863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Turpin, W.; Humblot, C.; Guyot, J.-P. Genetic Screening of Functional Properties of Lactic Acid Bacteria in a Fermented Pearl Millet Slurry and in the Metagenome of Fermented Starchy Foods. Appl. Environ. Microbiol. 2011, 77, 8722–8734. [Google Scholar] [CrossRef] [Green Version]
  25. Zhao, N.; Cai, J.; Zhang, C.; Guo, Z.; Lu, W.; Yang, B.; Tian, F.-W.; Liu, X.-M.; Zhang, H.; Chen, W. Suitability of Various DNA Extraction Methods for a Traditional Chinese Paocai System. Bioengineered 2017, 8, 642–650. [Google Scholar] [CrossRef] [Green Version]
  26. Keisam, S.; Romi, W.; Ahmed, G.; Jeyaram, K. Quantifying the Biases in Metagenome Mining for Realistic Assessment of Microbial Ecology of Naturally Fermented Foods. Sci. Rep. 2016, 6, 34155. [Google Scholar] [CrossRef] [Green Version]
  27. Shaffer, J.P.; Carpenter, C.S.; Martino, C.; Salido, R.A.; Minich, J.J.; Bryant, M.; Sanders, K.; Schwartz, T.; Humphrey, G.; Swafford, A.D.; et al. A Comparison of Six DNA Extraction Protocols for 16S, ITS, and Shotgun Metagenomic Sequencing of Microbial Communities. BioTechniques 2022, 73, 34–46. [Google Scholar] [CrossRef] [PubMed]
  28. Available online: https://www.pacb.com/wp-content/uploads/Technical-Note-Preparing-DNA-for-PacBio-HiFi-Sequencing-Extraction-and-Quality-Control.pdf (accessed on 29 July 2022).
  29. Cai, W.; Wang, Y.; Hou, Q.; Zhang, Z.; Tang, F.; Shan, C.; Yang, X.; Guo, Z. PacBio Sequencing Combined with Metagenomic Shotgun Sequencing Provides Insight into the Microbial Diversity of Zha-Chili. Food BioSci. 2021, 40, 100884. [Google Scholar] [CrossRef]
  30. Quijada, N.M.; Schmitz-Esser, S.; Zwirzitz, B.; Guse, C.; Strachan, C.R.; Wagner, M.; Wetzels, S.U.; Selberherr, E.; Dzieciol, M. Austrian Raw-Milk Hard-Cheese Ripening Involves Successional Dynamics of Non-Inoculated Bacteria and Fungi. Foods 2020, 9, 1851. [Google Scholar] [CrossRef] [PubMed]
  31. Yang, C.; Zhao, F.; Hou, Q.; Wang, J.; Li, M.; Sun, Z. PacBio Sequencing Reveals Bacterial Community Diversity in Cheeses Collected from Different Regions. J. Dairy Sci. 2020, 103, 1238–1249. [Google Scholar] [CrossRef]
  32. Jones, A.; Torkel, C.; Stanley, D.; Nasim, J.; Borevitz, J.; Schwessinger, B. High-Molecular Weight DNA Extraction, Clean-up and Size Selection for Long-Read Sequencing. PLoS ONE 2021, 16, e0253830. [Google Scholar] [CrossRef] [PubMed]
  33. Mayjonade, B.; Gouzy, J.; Donnadieu, C.; Pouilly, N.; Marande, W.; Callot, C.; Langlade, N.; Muños, S. Extraction of High-Molecular-Weight Genomic DNA for Long-Read Sequencing of Single Molecules. BioTechniques 2016, 61, 203–205. [Google Scholar] [CrossRef] [PubMed]
  34. Ganda, E.; Beck, K.L.; Haiminen, N.; Silverman, J.D.; Kawas, B.; Cronk, B.D.; Anderson, R.R.; Goodman, L.B.; Wiedmann, M. DNA Extraction and Host Depletion Methods Significantly Impact and Potentially Bias Bacterial Detection in a Biological Fluid. mSystems 2021, 6, e00619-21. [Google Scholar] [CrossRef] [PubMed]
  35. Lim, M.Y.; Song, E.-J.; Kim, S.H.; Lee, J.; Nam, Y.-D. Comparison of DNA Extraction Methods for Human Gut Microbial Community Profiling. Syst. Appl. Microbiol. 2018, 41, 151–157. [Google Scholar] [CrossRef]
  36. Werner, D.; Acharya, K.; Blackburn, A.; Zan, R.; Plaimart, J.; Allen, B.; Mgana, S.M.; Sabai, S.M.; Halla, F.F.; Massawa, S.M.; et al. MinION Nanopore Sequencing Accelerates Progress towards Ubiquitous Genetics in Water Research. Water 2022, 14, 2491. [Google Scholar] [CrossRef]
  37. Tighe, S.; Afshinnekoo, E.; Rock, T.M.; McGrath, K.; Alexander, N.; McIntyre, A.; Ahsanuddin, S.; Bezdan, D.; Green, S.J.; Joye, S.; et al. Genomic Methods and Microbiological Technologies for Profiling Novel and Extreme Environments for the Extreme Microbiome Project (XMP). J. Biomol. Tech. 2017, 28, 31–39. [Google Scholar] [CrossRef]
  38. Feehery, G.R.; Yigit, E.; Oyola, S.O.; Langhorst, B.W.; Schmidt, V.T.; Stewart, F.J.; Dimalanta, E.T.; Amaral-Zettler, L.A.; Davis, T.; Quail, M.A.; et al. A Method for Selectively Enriching Microbial DNA from Contaminating Vertebrate Host DNA. PLoS ONE 2013, 8, e76096. [Google Scholar] [CrossRef] [PubMed]
  39. McHugh, A.J.; Feehily, C.; Hill, C.; Cotter, P.D. Detection and Enumeration of Spore-Forming Bacteria in Powdered Dairy Products. Front. Microbiol. 2017, 8, 109. [Google Scholar] [PubMed] [Green Version]
  40. Schuele, L.; Cassidy, H.; Peker, N.; Rossen, J.W.A.; Couto, N. Future Potential of Metagenomics in Microbiology Laboratories. Expert Rev. Mol. Diagn. 2021, 21, 1273–1285. [Google Scholar] [CrossRef]
  41. Shi, Y.; Wang, G.; Lau, H.C.-H.; Yu, J. Metagenomic Sequencing for Microbial DNA in Human Samples: Emerging Technological Advances. Int. J. Mol. Sci. 2022, 23, 2181. [Google Scholar] [CrossRef]
  42. McHugh, A.J.; Feehily, C.; Fenelon, M.A.; Gleeson, D.; Hill, C.; Cotter, P.D. Tracking the Dairy Microbiota from Farm Bulk Tank to Skimmed Milk Powder. mSystems 2020, 5, e00226-20. [Google Scholar] [CrossRef] [PubMed]
  43. Pereira-Marques, J.; Hout, A.; Ferreira, R.M.; Weber, M.; Pinto-Ribeiro, I.; van Doorn, L.-J.; Knetsch, C.W.; Figueiredo, C. Impact of Host DNA and Sequencing Depth on the Taxonomic Resolution of Whole Metagenome Sequencing for Microbiome Analysis. Front. Microbiol. 2019, 10, 1277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rubiola, S.; Chiesa, F.; Dalmasso, A.; Di Ciccio, P.; Civera, T. Detection of Antimicrobial Resistance Genes in the Milk Production Environment: Impact of Host DNA and Sequencing Depth. Front. Microbiol. 2020, 11, 1983. [Google Scholar] [CrossRef] [PubMed]
  45. Heravi, F.S.; Zakrzewski, M.; Vickery, K.; Hu, H. Host DNA Depletion Efficiency of Microbiome DNA Enrichment Methods in Infected Tissue Samples. J. Microbiol. Methods 2020, 170, 105856. [Google Scholar] [CrossRef] [PubMed]
  46. Marotz, C.A.; Sanders, J.G.; Zuniga, C.; Zaramela, L.S.; Knight, R.; Zengler, K. Improving Saliva Shotgun Metagenomics by Chemical Host DNA Depletion. Microbiome 2018, 6, 42. [Google Scholar] [CrossRef] [Green Version]
  47. Mo, L.; Yu, J.; Jin, H.; Hou, Q.; Yao, C.; Ren, D.; An, X.; Tsogtgerel, T.; Zhang, H. Investigating the Bacterial Microbiota of Traditional Fermented Dairy Products Using Propidium Monoazide with Single-Molecule Real-Time Sequencing. J. Dairy Sci. 2019, 102, 3912–3923. [Google Scholar] [CrossRef]
  48. Tantikachornkiat, M.; Sakakibara, S.; Neuner, M.; Durall, D.M. The Use of Propidium Monoazide in Conjunction with QPCR and Illumina Sequencing to Identify and Quantify Live Yeasts and Bacteria. Int. J. Food Microbiol. 2016, 234, 53–59. [Google Scholar] [CrossRef]
  49. Thoendel, M.; Jeraldo, P.R.; Greenwood-Quaintance, K.E.; Yao, J.Z.; Chia, N.; Hanssen, A.D.; Abdel, M.P.; Patel, R. Comparison of Microbial DNA Enrichment Tools for Metagenomic Whole Genome Sequencing. J. Microbiol. Methods 2016, 127, 141–145. [Google Scholar] [CrossRef] [Green Version]
