To date, the vast majority of prokaryotic organisms escapes detailed characterization because they cannot be isolated in axenic cultures. These organisms are referred to as microbial dark matter. Targeted labelling and sorting of these microorganisms pave the way for single-cell, enrichment, or cultivation approaches. In this review, we describe an array of different methods ranging from labeling-free to specific labelling techniques. In addition, different cell sorting methods and their combinations with targeting strategies are summarized and downstream applications like sequencing and cultivation are reviewed. Recent advances, challenges, and limitations of the particular methods are discussed with respect to cell viability, genome integrity as well as throughput, in order to help researchers select the most suitable methods for their specific research questions.

Microorganisms harbor a huge potential for biotechnological applications, e.g., the production of novel antibiotics, the conservation and storage of bioenergy, or the remediation of hazardous compounds [Mullis et al., 2019; Jeyavani et al., 2021; Wan et al., 2021]. Since the late 1980s, studies of prokaryotic communities based on 16S rRNA gene sequences have shown a vast diversity of microorganisms in the environment [Cho, 2021]. However, the majority of prokaryotic species has yet to be characterized and is therefore referred to as microbial dark matter (MDM) [Lok, 2015; Kaster and Sobol, 2020; Jiao et al., 2021]. Despite breakthroughs in cultivation techniques [Berdy et al., 2017; Zoheir et al., 2022], attempts to grow the majority of microorganisms and to generate sufficient biomass for further analysis have mostly failed (also known as the great plate count anomaly) [Staley and Konopka, 1985]. This is why researchers still primarily rely on omics methods, such as meta- and single-cell omics to study MDM [Kaster and Sobol, 2020; Jiao et al., 2021; Kumar et al., 2021; Münch et al., 2023].

Genomes derived from metagenomics data – so-called metagenome-assembled genomes (MAGs) – are useful for understanding microbial community compositions and metabolic potentials. Nevertheless, creating reference-quality genomes at the species level can be difficult and expensive with this approach. The quality of genome reconstruction is largely dependent on sequencing coverage and covariance-based binning [Amarasinghe et al., 2020; Dam et al., 2020; Kaster and Sobol, 2020], which renders the characterization of low-abundance microorganisms in a habitat a challenge, particularly in ecosystems with great diversity [Bickhart et al., 2022]. Despite advances in sequencing technologies (e.g., long-read sequencing) [Amarasinghe et al., 2020] and binning algorithms [Arisdakessian et al., 2021; Nissen et al., 2021; Zhang and Zhang, 2021], MAGs still are often consensus genomes derived from cells with high homologous regions. In addition, mobile genetic elements such as plasmids cannot be assigned to particular MAGs [Vollmers et al., 2022].

Single-cell genomics (SCG) on the other hand is the separation of single cells from a community prior to analysis, followed by sequencing and data analyses, which results in so-called single amplified genomes (SAGs) [Kaster and Sobol, 2020; Jiao et al., 2021]. However, obtaining SAGs can be difficult and expensive, too. This is especially true for minority community members that are statistically rarely sorted. In addition, cells might already lyse during the sorting process or do not lyse at all when applying standard lysis conditions [Rinke et al., 2014; Dam et al., 2020; Wiegand et al., 2021]. Moreover, SCG requires amplification of DNA since a single cell only contains femtograms of nucleic acids, which is too little to be directly used in sequencing. Multiple displacement amplification (MDA) is the most common method used; however, it suffers from random biases as well as biases against high GC regions and chimera formation, which often results in incomplete genomes [Parras-Moltó et al., 2018; Tu et al., 2021; Lawrence et al., 2