  50. Marquet, M.; Zöllkau, J.; Pastuschek, J.; Viehweger, A.; Schleußner, E.; Makarewicz, O.; Pletz, M.W.; Ehricht, R.; Brandt, C. Evaluation of Microbiome Enrichment and Host DNA Depletion in Human Vaginal Samples Using Oxford Nanopore’s Adaptive Sequencing. Sci. Rep. 2022, 12, 4000. [Google Scholar] [CrossRef]
  51. Erkus, O.; de Jager, V.C.L.; Geene, R.T.C.M.; van Alen-Boerrigter, I.; Hazelwood, L.; van Hijum, S.A.F.T.; Kleerebezem, M.; Smid, E.J. Use of Propidium Monoazide for Selective Profiling of Viable Microbial Cells during Gouda Cheese Ripening. Int. J. Food Microbiol. 2016, 228, 1–9. [Google Scholar] [CrossRef]
  52. Cangelosi, G.A.; Meschke, J.S. Dead or Alive: Molecular Assessment of Microbial Viability. Appl. Environ. Microbiol. 2014, 80, 5884–5891. [Google Scholar] [CrossRef] [PubMed]
  53. Li, R.; Tun, H.M.; Jahan, M.; Zhang, Z.; Kumar, A.; Dilantha Fernando, W.G.; Farenhorst, A.; Khafipour, E. Comparison of DNA-, PMA-, and RNA-Based 16S RRNA Illumina Sequencing for Detection of Live Bacteria in Water. Sci. Rep. 2017, 7, 5752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Mancabelli, L.; Milani, C.; Anzalone, R.; Alessandri, G.; Lugli, G.A.; Tarracchini, C.; Fontana, F.; Turroni, F.; Ventura, M. Free DNA and Metagenomics Analyses: Evaluation of Free DNA Inactivation Protocols for Shotgun Metagenomics Analysis of Human Biological Matrices. Front. Microbiol. 2021, 12, 749373. [Google Scholar] [CrossRef] [PubMed]
  55. Shen, J.; McFarland, A.G.; Young, V.B.; Hayden, M.K.; Hartmann, E.M. Toward Accurate and Robust Environmental Surveillance Using Metagenomics. Front. Genet. 2021, 12, 600111. [Google Scholar] [CrossRef]
  56. Emerson, J.B.; Adams, R.I.; Román, C.M.B.; Brooks, B.; Coil, D.A.; Dahlhausen, K.; Ganz, H.H.; Hartmann, E.M.; Hsu, T.; Justice, N.B.; et al. Schrödinger’s Microbes: Tools for Distinguishing the Living from the Dead in Microbial Ecosystems. Microbiome 2017, 5, 86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Stinson, L.F.; Trevenen, M.L.; Geddes, D.T. The Viable Microbiome of Human Milk Differs from the Metataxonomic Profile. Nutrients 2021, 13, 4445. [Google Scholar] [CrossRef]
  58. Chen, W. Demystification of Fermented Foods by Omics Technologies. Curr. Opin. Food Sci. 2022, 46, 100845. [Google Scholar] [CrossRef]
  59. Balkir, P.; Kemahlioglu, K.; Yucel, U. Foodomics: A New Approach in Food Quality and Safety. Trends Food Sci. Technol. 2021, 108, 49–57. [Google Scholar] [CrossRef]
  60. Okeke, E.S.; Ita, R.E.; Egong, E.J.; Udofia, L.E.; Mgbechidinma, C.L.; Akan, O.D. Metaproteomics Insights into Fermented Fish and Vegetable Products and Associated Microbes. Food Chem. 2021, 3, 100045. [Google Scholar] [CrossRef]
  61. Zhao, Y.; Wu, Z.; Miyao, S.; Zhang, W. Unraveling the Flavor Profile and Microbial Roles during Industrial Sichuan Radish Paocai Fermentation by Molecular Sensory Science and Metatranscriptomics. Food Biosci. 2022, 48, 101815. [Google Scholar] [CrossRef]
  62. Heather, J.M.; Chain, B. The Sequence of Sequencers: The History of Sequencing DNA. Genomics 2016, 107, 1–8. [Google Scholar] [CrossRef] [PubMed]
  63. Allali, I.; Arnold, J.W.; Roach, J.; Cadenas, M.B.; Butz, N.; Hassan, H.M.; Koci, M.; Ballou, A.; Mendoza, M.; Ali, R.; et al. A Comparison of Sequencing Platforms and Bioinformatics Pipelines for Compositional Analysis of the Gut Microbiome. BMC Microbiol. 2017, 17, 194. [Google Scholar] [CrossRef] [PubMed]
  64. Gupta, A.K.; Gupta, U. Chapter 20—Next Generation Sequencing and Its Applications. In Animal Biotechnology, 2nd ed.; Verma, A.S., Singh, A., Eds.; Academic Press: Boston, MA, USA, 2020; pp. 395–421. ISBN 978-0-12-811710-1. [Google Scholar]
  65. Thomas, T.; Gilbert, J.; Meyer, F. Metagenomics—A Guide from Sampling to Data Analysis. Microb. Inform. Exp. 2012, 2, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ma, X.; Shao, Y.; Tian, L.; Flasch, D.A.; Mulder, H.L.; Edmonson, M.N.; Liu, Y.; Chen, X.; Newman, S.; Nakitandwe, J.; et al. Analysis of Error Profiles in Deep Next-Generation Sequencing Data. Genome Biol. 2019, 20, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Stoler, N.; Nekrutenko, A. Sequencing Error Profiles of Illumina Sequencing Instruments. NAR Genom. Bioinform. 2021, 3, lqab019. [Google Scholar] [CrossRef] [PubMed]
  68. Kanwar, N.; Blanco, C.; Chen, I.A.; Seelig, B. PacBio Sequencing Output Increased through Uniform and Directional Fivefold Concatenation. Sci. Rep. 2021, 11, 18065. [Google Scholar] [CrossRef] [PubMed]
  69. Aunin, E.; Böhme, U.; Blake, D.; Dove, A.; Smith, M.; Corton, C.; Oliver, K.; Betteridge, E.; Quail, M.A.; McCarthy, S.A.; et al. The Complete Genome Sequence of Eimeria Tenella (Tyzzer 1929), a Common Gut Parasite of Chickens. Wellcome Open Res. 2021, 6, 225. [Google Scholar] [CrossRef]
  70. Kenny, N.J.; McCarthy, S.A.; Dudchenko, O.; James, K.; Betteridge, E.; Corton, C.; Dolucan, J.; Mead, D.; Oliver, K.; Omer, A.D.; et al. The Gene-Rich Genome of the Scallop Pecten Maximus. GigaScience 2020, 9, giaa037. [Google Scholar] [CrossRef]
  71. Rhie, A.; McCarthy, S.A.; Fedrigo, O.; Damas, J.; Formenti, G.; Koren, S.; Uliano-Silva, M.; Chow, W.; Fungtammasan, A.; Kim, J.; et al. Towards Complete and Error-Free Genome Assemblies of All Vertebrate Species. Nature 2021, 592, 737–746. [Google Scholar] [CrossRef]
  72. Hu, T.; Chitnis, N.; Monos, D.; Dinh, A. Next-Generation Sequencing Technologies: An Overview. Hum. Immunol. 2021, 82, 801–811. [Google Scholar] [CrossRef]
  73. Zhang, H.; Jain, C.; Aluru, S. A Comprehensive Evaluation of Long Read Error Correction Methods. BMC Genom. 2020, 21, 889. [Google Scholar] [CrossRef] [PubMed]
  74. Amarasinghe, S.L.; Su, S.; Dong, X.; Zappia, L.; Ritchie, M.E.; Gouil, Q. Opportunities and Challenges in Long-Read Sequencing Data Analysis. Genome Biol. 2020, 21, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Dohm, J.C.; Peters, P.; Stralis-Pavese, N.; Himmelbauer, H. Benchmarking of Long-Read Correction Methods. NAR Genom. Bioinform. 2020, 2, lqaa037. [Google Scholar] [CrossRef] [PubMed]
  76. Cornelis, S. Forensic Lab-on-a-Chip DNA Analysis. Ph.D. Thesis, Ghent University, Ghent, Belgium, 2019. [Google Scholar]
  77. Lin, B.; Hui, J.; Mao, H. Nanopore Technology and Its Applications in Gene Sequencing. Biosensors 2021, 11, 214. [Google Scholar] [CrossRef] [PubMed]
  78. Available online: https://www.pacb.com/wp-content/uploads/Baybayan-PAG-2017-Best-Practices-for-Whole-Genome-Sequencing-Using-the-Sequel-System.pdf (accessed on 18 July 2022).
  79. Kim, K.E.; Peluso, P.; Babayan, P.; Yeadon, P.J.; Yu, C.; Fisher, W.W.; Chin, C.-S.; Rapicavoli, N.A.; Rank, D.R.; Li, J.; et al. Long-Read, Whole-Genome Shotgun Sequence Data for Five Model Organisms. Sci. Data 2014, 1, 140045. [Google Scholar] [CrossRef] [Green Version]
  80. Available online: https://www.pacb.com/wp-content/uploads/Procedure-Checklist-%E2%80%93-Preparing-10-kb-Library-Using-SMRTbell-Express-Template-Prep-Kit-2.0-for-Metagenomics-Shotgun-Sequencing.pdf (accessed on 29 July 2022).
  81. Kircher, M.; Sawyer, S.; Meyer, M. Double Indexing Overcomes Inaccuracies in Multiplex Sequencing on the Illumina Platform. Nucleic Acids Res. 2012, 40, e3. [Google Scholar] [CrossRef] [Green Version]
  82. Aigrain, L. Beginner’s Guide to next-Generation Sequencing. Biochem 2021, 43, 58–64. [Google Scholar] [CrossRef]
  83. MacConaill, L.E.; Burns, R.T.; Nag, A.; Coleman, H.A.; Slevin, M.K.; Giorda, K.; Light, M.; Lai, K.; Jarosz, M.; McNeill, M.S.; et al. Unique, Dual-Indexed Sequencing Adapters with UMIs Effectively Eliminate Index Cross-Talk and Significantly Improve Sensitivity of Massively Parallel Sequencing. BMC Genom. 2018, 19, 30. [Google Scholar] [CrossRef] [Green Version]
  84. Sinha, R.; Stanley, G.; Gulati, G.S.; Ezran, C.; Travaglini, K.J.; Wei, E.; Chan, C.K.F.; Nabhan, A.N.; Su, T.; Morganti, R.M.; et al. Index Switching Causes “Spreading-of-Signal” among Multiplexed Samples in Illumina HiSeq 4000 DNA Sequencing. BioRxiv 2017. [Google Scholar] [CrossRef]
  85. Ros-Freixedes, R.; Battagin, M.; Johnsson, M.; Gorjanc, G.; Mileham, A.J.; Rounsley, S.D.; Hickey, J.M. Impact of Index Hopping and Bias towards the Reference Allele on Accuracy of Genotype Calls from Low-Coverage Sequencing. Genet. Sel. Evol. 2018, 50, 64. [Google Scholar] [CrossRef] [Green Version]
  86. 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. 2020, 20, 1171–1181. [Google Scholar] [CrossRef] [PubMed]
  87. Wright, E.S.; Vetsigian, K.H. Quality Filtering of Illumina Index Reads Mitigates Sample Cross-Talk. BMC Genom. 2016, 17, 876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Guenay-Greunke, Y.; Bohan, D.A.; Traugott, M.; Wallinger, C. Handling of Targeted Amplicon Sequencing Data Focusing on Index Hopping and Demultiplexing Using a Nested Metabarcoding Approach in Ecology. Sci. Rep. 2021, 11, 19510. [Google Scholar] [CrossRef]
  89. Li, Q.; Zhao, X.; Zhang, W.; Wang, L.; Wang, J.; Xu, D.; Mei, Z.; Liu, Q.; Du, S.; Li, Z.; et al. Reliable Multiplex Sequencing with Rare Index Mis-Assignment on DNB-Based NGS Platform. BMC Genom. 2019, 20, 215. [Google Scholar] [CrossRef] [Green Version]
  90. Salter, S.J.; Cox, M.J.; Turek, E.M.; Calus, S.T.; Cookson, W.O.; Moffatt, M.F.; Turner, P.; Parkhill, J.; Loman, N.J.; Walker, A.W. Reagent and Laboratory Contamination Can Critically Impact Sequence-Based Microbiome Analyses. BMC Biol. 2014, 12, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Yeh, Y.-C.; Needham, D.M.; Sieradzki, E.T.; Fuhrman, J.A. Taxon Disappearance from Microbiome Analysis Reinforces the Value of Mock Communities as a Standard in Every Sequencing Run. mSystems 2018, 3, e00023-18. [Google Scholar] [CrossRef] [Green Version]
  92. Frau, A.; Kenny, J.G.; Lenzi, L.; Campbell, B.J.; Ijaz, U.Z.; Duckworth, C.A.; Burkitt, M.D.; Hall, N.; Anson, J.; Darby, A.C.; et al. DNA Extraction and Amplicon Production Strategies Deeply Inf Luence the Outcome of Gut Mycobiome Studies. Sci. Rep. 2019, 9, 9328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Tedersoo, L.; Bahram, M.; Zinger, L.; Nilsson, R.H.; Kennedy, P.G.; Yang, T.; Anslan, S.; Mikryukov, V. Best Practices in Metabarcoding of Fungi: From Experimental Design to Results. Mol. Ecol. 2022, 31, 2769–2795. [Google Scholar] [CrossRef]
  94. Jay, Z.J.; Inskeep, W.P. The Distribution, Diversity, and Importance of 16S RRNA Gene Introns in the Order Thermoproteales. Biol. Direct 2015, 10, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Bukin, Y.S.; Galachyants, Y.P.; Morozov, I.V.; Bukin, S.V.; Zakharenko, A.S.; Zemskaya, T.I. The Effect of 16S RRNA Region Choice on Bacterial Community Metabarcoding Results. Sci. Data 2019, 6, 190007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Chakravorty, S.; Helb, D.; Burday, M.; Connell, N.; Alland, D. A Detailed Analysis of 16S Ribosomal RNA Gene Segments for the Diagnosis of Pathogenic Bacteria. J. Microbiol. Methods 2007, 69, 330–339. [Google Scholar] [CrossRef]
  97. Stefanini, I.; Cavalieri, D. Metagenomic Approaches to Investigate the Contribution of the Vineyard Environment to the Quality of Wine Fermentation: Potentials and Difficulties. Front. Microbiol. 2018, 9, 991. [Google Scholar] [CrossRef] [PubMed]
  98. Amrouche, T.; Mounier, J.; Pawtowski, A.; Thomas, F.; Picot, A. Microbiota Associated with Dromedary Camel Milk from Algerian Sahara. Curr. Microbiol. 2020, 77, 24–31. [Google Scholar] [CrossRef]
  99. Maillet, A.; Bouju-Albert, A.; Roblin, S.; Vaissié, P.; Leuillet, S.; Dousset, X.; Jaffrès, E.; Combrisson, J.; Prévost, H. Impact of DNA Extraction and Sampling Methods on Bacterial Communities Monitored by 16S RDNA Metabarcoding in Cold-Smoked Salmon and Processing Plant Surfaces. Food Microbiol. 2021, 95, 103705. [Google Scholar] [CrossRef] [PubMed]
  100. Michailidou, S.; Petrovits, G.E.; Kyritsi, M.; Argiriou, A. Amplicon Metabarcoding Data of Prokaryotes and Eukaryotes Present in ‘Kalamata’ Table Olives Packaged under Modified Atmosphere. Data Brief 2021, 38, 107314. [Google Scholar] [CrossRef]
  101. Penland, M.; Mounier, J.; Pawtowski, A.; Tréguer, S.; Deutsch, S.-M.; Coton, M. Use of Metabarcoding and Source Tracking to Identify Desirable or Spoilage Autochthonous Microorganism Sources during Black Olive Fermentations. Food Res. Int. 2021, 144, 110344. [Google Scholar] [CrossRef] [PubMed]
  102. Hall, M.; Beiko, R.G. 16S RRNA Gene Analysis with QIIME2. In Microbiome Analysis: Methods and Protocols; Beiko, R.G., Hsiao, W., Parkinson, J., Eds.; Methods in Molecular Biology; Springer: New York, NY, USA, 2018; pp. 113–129. ISBN 978-1-4939-8728-3. [Google Scholar]
  103. Schloss, P.D.; Westcott, S.L.; Ryabin, T.; Hall, J.R.; Hartmann, M.; Hollister, E.B.; Lesniewski, R.A.; Oakley, B.B.; Parks, D.H.; Robinson, C.J.; et al. Introducing Mothur: Open-Source, Platform-Independent, Community-Supported Software for Describing and Comparing Microbial Communities. Appl. Environ. Microbiol. 2009, 75, 7537–7541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Schloss, P.D. Reintroducing Mothur: 10 Years Later. Appl. Environ. Microbiol. 2020, 86, e02343-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Meyer, F.; Paarmann, D.; D’Souza, M.; Olson, R.; Glass, E.; Kubal, M.; Paczian, T.; Rodriguez, A.; Stevens, R.; Wilke, A.; et al. The Metagenomics RAST Server—A Public Resource for the Automatic Phylogenetic and Functional Analysis of Metagenomes. BMC Bioinform. 2008, 9, 386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Edgar, R.C. UPARSE: Highly Accurate OTU Sequences from Microbial Amplicon Reads. Nat. Methods 2013, 10, 996–998. [Google Scholar] [CrossRef] [PubMed]
  107. Escudié, F.; Auer, L.; Bernard, M.; Mariadassou, M.; Cauquil, L.; Vidal, K.; Maman, S.; Hernandez-Raquet, G.; Combes, S.; Pascal, G. FROGS: Find, Rapidly, OTUs with Galaxy Solution. Bioinformatics 2018, 34, 1287–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Gentleman, R.C.; Carey, V.J.; Bates, D.M.; Bolstad, B.; Dettling, M.; Dudoit, S.; Ellis, B.; Gautier, L.; Ge, Y.; Gentry, J.; et al. Bioconductor: Open Software Development for Computational Biology and Bioinformatics. Genome Biol. 2004, 5, R80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Amir, A.; McDonald, D.; Navas-Molina, J.A.; Kopylova, E.; Morton, J.T.; Zech Xu, Z.; Kightley, E.P.; Thompson, L.R.; Hyde, E.R.; Gonzalez, A.; et al. Deblur Rapidly Resolves Single-Nucleotide Community Sequence Patterns. mSystems 2017, 2, e00191-16. [Google Scholar] [CrossRef] [Green Version]
  110. Callahan, B.J.; McMurdie, P.J.; Rosen, M.J.; Han, A.W.; Johnson, A.J.A.; Holmes, S.P. DADA2: High-Resolution Sample Inference from Illumina Amplicon Data. Nat. Methods 2016, 13, 581–583. [Google Scholar] [CrossRef] [Green Version]
  111. Liu, Y.-X.; Qin, Y.; Chen, T.; Lu, M.; Qian, X.; Guo, X.; Bai, Y. A Practical Guide to Amplicon and Metagenomic Analysis of Microbiome Data. Protein Cell 2021, 12, 315–330. [Google Scholar] [CrossRef] [PubMed]
  112. Gao, B.; Chi, L.; Zhu, Y.; Shi, X.; Tu, P.; Li, B.; Yin, J.; Gao, N.; Shen, W.; Schnabl, B. An Introduction to Next Generation Sequencing Bioinformatic Analysis in Gut Microbiome Studies. Biomolecules 2021, 11, 530. [Google Scholar] [CrossRef] [PubMed]
  113. Wajid, B.; Anwar, F.; Wajid, I.; Nisar, H.; Meraj, S.; Zafar, A.; Al-Shawaqfeh, M.K.; Ekti, A.R.; Khatoon, A.; Suchodolski, J.S. Music of Metagenomics—A Review of Its Applications, Analysis Pipeline, and Associated Tools. Funct. Integr. Genom. 2022, 22, 3–26. [Google Scholar] [CrossRef]
  114. Nakano, K.; Shiroma, A.; Shimoji, M.; Tamotsu, H.; Ashimine, N.; Ohki, S.; Shinzato, M.; Minami, M.; Nakanishi, T.; Teruya, K.; et al. Advantages of Genome Sequencing by Long-Read Sequencer Using SMRT Technology in Medical Area. Hum. Cell 2017, 30, 149–161. [Google Scholar] [CrossRef] [Green Version]
  115. Jin, H.; Mo, L.; Pan, L.; Hou, Q.; Li, C.; Darima, I.; Yu, J. Using PacBio Sequencing to Investigate the Bacterial Microbiota of Traditional Buryatian Cottage Cheese and Comparison with Italian and Kazakhstan Artisanal Cheeses. J. Dairy Sci. 2018, 101, 6885–6896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yang, J.; Cao, J.; Xu, H.; Hou, Q.; Yu, Z.; Zhang, H.; Sun, Z. Bacterial Diversity and Community Structure in Chongqing Radish Paocai Brines Revealed Using PacBio Single-Molecule Real-Time Sequencing Technology. J. Sci. Food Agric. 2018, 98, 3234–3245. [Google Scholar] [CrossRef] [PubMed]
  117. Cuscó, A.; Catozzi, C.; Viñes, J.; Sanchez, A.; Francino, O. Microbiota Profiling with Long Amplicons Using Nanopore Sequencing: Full-Length 16S RRNA Gene and the 16S-ITS-23S of the rrn Operon. F1000Research 2019, 7, 1755. [Google Scholar] [CrossRef] [PubMed]
  118. De Oliveira Martins, L.; Page, A.J.; Mather, A.E.; Charles, I.G. Taxonomic Resolution of the Ribosomal RNA Operon in Bacteria: Implications for Its Use with Long-Read Sequencing. NAR Genom. Bioinform. 2020, 2, lqz016. [Google Scholar] [CrossRef] [Green Version]
  119. Gehrig, J.L.; Portik, D.M.; Driscoll, M.D.; Jackson, E.; Chakraborty, S.; Gratalo, D.; Ashby, M.; Valladares, R. Finding the Right Fit: Evaluation of Short-Read and Long-Read Sequencing Approaches to Maximize the Utility of Clinical Microbiome Data. Microb. Genom. 2022, 8, 000794. [Google Scholar] [CrossRef] [PubMed]
  120. Kerkhof, L.J.; Dillon, K.P.; Häggblom, M.M.; McGuinness, L.R. Profiling Bacterial Communities by MinION Sequencing of Ribosomal Operons. Microbiome 2017, 5, 116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Kinoshita, Y.; Niwa, H.; Uchida-Fujii, E.; Nukada, T. Establishment and Assessment of an Amplicon Sequencing Method Targeting the 16S-ITS-23S RRNA Operon for Analysis of the Equine Gut Microbiome. Sci. Rep. 2021, 11, 11884. [Google Scholar] [CrossRef] [PubMed]
  122. Martijn, J.; Lind, A.E.; Schön, M.E.; Spiertz, I.; Juzokaite, L.; Bunikis, I.; Pettersson, O.V.; Ettema, T.J.G. Confident Phylogenetic Identification of Uncultured Prokaryotes through Long Read Amplicon Sequencing of the 16S-ITS-23S RRNA Operon. Environ. Microbiol. 2019, 21, 2485–2498. [Google Scholar] [CrossRef]
  123. Karst, S.M.; Ziels, R.M.; Kirkegaard, R.H.; Sørensen, E.A.; McDonald, D.; Zhu, Q.; Knight, R.; Albertsen, M. High-Accuracy Long-Read Amplicon Sequences Using Unique Molecular Identifiers with Nanopore or PacBio Sequencing. Nat. Methods 2021, 18, 165–169. [Google Scholar] [CrossRef]
  124. Brewer, T.E.; Albertsen, M.; Edwards, A.; Kirkegaard, R.H.; Rocha, E.P.C.; Fierer, N. Unlinked RRNA Genes Are Widespread among Bacteria and Archaea. ISME J. 2020, 14, 597–608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Bik, H.M. Just Keep It Simple? Benchmarking the Accuracy of Taxonomy Assignment Software in Metabarcoding Studies. Mol. Ecol. Resour. 2021, 21, 2187–2189. [Google Scholar] [CrossRef] [PubMed]
  126. Stoddard, S.F.; Smith, B.J.; Hein, R.; Roller, B.R.K.; Schmidt, T.M. RrnDB: Improved Tools for Interpreting RRNA Gene Abundance in Bacteria and Archaea and a New Foundation for Future Development. Nucleic Acids Res. 2015, 43, D593–D598. [Google Scholar] [CrossRef] [Green Version]
  127. Available online: https://www.pacb.com/wp-content/uploads/Driscoll-ASM-Microbe-2019-Microbiome-Profiling-at-the-Strain-Level-Using-rRNA-Amplicons.pdf (accessed on 18 July 2022).
  128. Seol, D.; Lim, J.S.; Sung, S.; Lee, Y.H.; Jeong, M.; Cho, S.; Kwak, W.; Kim, H. Microbial Identification Using RRNA Operon Region: Database and Tool for Metataxonomics with Long-Read Sequence. Microbiol. Spectr. 2022, 10, e02017–e02021. [Google Scholar] [CrossRef] [PubMed]
  129. Benítez-Páez, A.; Portune, K.J.; Sanz, Y. Species-Level Resolution of 16S RRNA Gene Amplicons Sequenced through the MinIONTM Portable Nanopore Sequencer. GigaScience 2016, 5, s13742-016-0111-z. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Ciuffreda, L.; Rodríguez-Pérez, H.; Flores, C. Nanopore Sequencing and Its Application to the Study of Microbial Communities. Comput. Struct. Biotechnol. J. 2021, 19, 1497–1511. [Google Scholar] [CrossRef] [PubMed]
  131. Li, H. Minimap2: Pairwise Alignment for Nucleotide Sequences. Bioinformatics 2018, 34, 3094–3100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Douglas, G.M.; Maffei, V.J.; Zaneveld, J.R.; Yurgel, S.N.; Brown, J.R.; Taylor, C.M.; Huttenhower, C.; Langille, M.G.I. PICRUSt2 for Prediction of Metagenome Functions. Nat. Biotechnol. 2020, 38, 685–688. [Google Scholar] [CrossRef]
  133. Aßhauer, K.P.; Wemheuer, B.; Daniel, R.; Meinicke, P. Tax4Fun: Predicting Functional Profiles from Metagenomic 16S RRNA Data. Bioinformatics 2015, 31, 2882–2884. [Google Scholar] [CrossRef] [Green Version]
  134. Zotta, T.; Ricciardi, A.; Condelli, N.; Parente, E. Metataxonomic and Metagenomic Approaches for the Study of Undefined Strain Starters for Cheese Manufacture. Crit. Rev. Food Sci. Nutr. 2022, 62, 3898–3912. [Google Scholar] [CrossRef]
  135. Langmead, B.; Salzberg, S.L. Fast Gapped-Read Alignment with Bowtie 2. Nat. Methods 2012, 9, 357–359. [Google Scholar] [CrossRef] [Green Version]
  136. Menzel, P.; Ng, K.L.; Krogh, A. Fast and Sensitive Taxonomic Classification for Metagenomics with Kaiju. Nat. Commun. 2016, 7, 11257. [Google Scholar] [CrossRef] [Green Version]
  137. Wood, D.E.; Salzberg, S.L. Kraken: Ultrafast Metagenomic Sequence Classification Using Exact Alignments. Genome Biol. 2014, 15, R46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Wood, D.E.; Lu, J.; Langmead, B. Improved Metagenomic Analysis with Kraken 2. Genome Biol. 2019, 20, 257. [Google Scholar] [CrossRef] [PubMed]
  139. Segata, N.; Waldron, L.; Ballarini, A.; Narasimhan, V.; Jousson, O.; Huttenhower, C. Metagenomic Microbial Community Profiling Using Unique Clade-Specific Marker Genes. Nat. Methods 2012, 9, 811–814. [Google Scholar] [CrossRef] [PubMed]
  140. Tovo, A.; Menzel, P.; Krogh, A.; Cosentino Lagomarsino, M.; Suweis, S. Taxonomic Classification Method for Metagenomics Based on Core Protein Families with Core-Kaiju. Nucleic Acids Res. 2020, 48, e93. [Google Scholar] [CrossRef] [PubMed]
  141. Ye, S.H.; Siddle, K.J.; Park, D.J.; Sabeti, P.C. Benchmarking Metagenomics Tools for Taxonomic Classification. Cell 2019, 178, 779–794. [Google Scholar] [CrossRef]
  142. Nasko, D.J.; Koren, S.; Phillippy, A.M.; Treangen, T.J. RefSeq Database Growth Influences the Accuracy of K-Mer-Based Lowest Common Ancestor Species Identification. Genome Biol. 2018, 19, 165. [Google Scholar] [CrossRef] [Green Version]
  143. Zhou, Y.; Liu, M.; Yang, J. Recovering Metagenome-Assembled Genomes from Shotgun Metagenomic Sequencing Data: Methods, Applications, Challenges, and Opportunities. Microbiol. Res. 2022, 260, 127023. [Google Scholar] [CrossRef]
  144. Ayling, M.; Clark, M.D.; Leggett, R.M. New Approaches for Metagenome Assembly with Short Reads. Brief Bioinform. 2020, 21, 584–594. [Google Scholar] [CrossRef] [Green Version]
  145. Nurk, S.; Meleshko, D.; Korobeynikov, A.; Pevzner, P.A. MetaSPAdes: A New Versatile Metagenomic Assembler. Genome Res. 2017, 27, 824–834. [Google Scholar] [CrossRef] [Green Version]
  146. Peng, Y.; Leung, H.C.M.; Yiu, S.M.; Chin, F.Y.L. Meta-IDBA: A de Novo Assembler for Metagenomic Data. Bioinformatics 2011, 27, i94–i101. [Google Scholar] [CrossRef] [Green Version]
  147. Kang, D.D.; Froula, J.; Egan, R.; Wang, Z. MetaBAT, an Efficient Tool for Accurately Reconstructing Single Genomes from Complex Microbial Communities. PeerJ 2015, 3, e1165. [Google Scholar] [CrossRef] [Green Version]
  148. Alneberg, J.; Bjarnason, B.S.; de Bruijn, I.; Schirmer, M.; Quick, J.; Ijaz, U.Z.; Loman, N.J.; Andersson, A.F.; Quince, C. CONCOCT: Clustering CONtigs on COverage and ComposiTion. arXiv 2013, arXiv:1312.4038. [Google Scholar]
  149. Li, D.; Liu, C.-M.; Luo, R.; Sadakane, K.; Lam, T.-W. MEGAHIT: An Ultra-Fast Single-Node Solution for Large and Complex Metagenomics Assembly via Succinct de Bruijn Graph. Bioinformatics 2015, 31, 1674–1676. [Google Scholar] [CrossRef] [PubMed]
  150. Wu, Y.-W.; Tang, Y.-H.; Tringe, S.G.; Simmons, B.A.; Singer, S.W. MaxBin: An Automated Binning Method to Recover Individual Genomes from Metagenomes Using an Expectation-Maximization Algorithm. Microbiome 2014, 2, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Maguire, F.; Jia, B.; Gray, K.L.; Lau, W.Y.V.; Beiko, R.G.; Brinkman, F.S.L.Y. Metagenome-Assembled Genome Binning Methods with Short Reads Disproportionately Fail for Plasmids and Genomic Islands. Microb. Genom. 2020, 6, e000436. [Google Scholar] [CrossRef]
  152. Xie, H.; Yang, C.; Sun, Y.; Igarashi, Y.; Jin, T.; Luo, F. PacBio Long Reads Improve Metagenomic Assemblies, Gene Catalogs, and Genome Binning. Front. Genet. 2020, 11, 516269. [Google Scholar] [CrossRef] [PubMed]
  153. De Coster, W.; Weissensteiner, M.H.; Sedlazeck, F.J. Towards Population-Scale Long-Read Sequencing. Nat. Rev. Genet. 2021, 22, 572–587. [Google Scholar] [CrossRef] [PubMed]
  154. Tedersoo, L.; Albertsen, M.; Anslan, S.; Callahan, B. Perspectives and Benefits of High-Throughput Long-Read Sequencing in Microbial Ecology. Appl. Environ. Microbiol. 2021, 87, e00626-21. [Google Scholar] [CrossRef]
  155. Hu, Y.; Fang, L.; Nicholson, C.; Wang, K. Implications of Error-Prone Long-Read Whole-Genome Shotgun Sequencing on Characterizing Reference Microbiomes. iScience 2020, 23, 101223. [Google Scholar] [CrossRef]
  156. Portik, D.M.; Brown, C.T.; Pierce-Ward, N.T. Evaluation of Taxonomic Profiling Methods for Long-Read Shotgun Metagenomic Sequencing Datasets. bioRxiv 2022. [Google Scholar] [CrossRef]
  157. Fu, S.; Wang, A.; Au, K.F. A Comparative Evaluation of Hybrid Error Correction Methods for Error-Prone Long Reads. Genome Biol. 2019, 20, 26. [Google Scholar] [CrossRef] [Green Version]
  158. Wang, Y.; Zhao, Y.; Bollas, A.; Wang, Y.; Au, K.F. Nanopore Sequencing Technology, Bioinformatics and Applications. Nat. Biotechnol. 2021, 39, 1348–1365. [Google Scholar] [CrossRef] [PubMed]
  159. De Filippis, F.; Parente, E.; Ercolini, D. Recent Past, Present, and Future of the Food Microbiome. Available online: https://www.annualreviews.org/doi/epdf/10.1146/annurev-food-030117-012312 (accessed on 18 July 2022).
  160. Durazzi, F.; Sala, C.; Castellani, G.; Manfreda, G.; Remondini, D.; De Cesare, A. Comparison between 16S RRNA and Shotgun Sequencing Data for the Taxonomic Characterization of the Gut Microbiota. Sci. Rep. 2021, 11, 3030. [Google Scholar] [CrossRef] [PubMed]
  161. Arıkan, M.; Mitchell, A.L.; Finn, R.D.; Gürel, F. Microbial Composition of Kombucha Determined Using Amplicon Sequencing and Shotgun Metagenomics. J. Food Sci. 2020, 85, 455–464. [Google Scholar] [CrossRef]
  162. You, L.; Yang, C.; Jin, H.; Kwok, L.-Y.; Sun, Z.; Zhang, H. Metagenomic Features of Traditional Fermented Milk Products. LWT 2022, 155, 112945. [Google Scholar] [CrossRef]
  163. Maske, B.L.; de Melo Pereira, G.V.; da Silva Vale, A.; Marques Souza, D.S.; De Dea Lindner, J.; Soccol, C.R. Viruses in Fermented Foods: Are They Good or Bad? Two Sides of the Same Coin. Food Microbiol. 2021, 98, 103794. [Google Scholar] [CrossRef] [PubMed]
  164. Tagirdzhanova, G.; Saary, P.; Tingley, J.P.; Díaz-Escandón, D.; Abbott, D.W.; Finn, R.D.; Spribille, T. Predicted Input of Uncultured Fungal Symbionts to a Lichen Symbiosis from Metagenome-Assembled Genomes. Genome Biol. Evol. 2021, 13, evab047. [Google Scholar] [CrossRef] [PubMed]
  165. Antipov, D.; Raiko, M.; Lapidus, A.; Pevzner, P.A. Plasmid Detection and Assembly in Genomic and Metagenomic Data Sets. Genome Res. 2019, 29, 961–968. [Google Scholar] [CrossRef] [Green Version]
  166. Beaulaurier, J.; Zhu, S.; Deikus, G.; Mogno, I.; Zhang, X.-S.; Davis-Richardson, A.; Canepa, R.; Triplett, E.W.; Faith, J.J.; Sebra, R.; et al. Metagenomic Binning and Association of Plasmids with Bacterial Host Genomes Using DNA Methylation. Nat. Biotechnol. 2018, 36, 61–69. [Google Scholar] [CrossRef]
  167. Hilpert, C.; Bricheux, G.; Debroas, D. Reconstruction of Plasmids by Shotgun Sequencing from Environmental DNA: Which Bioinformatic Workflow? Brief Bioinform. 2021, 22, bbaa059. [Google Scholar] [CrossRef]
  168. Callahan, B.J.; Grinevich, D.; Thakur, S.; Balamotis, M.A.; Yehezkel, T.B. Ultra-Accurate Microbial Amplicon Sequencing with Synthetic Long Reads. Microbiome 2021, 9, 130. [Google Scholar] [CrossRef]
  169. Liu, S.; Wu, I.; Yu, Y.-P.; Balamotis, M.; Ren, B.; Ben Yehezkel, T.; Luo, J.-H. Targeted Transcriptome Analysis Using Synthetic Long Read Sequencing Uncovers Isoform Reprograming in the Progression of Colon Cancer. Commun Biol. 2021, 4, 506. [Google Scholar] [CrossRef] [PubMed]
  170. Li, R.; Hsieh, C.-L.; Young, A.; Zhang, Z.; Ren, X.; Zhao, Z. Illumina Synthetic Long Read Sequencing Allows Recovery of Missing Sequences Even in the “Finished” C. Elegans Genome. Sci. Rep. 2015, 5, 10814. [Google Scholar] [CrossRef]
  171. Burton, J.N.; Liachko, I.; Dunham, M.J.; Shendure, J. Species-Level Deconvolution of Metagenome Assemblies with Hi-C–Based Contact Probability Maps. G3 2014, 4, 1339–1346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Elbers, J.P.; Rogers, M.F.; Perelman, P.L.; Proskuryakova, A.A.; Serdyukova, N.A.; Johnson, W.E.; Horin, P.; Corander, J.; Murphy, D.; Burger, P.A. Improving Illumina Assemblies with Hi-C and Long Reads: An Example with the North African Dromedary. Mol. Ecol. Resour. 2019, 19, 1015–1026. [Google Scholar] [CrossRef] [PubMed]
  173. Kong, S.; Zhang, Y. Deciphering Hi-C: From 3D Genome to Function. Cell Biol. Toxicol. 2019, 35, 15–32. [Google Scholar] [CrossRef] [PubMed]
  174. Ning, D.-L.; Wu, T.; Xiao, L.-J.; Ma, T.; Fang, W.-L.; Dong, R.-Q.; Cao, F.-L. Chromosomal-Level Assembly of Juglans Sigillata Genome Using Nanopore, BioNano, and Hi-C Analysis. GigaScience 2020, 9, giaa006. [Google Scholar] [CrossRef]
  175. Bickhart, D.M.; Kolmogorov, M.; Tseng, E.; Portik, D.M.; Korobeynikov, A.; Tolstoganov, I.; Uritskiy, G.; Liachko, I.; Sullivan, S.T.; Shin, S.B.; et al. Generating Lineage-Resolved, Complete Metagenome-Assembled Genomes from Complex Microbial Communities. Nat. Biotechnol. 2022, 40, 711–719. [Google Scholar] [CrossRef]
  176. Jagadeesan, B.; Gerner-Smidt, P.; Allard, M.W.; Leuillet, S.; Winkler, A.; Xiao, Y.; Chaffron, S.; Van Der Vossen, J.; Tang, S.; Katase, M.; et al. The Use of next Generation Sequencing for Improving Food Safety: Translation into Practice. Food Microbiol. 2019, 79, 96–115. [Google Scholar] [CrossRef]
  177. Bao, Y.; Wadden, J.; Erb-Downward, J.R.; Ranjan, P.; Zhou, W.; McDonald, T.L.; Mills, R.E.; Boyle, A.P.; Dickson, R.P.; Blaauw, D.; et al. SquiggleNet: Real-Time, Direct Classification of Nanopore Signals. Genome Biol. 2021, 22, 298. [Google Scholar] [CrossRef]
  178. Cao, M.D.; Ganesamoorthy, D.; Elliott, A.G.; Zhang, H.; Cooper, M.A.; Coin, L.J.M. Streaming Algorithms for Identification Pathogens and Antibiotic Resistance Potential from Real-Time MinIONTM Sequencing. GigaScience 2016, 5, 32. [Google Scholar] [CrossRef] [Green Version]
  179. Juul, S.; Izquierdo, F.; Hurst, A.; Dai, X.; Wright, A.; Kulesha, E.; Pettett, R.; Turner, D.J. What’s in My Pot? Real-Time Species Identification on the MinIONTM. bioRxiv 2015. [Google Scholar] [CrossRef] [Green Version]
  180. Edwards, H.S.; Krishnakumar, R.; Sinha, A.; Bird, S.W.; Patel, K.D.; Bartsch, M.S. Real-Time Selective Sequencing with RUBRIC: Read Until with Basecall and Reference-Informed Criteria. Sci. Rep. 2019, 9, 11475. [Google Scholar] [CrossRef] [PubMed]
  181. Payne, A.; Holmes, N.; Clarke, T.; Munro, R.; Debebe, B.J.; Loose, M. Readfish Enables Targeted Nanopore Sequencing of Gigabase-Sized Genomes. Nat. Biotechnol. 2021, 39, 442–450. [Google Scholar] [CrossRef] [PubMed]
  182. EFSA Panel on Biological Hazards (BIOHAZ); Koutsoumanis, K.; Allende, A.; Álvarez-Ordóñez, A.; Bolton, D.; Bover-Cid, S.; Chemaly, M.; Davies, R.; De Cesare, A.; Herman, L.; et al. Role Played by the Environment in the Emergence and Spread of Antimicrobial Resistance (AMR) through the Food Chain. EFSA J. 2021, 19, e06651. [Google Scholar] [CrossRef] [PubMed]
  183. Walsh, A.M.; Macori, G.; Kilcawley, K.N.; Cotter, P.D. Meta-Analysis of Cheese Microbiomes Highlights Contributions to Multiple Aspects of Quality. Nat. Food 2020, 1, 500–510. [Google Scholar] [CrossRef]
  184. Devirgiliis, C.; Barile, S.; Perozzi, G. Antibiotic Resistance Determinants in the Interplay between Food and Gut Microbiota. Genes Nutr. 2011, 6, 275–284. [Google Scholar] [CrossRef] [Green Version]
  185. Tan, G.; Hu, M.; Li, X.; Pan, Z.; Li, M.; Li, L.; Zheng, Z.; Yang, M. Metagenomics Reveals the Diversity and Taxonomy of Antibiotic Resistance Genes in Sufu Bacterial Communities. Food Control 2021, 121, 107641. [Google Scholar] [CrossRef]
  186. Song, Q.; Wang, B.; Han, Y.; Zhou, Z. Metagenomics Reveals the Diversity and Taxonomy of Carbohydrate-Active Enzymes and Antibiotic Resistance Genes in Suancai Bacterial Communities. Genes 2022, 13, 773. [Google Scholar] [CrossRef]
  187. Leech, J.; Cabrera-Rubio, R.; Walsh, A.M.; Macori, G.; Walsh, C.J.; Barton, W.; Finnegan, L.; Crispie, F.; O’Sullivan, O.; Claesson, M.J.; et al. Fermented-Food Metagenomics Reveals Substrate-Associated Differences in Taxonomy and Health-Associated and Antibiotic Resistance Determinants. mSystems 2020, 5, e00522-20. [Google Scholar] [CrossRef]
  188. McArthur, A.G.; Waglechner, N.; Nizam, F.; Yan, A.; Azad, M.A.; Baylay, A.J.; Bhullar, K.; Canova, M.J.; De Pascale, G.; Ejim, L.; et al. The Comprehensive Antibiotic Resistance Database. Antimicrob. Agents Chemother. 2013, 57, 3348–3357. [Google Scholar] [CrossRef] [Green Version]
  189. Florensa, A.F.; Kaas, R.S.; Clausen, P.T.L.C.; Aytan-Aktug, D.; Aarestrup, F.M. ResFinder—An Open Online Resource for Identification of Antimicrobial Resistance Genes in next-Generation Sequencing Data and Prediction of Phenotypes from Genotypes. Microb. Genom. 2022, 8, 000748. [Google Scholar] [CrossRef] [PubMed]
  190. Walsh, A.M.; Crispie, F.; Kilcawley, K.; O’Sullivan, O.; O’Sullivan, M.G.; Claesson, M.J.; Cotter, P.D. Microbial Succession and Flavor Production in the Fermented Dairy Beverage Kefir. mSystems 2016, 1, e00052-16. [Google Scholar] [CrossRef] [PubMed]
  191. Chen, G.; Chen, C.; Lei, Z. Meta-Omics Insights in the Microbial Community Profiling and Functional Characterization of Fermented Foods. Trends Food Sci. Technol. 2017, 65, 23–31. [Google Scholar] [CrossRef]
  192. Dimidi, E.; Cox, S.R.; Rossi, M.; Whelan, K. Fermented Foods: Definitions and Characteristics, Impact on the Gut Microbiota and Effects on Gastrointestinal Health and Disease. Nutrients 2019, 11, 1806. [Google Scholar] [CrossRef] [Green Version]
  193. EFSA Panel on Dietetic Products, Nutrition and Allergies (NDA). Scientific Opinion on the Substantiation of Health Claims Related to Live Yoghurt Cultures and Improved Lactose Digestion (ID 1143, 2976) Pursuant to Article 13(1) of Regulation (EC) No 1924/2006. EFSA J. 2010, 8, 1763. [Google Scholar] [CrossRef]
  194. De Filippis, F.; Pasolli, E.; Ercolini, D. The Food-Gut Axis: Lactic Acid Bacteria and Their Link to Food, the Gut Microbiome and Human Health. FEMS Microbiol. Rev. 2020, 44, 454–489. [Google Scholar] [CrossRef]
  195. Kok, C.R.; Hutkins, R. Yogurt and Other Fermented Foods as Sources of Health-Promoting Bacteria. Nutr. Rev. 2018, 76, 4–15. [Google Scholar] [CrossRef] [Green Version]
  196. Aslam, H.; Green, J.; Jacka, F.N.; Collier, F.; Berk, M.; Pasco, J.; Dawson, S.L. Fermented Foods, the Gut and Mental Health: A Mechanistic Overview with Implications for Depression and Anxiety. Nutr. Neurosci. 2020, 23, 659–671. [Google Scholar] [CrossRef]
  197. Wang, D.H.; Yang, Y.; Wang, Z.; Lawrence, P.; Worobo, R.W.; Brenna, J.T. High Levels of Branched Chain Fatty Acids in Nātto and Other Asian Fermented Foods. Food Chem. 2019, 286, 428–433. [Google Scholar] [CrossRef]
  198. Hati, S.; Patel, M.; Mishra, B.K.; Das, S. Short-Chain Fatty Acid and Vitamin Production Potentials of Lactobacillus Isolated from Fermented Foods of Khasi Tribes, Meghalaya, India. Ann. Microbiol. 2019, 69, 1191–1199. [Google Scholar] [CrossRef]
  199. Barbara, G.; Feinle-Bisset, C.; Ghoshal, U.C.; Santos, J.; Vanner, S.J.; Vergnolle, N.; Zoetendal, E.G.; Quigley, E.M. The Intestinal Microenvironment and Functional Gastrointestinal Disorders. Gastroenterology 2016, 150, 1305–1318.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Harper, A.; Naghibi, M.M.; Garcha, D. The Role of Bacteria, Probiotics and Diet in Irritable Bowel Syndrome. Foods 2018, 7, 13. [Google Scholar] [CrossRef] [PubMed]
  201. Nayfach, S.; Pollard, K.S. Toward Accurate and Quantitative Comparative Metagenomics. Cell 2016, 166, 1103–1116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Lee, S.H.; Whon, T.W.; Roh, S.W.; Jeon, C.O. Unraveling Microbial Fermentation Features in Kimchi: From Classical to Meta-Omics Approaches. Appl. Microbiol. Biotechnol. 2020, 104, 7731–7744. [Google Scholar] [CrossRef]
  203. Blasche, S.; Kim, Y.; Mars, R.A.T.; Machado, D.; Maansson, M.; Kafkia, E.; Milanese, A.; Zeller, G.; Teusink, B.; Nielsen, J.; et al. Metabolic Cooperation and Spatiotemporal Niche Partitioning in a Kefir Microbial Community. Nat. Microbiol. 2021, 6, 196–208. [Google Scholar] [CrossRef]
  204. Kamilari, E.; Tomazou, M.; Antoniades, A.; Tsaltas, D. High Throughput Sequencing Technologies as a New Toolbox for Deep Analysis, Characterization and Potentially Authentication of Protection Designation of Origin Cheeses? Int. J. Food Sci. 2019, 2019, e5837301. [Google Scholar] [CrossRef] [Green Version]
  205. O’Sullivan, D.J.; Cotter, P.D.; O’Sullivan, O.; Giblin, L.; McSweeney, P.L.H.; Sheehan, J.J. Temporal and Spatial Differences in Microbial Composition during the Manufacture of a Continental-Type Cheese. Appl. Environ. Microbiol. 2015, 81, 2525–2533. [Google Scholar] [CrossRef] [Green Version]
  206. Pierce, E.C.; Morin, M.; Little, J.C.; Liu, R.B.; Tannous, J.; Keller, N.P.; Pogliano, K.; Wolfe, B.E.; Sanchez, L.M.; Dutton, R.J. Bacterial–Fungal Interactions Revealed by Genome-Wide Analysis of Bacterial Mutant Fitness. Nat. Microbiol. 2021, 6, 87–102. [Google Scholar] [CrossRef]
  207. Wolfe, B.E.; Button, J.E.; Santarelli, M.; Dutton, R.J. Cheese Rind Communities Provide Tractable Systems for In Situ and In Vitro Studies of Microbial Diversity. Cell 2014, 158, 422–433. [Google Scholar] [CrossRef] [Green Version]
  208. Paillet, T.; Dugat-Bony, E. Bacteriophage Ecology of Fermented Foods: Anything New under the Sun? Curr. Opin. Food Sci. 2021, 40, 102–111. [Google Scholar] [CrossRef]
  209. Roux, S.; Matthijnssens, J.; Dutilh, B.E. Metagenomics in Virology. Encycl. Virol. 2021, 1, 133–140. [Google Scholar] [CrossRef]
  210. Tamang, J.P.; Das, S.; Kharnaior, P.; Pariyar, P.; Thapa, N.; Jo, S.-W.; Yim, E.-J.; Shin, D.-H. Shotgun Metagenomics of Cheonggukjang, a Fermented Soybean Food of Korea: Community Structure, Predictive Functionalities and Amino Acids Profile. Food Res. Int. 2022, 151, 110904. [Google Scholar] [CrossRef] [PubMed]
  211. Kumar, J.; Sharma, N.; Kaushal, G.; Samurailatpam, S.; Sahoo, D.; Rai, A.K.; Singh, S.P. Metagenomic Insights Into the Taxonomic and Functional Features of Kinema, a Traditional Fermented Soybean Product of Sikkim Himalaya. Front. Microbiol. 2019, 10, 1744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Ibrahim, F.; Oppelt, J.; Maragkakis, M.; Mourelatos, Z. TERA-Seq: True End-to-End Sequencing of Native RNA Molecules for Transcriptome Characterization. Nucleic Acids Res. 2021, 49, e115. [Google Scholar] [CrossRef] [PubMed]
  213. Greninger, A.L. A Decade of RNA Virus Metagenomics Is (Not) Enough. Virus Res. 2018, 244, 218–229. [Google Scholar] [CrossRef] [PubMed]
  214. Van Reckem, E.; De Vuyst, L.; Weckx, S.; Leroy, F. Next-Generation Sequencing to Enhance the Taxonomic Resolution of the Microbiological Analysis of Meat and Meat-Derived Products. Curr. Opin. Food Sci. 2021, 37, 58–65. [Google Scholar] [CrossRef]
  215. Suminda, G.G.D.; Bhandari, S.; Won, Y.; Goutam, U.; Kanth Pulicherla, K.; Son, Y.-O.; Ghosh, M. High-Throughput Sequencing Technologies in the Detection of Livestock Pathogens, Diagnosis, and Zoonotic Surveillance. Comput. Struct. Biotechnol. J. 2022, 20, 5378–5392. [Google Scholar] [CrossRef] [PubMed]
  216. Gołębiewski, M.; Tretyn, A. Generating Amplicon Reads for Microbial Community Assessment with Next-Generation Sequencing. J. Appl. Microbiol. 2020, 128, 330–354. [Google Scholar] [CrossRef] [Green Version]
  217. Berman, H.; McLaren, M.; Callahan, B. Understanding and Interpreting Community Sequencing Measurements of the Vaginal Microbiome. BJOG 2020, 127, 139–146. [Google Scholar] [CrossRef]
  218. Weinroth, M.D.; Belk, A.D.; Dean, C.; Noyes, N.; Dittoe, D.K.; Rothrock, M.J., Jr.; Ricke, S.C.; Myer, P.R.; Henniger, M.T.; Ramírez, G.A.; et al. Considerations and Best Practices in Animal Science 16S Ribosomal RNA Gene Sequencing Microbiome Studies. J. Anim. Sci. 2022, 100, skab346. [Google Scholar] [CrossRef] [PubMed]
  219. Delbeke, H.; Younas, S.; Casteels, I.; Joossens, M. Current Knowledge on the Human Eye Microbiome: A Systematic Review of Available Amplicon and Metagenomic Sequencing Data. Acta Ophthalmol. 2021, 99, 16–25. [Google Scholar] [CrossRef] [PubMed]
  220. Wensel, C.R.; Pluznick, J.L.; Salzberg, S.L.; Sears, C.L. Next-Generation Sequencing: Insights to Advance Clinical Investigations of the Microbiome. J. Clin. Investig. 2022, 132, e154944. [Google Scholar] [CrossRef] [PubMed]
  221. Joseph, T.A.; Pe’er, I. An Introduction to Whole-Metagenome Shotgun Sequencing Studies. In Deep Sequencing Data Analysis; Shomron, N., Ed.; Methods in Molecular Biology; Springer: New York, NY, USA, 2021; pp. 107–122. ISBN 978-1-07-161103-6. [Google Scholar]
  222. Casertano, M.; Fogliano, V.; Ercolini, D. Psychobiotics, Gut Microbiota and Fermented Foods Can Help Preserving Mental Health. Food Res. Int. 2022, 152, 110892. [Google Scholar] [CrossRef] [PubMed]
  223. Van de Wouw, M.; Walsh, A.M.; Crispie, F.; van Leuven, L.; Lyte, J.M.; Boehme, M.; Clarke, G.; Dinan, T.G.; Cotter, P.D.; Cryan, J.F. Distinct Actions of the Fermented Beverage Kefir on Host Behaviour, Immunity and Microbiome Gut-Brain Modules in the Mouse. Microbiome 2020, 8, 67. [Google Scholar] [CrossRef]
  224. Dai, S.; Pan, M.; El-Nezami, H.S.; Wan, J.M.F.; Wang, M.F.; Habimana, O.; Lee, J.C.Y.; Louie, J.C.Y.; Shah, N.P. Effects of Lactic Acid Bacteria-Fermented Soymilk on Isoflavone Metabolites and Short-Chain Fatty Acids Excretion and Their Modulating Effects on Gut Microbiota. J. Food Sci. 2019, 84, 1854–1863. [Google Scholar] [CrossRef]
  225. Shimizu, H.; Masujima, Y.; Ushiroda, C.; Mizushima, R.; Taira, S.; Ohue-Kitano, R.; Kimura, I. Dietary Short-Chain Fatty Acid Intake Improves the Hepatic Metabolic Condition via FFAR3. Sci. Rep. 2019, 9, 16574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Vital, M.; Howe, A.; Bergeron, N.; Krauss, R.M.; Jansson, J.K.; Tiedje, J.M. Metagenomic Insights into the Degradation of Resistant Starch by Human Gut Microbiota. Appl. Environ. Microbiol. 2018, 84, e01562-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Zhao, L.; Zhang, F.; Ding, X.; Wu, G.; Lam, Y.Y.; Wang, X.; Fu, H.; Xue, X.; Lu, C.; Ma, J.; et al. Gut Bacteria Selectively Promoted by Dietary Fibers Alleviate Type 2 Diabetes. Science 2018, 359, 1151–1156. [Google Scholar] [CrossRef] [Green Version]
  228. Figueroa-Hernández, C.; Mota-Gutierrez, J.; Ferrocino, I.; Hernández-Estrada, Z.J.; González-Ríos, O.; Cocolin, L.; Suárez-Quiroz, M.L. The Challenges and Perspectives of the Selection of Starter Cultures for Fermented Cocoa Beans. Int. J. Food Microbiol. 2019, 301, 41–50. [Google Scholar] [CrossRef]
  229. Ianni, A.; Di Domenico, M.; Bennato, F.; Peserico, A.; Martino, C.; Rinaldi, A.; Candeloro, L.; Grotta, L.; Cammà, C.; Pomilio, F.; et al. Metagenomic and Volatile Profiles of Ripened Cheese Obtained from Dairy Ewes Fed a Dietary Hemp Seed Supplementation. J. Dairy Sci. 2020, 103, 5882–5892. [Google Scholar] [CrossRef]
  230. Landis, E.A.; Oliverio, A.M.; McKenney, E.A.; Nichols, L.M.; Kfoury, N.; Biango-Daniels, M.; Shell, L.K.; Madden, A.A.; Shapiro, L.; Sakunala, S.; et al. The Diversity and Function of Sourdough Starter Microbiomes. Elife 2021, 10, e61644. [Google Scholar] [CrossRef] [PubMed]
  231. Milani, C.; Fontana, F.; Alessandri, G.; Mancabelli, L.; Lugli, G.A.; Longhi, G.; Anzalone, R.; Viappiani, A.; Duranti, S.; Turroni, F.; et al. Ecology of Lactobacilli Present in Italian Cheeses Produced from Raw Milk. Appl. Environ. Microbiol. 2020, 86, e00139-20. [Google Scholar] [CrossRef] [PubMed]
  232. Pacheco-Montealegre, M.E.; Dávila-Mora, L.L.; Botero-Rute, L.M.; Reyes, A.; Caro-Quintero, A. Fine Resolution Analysis of Microbial Communities Provides Insights Into the Variability of Cocoa Bean Fermentation. Front. Microbiol. 2020, 11, 650. [Google Scholar] [CrossRef] [Green Version]
  233. Casey, E.; McDonnell, B.; White, K.; Stamou, P.; Crowley, T.; O’Neill, I.; Lavelle, K.; Hayes, S.; Lugli, G.A.; Arboleya, S.; et al. Needle in a Whey-Stack: PhRACS as a Discovery Tool for Unknown Phage-Host Combinations. mBio 2022, 13, e03334-21. [Google Scholar] [CrossRef]
  234. Mahony, J.; van Sinderen, D. Virome Studies of Food Production Systems: Time for ‘Farm to Fork’ Analyses. Curr. Opin. Biotechnol. 2022, 73, 22–27. [Google Scholar] [CrossRef] [PubMed]
  235. Muhammed, M.K.; Kot, W.; Neve, H.; Mahony, J.; Castro-Mejía, J.L.; Krych, L.; Hansen, L.H.; Nielsen, D.S.; Sørensen, S.J.; Heller, K.J.; et al. Metagenomic Analysis of Dairy Bacteriophages: Extraction Method and Pilot Study on Whey Samples Derived from Using Undefined and Defined Mesophilic Starter Cultures. Appl. Environ. Microbiol. 2017, 83, e00888-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Johansen, P.; Vindeløv, J.; Arneborg, N.; Brockmann, E. Development of Quantitative PCR and Metagenomics-Based Approaches for Strain Quantification of a Defined Mixed-Strain Starter Culture. Syst. Appl. Microbiol. 2014, 37, 186–193. [Google Scholar] [CrossRef]
  237. Hussain, B.; Chen, J.-S.; Hsu, B.-M.; Chu, I.-T.; Koner, S.; Chen, T.-H.; Rathod, J.; Chan, M.W.Y. Deciphering Bacterial Community Structure, Functional Prediction and Food Safety Assessment in Fermented Fruits Using Next-Generation 16S RRNA Amplicon Sequencing. Microorganisms 2021, 9, 1574. [Google Scholar] [CrossRef]
  238. Walsh, A.M.; Crispie, F.; Daari, K.; O’Sullivan, O.; Martin, J.C.; Arthur, C.T.; Claesson, M.J.; Scott, K.P.; Cotter, P.D. Strain-Level Metagenomic Analysis of the Fermented Dairy Beverage Nunu Highlights Potential Food Safety Risks. Appl. Environ. Microbiol. 2017, 83, e01144-17. [Google Scholar] [CrossRef] [Green Version]
  239. Illeghems, K.; Weckx, S.; De Vuyst, L. Applying Meta-Pathway Analyses through Metagenomics to Identify the Functional Properties of the Major Bacterial Communities of a Single Spontaneous Cocoa Bean Fermentation Process Sample. Food Microbiol. 2015, 50, 54–63. [Google Scholar] [CrossRef]
  240. O’Connor, P.M.; Kuniyoshi, T.M.; Oliveira, R.P.; Hill, C.; Ross, R.P.; Cotter, P.D. Antimicrobials for Food and Feed; a Bacteriocin Perspective. Curr. Opin. Biotechnol. 2020, 61, 160–167. [Google Scholar] [CrossRef] [PubMed]
  241. Suárez, N.; Weckx, S.; Minahk, C.; Hebert, E.M.; Saavedra, L. Metagenomics-Based Approach for Studying and Selecting Bioprotective Strains from the Bacterial Community of Artisanal Cheeses. Int. J. Food Microbiol. 2020, 335, 108894. [Google Scholar] [CrossRef] [PubMed]
  242. Li, L.-G.; Huang, Q.; Yin, X.; Zhang, T. Source Tracking of Antibiotic Resistance Genes in the Environment—Challenges, Progress, and Prospects. Water Res. 2020, 185, 116127. [Google Scholar] [CrossRef] [PubMed]
  243. Yasir, M.; Al-Zahrani, I.A.; Bibi, F.; Abd El Ghany, M.; Azhar, E.I. New Insights of Bacterial Communities in Fermented Vegetables from Shotgun Metagenomics and Identification of Antibiotic Resistance Genes and Probiotic Bacteria. Food Res. Int. 2022, 157, 111190. [Google Scholar] [CrossRef]
  244. Haiminen, N.; Edlund, S.; Chambliss, D.; Kunitomi, M.; Weimer, B.C.; Ganesan, B.; Baker, R.; Markwell, P.; Davis, M.; Huang, B.C.; et al. Food Authentication from Shotgun Sequencing Reads with an Application on High Protein Powders. Npj Sci. Food 2019, 3, 24. [Google Scholar] [CrossRef] [Green Version]
  245. Jiang, M.; Xu, S.-F.; Tang, T.-S.; Miao, L.; Luo, B.-Z.; Ni, Y.; Kong, F.-D.; Liu, C. Development and Evaluation of a Meat Mitochondrial Metagenomic (3MG) Method for Composition Determination of Meat from Fifteen Mammalian and Avian Species. BMC Genom. 2022, 23, 36. [Google Scholar] [CrossRef]
  246. Kobus, R.; Abuín, J.M.; Müller, A.; Hellmann, S.L.; Pichel, J.C.; Pena, T.F.; Hildebrandt, A.; Hankeln, T.; Schmidt, B. A Big Data Approach to Metagenomics for All-Food-Sequencing. BMC Bioinform. 2020, 21, 102. [Google Scholar] [CrossRef]
  247. Voorhuijzen-Harink, M.M.; Hagelaar, R.; van Dijk, J.P.; Prins, T.W.; Kok, E.J.; Staats, M. Toward On-Site Food Authentication Using Nanopore Sequencing. Food Chem. X 2019, 2, 100035. [Google Scholar] [CrossRef]
  248. Bokulich, N.A.; Thorngate, J.H.; Richardson, P.M.; Mills, D.A. Microbial Biogeography of Wine Grapes Is Conditioned by Cultivar, Vintage, and Climate. Proc. Natl. Acad. Sci. USA 2014, 111, E139–E148. [Google Scholar] [CrossRef] [Green Version]
  249. Gul, O.; Atalar, I.; Mortas, M.; Dervisoglu, M. Rheological, Textural, Colour and Sensorial Properties of Kefir Produced with Buffalo Milk Using Kefir Grains and Starter Culture: A Comparison with Cows’ Milk Kefir. Int. J. Dairy Technol. 2018, 71, 73–80. [Google Scholar] [CrossRef]
  250. Vermote, L.; Verce, M.; De Vuyst, L.; Weckx, S. Amplicon and Shotgun Metagenomic Sequencing Indicates That Microbial Ecosystems Present in Cheese Brines Reflect Environmental Inoculation during the Cheese Production Process. Int. Dairy J. 2018, 87, 44–53. [Google Scholar] [CrossRef]
  251. Yang, X.; Hu, W.; Xiu, Z.; Jiang, A.; Yang, X.; Saren, G.; Ji, Y.; Guan, Y.; Feng, K. Microbial Community Dynamics and Metabolome Changes During Spontaneous Fermentation of Northeast Sauerkraut From Different Households. Front. Microbiol. 2020, 11, 1878. [Google Scholar] [CrossRef] [PubMed]
  252. Hananiah, N.; Rahim, A.A. The Application of Hurdle Technology in Extending the Shelf Life and Improving the Quality of Fermented Freshwater Fish (Pekasam): A Review. MJoSHT 2022, 8, 44–54. [Google Scholar] [CrossRef]
  253. Kazou, M.; Grafakou, A.; Tsakalidou, E.; Georgalaki, M. Zooming Into the Microbiota of Home-Made and Industrial Kefir Produced in Greece Using Classical Microbiological and Amplicon-Based Metagenomics Analyses. Front. Microbiol. 2021, 12, 621069. [Google Scholar] [CrossRef]
  254. Katz, L.; Chen, Y.Y.; Gonzalez, R.; Peterson, T.C.; Zhao, H.; Baltz, R.H. Synthetic Biology Advances and Applications in the Biotechnology Industry: A Perspective. J. Ind. Microbiol. Biotechnol. 2018, 45, 449–461. [Google Scholar] [CrossRef]
  255. Son, J.; Jeong, K. Recent Advances in Synthetic Biology for the Engineering of Lactic Acid Bacteria. Biotechnol. Bioprocess. Eng. 2020, 25, 962–973. [Google Scholar] [CrossRef]
  256. Sambyal, K.; Singh, R.V. Production Aspects of Testosterone by Microbial Biotransformation and Future Prospects. Steroids 2020, 159, 108651. [Google Scholar] [CrossRef]
  257. Sharma, A.; Sharma, P.; Singh, J.; Singh, S.; Nain, L. Prospecting the Potential of Agroresidues as Substrate for Microbial Flavor Production. Front. Sustain. Food Syst. 2020, 4, 18. [Google Scholar] [CrossRef] [Green Version]
  258. Amicarelli, V.; Lagioia, G.; Bux, C. Global Warming Potential of Food Waste through the Life Cycle Assessment: An Analytical Review. Environ. Impact. Assess Rev. 2021, 91, 106677. [Google Scholar] [CrossRef]
  259. Wesana, J.; Gellynck, X.; Dora, M.K.; Pearce, D.; De Steur, H. Measuring Food and Nutritional Losses through Value Stream Mapping along the Dairy Value Chain in Uganda. Resour. Conserv. Recycl. 2019, 150, 104416. [Google Scholar] [CrossRef]
  260. Calvete-Torre, I.; Sabater, C.; Antón, M.J.; Moreno, F.J.; Riestra, S.; Margolles, A.; Ruiz, L. Prebiotic Potential of Apple Pomace and Pectins from Different Apple Varieties: Modulatory Effects on Key Target Commensal Microbial Populations. Food Hydrocoll. 2022, 133, 107958. [Google Scholar] [CrossRef]
  261. Hyun Chung, T.; Ranjan Dhar, B. A Multi-Perspective Review on Microbial Electrochemical Technologies for Food Waste Valorization. Bioresour. Technol. 2021, 342, 125950. [Google Scholar] [CrossRef] [PubMed]
  262. Mehmood, T.; Nadeem, F.; Bilal, M.; Iqbal, H.M.N. 8—Recent Trends on the Food Wastes Valorization to Value-Added Commodities. In Advanced Technology for the Conversion of Waste into Fuels and Chemicals; Khan, A., Jawaid, M., Pizzi, A., Azum, N., Asiri, A., Isa, I., Eds.; Woodhead Publishing: Sawston, UK, 2021; pp. 171–196. ISBN 978-0-12-823139-5. [Google Scholar]
  263. Sabater, C.; Calvete-Torre, I.; Villamiel, M.; Moreno, F.J.; Margolles, A.; Ruiz, L. Vegetable Waste and By-Products to Feed a Healthy Gut Microbiota: Current Evidence, Machine Learning and Computational Tools to Design Novel Microbiome-Targeted Foods. Trends Food Sci. Technol. 2021, 118, 399–417. [Google Scholar] [CrossRef]
  264. Socas-Rodríguez, B.; Álvarez-Rivera, G.; Valdés, A.; Ibáñez, E.; Cifuentes, A. Food By-Products and Food Wastes: Are They Safe Enough for Their Valorization? Trends Food Sci. Technol. 2021, 114, 133–147. [Google Scholar] [CrossRef]
  265. Talan, A.; Tiwari, B.; Yadav, B.; Tyagi, R.D.; Wong, J.W.C.; Drogui, P. Food Waste Valorization: Energy Production Using Novel Integrated Systems. Bioresour. Technol. 2021, 322, 124538. [Google Scholar] [CrossRef]
  266. Iquebal, M.A.; Jagannadham, J.; Jaiswal, S.; Prabha, R.; Rai, A.; Kumar, D. Potential Use of Microbial Community Genomes in Various Dimensions of Agriculture Productivity and Its Management: A Review. Front. Microbiol. 2022, 13, 708335. [Google Scholar] [CrossRef]
  267. Eckstrom, K.; Barlow, J.W. Resistome Metagenomics from Plate to Farm: The Resistome and Microbial Composition during Food Waste Feeding and Composting on a Vermont Poultry Farm. PLoS ONE 2019, 14, e0219807. [Google Scholar] [CrossRef] [Green Version]
  268. Bianco, A.; Budroni, M.; Zara, S.; Mannazzu, I.; Fancello, F.; Zara, G. The Role of Microorganisms on Biotransformation of Brewers’ Spent Grain. Appl. Microbiol. Biotechnol. 2020, 104, 8661–8678. [Google Scholar] [CrossRef]
  269. Crognale, S.; Braguglia, C.M.; Gallipoli, A.; Gianico, A.; Rossetti, S.; Montecchio, D. Direct Conversion of Food Waste Extract into Caproate: Metagenomics Assessment of Chain Elongation Process. Microorganisms 2021, 9, 327. [Google Scholar] [CrossRef]
  270. Zhang, L.; Loh, K.-C.; Kuroki, A.; Dai, Y.; Tong, Y.W. Microbial Biodiesel Production from Industrial Organic Wastes by Oleaginous Microorganisms: Current Status and Prospects. J. Hazard. Mater. 2021, 402, 123543. [Google Scholar] [CrossRef]
  271. Javourez, U.; O’Donohue, M.; Hamelin, L. Waste-to-Nutrition: A Review of Current and Emerging Conversion Pathways. Biotechnol. Adv. 2021, 53, 107857. [Google Scholar] [CrossRef] [PubMed]
  272. Lv, X.; Wu, Y.; Gong, M.; Deng, J.; Gu, Y.; Liu, Y.; Li, J.; Du, G.; Ledesma-Amaro, R.; Liu, L.; et al. Synthetic Biology for Future Food: Research Progress and Future Directions. Future Foods 2021, 3, 100025. [Google Scholar] [CrossRef]
  273. Galimberti, A.; Bruno, A.; Agostinetto, G.; Casiraghi, M.; Guzzetti, L.; Labra, M. Fermented Food Products in the Era of Globalization: Tradition Meets Biotechnology Innovations. Curr. Opin. Biotechnol. 2021, 70, 36–41. [Google Scholar] [CrossRef]
  274. Branduardi, P. Closing the Loop: The Power of Microbial Biotransformations from Traditional Bioprocesses to Biorefineries, and Beyond. Microb. Biotechnol. 2021, 14, 68–73. [Google Scholar] [CrossRef] [PubMed]
  275. Ubando, A.T.; Felix, C.B.; Chen, W.-H. Biorefineries in Circular Bioeconomy: A Comprehensive Review. Bioresour. Technol. 2020, 299, 122585. [Google Scholar] [CrossRef] [PubMed]
  276. Chavan, S.; Yadav, B.; Atmakuri, A.; Tyagi, R.D.; Wong, J.W.C.; Drogui, P. Bioconversion of Organic Wastes into Value-Added Products: A Review. Bioresour. Technol. 2022, 344, 126398. [Google Scholar] [CrossRef]
  277. Jayasekara, S.; Dissanayake, L.; Jayakody, L.N. Opportunities in the Microbial Valorization of Sugar Industrial Organic Waste to Biodegradable Smart Food Packaging Materials. Int. J. Food Microbiol. 2022, 377, 109785. [Google Scholar] [CrossRef]
Figure 1. Overview of library preparation steps for amplicon and shotgun sequencing on (a) short read platforms such as Illumina, and (b) long read platforms such as PacBio and ONT [72,82]. This figure was created with BioRender.com.
Figure 1. Overview of library preparation steps for amplicon and shotgun sequencing on (a) short read platforms such as Illumina, and (b) long read platforms such as PacBio and ONT [72,82]. This figure was created with BioRender.com.
Foods 11 03297 g001aFoods 11 03297 g001b
Figure 2. Overview of the potential value-added products that can be obtained through the combined application of metagenomics, synthetic biology and microbial biotransformation, enabling the establishment of circular bioeconomies [276,277]. In this process, metagenomics can be applied to understand the functional roles within the microbial communities to allow their application in industry through microbial biotransformation. This figure was created with BioRender.com.
Figure 2. Overview of the potential value-added products that can be obtained through the combined application of metagenomics, synthetic biology and microbial biotransformation, enabling the establishment of circular bioeconomies [276,277]. In this process, metagenomics can be applied to understand the functional roles within the microbial communities to allow their application in industry through microbial biotransformation. This figure was created with BioRender.com.
Foods 11 03297 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Srinivas, M.; O’Sullivan, O.; Cotter, P.D.; Sinderen, D.v.; Kenny, J.G. The Application of Metagenomics to Study Microbial Communities and Develop Desirable Traits in Fermented Foods. Foods 2022, 11, 3297. https://doi.org/10.3390/foods11203297

AMA Style

Srinivas M, O’Sullivan O, Cotter PD, Sinderen Dv, Kenny JG. The Application of Metagenomics to Study Microbial Communities and Develop Desirable Traits in Fermented Foods. Foods. 2022; 11(20):3297. https://doi.org/10.3390/foods11203297

Chicago/Turabian Style

Srinivas, Meghana, Orla O’Sullivan, Paul D. Cotter, Douwe van Sinderen, and John G. Kenny. 2022. "The Application of Metagenomics to Study Microbial Communities and Develop Desirable Traits in Fermented Foods" Foods 11, no. 20: 3297. https://doi.org/10.3390/foods11203297

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop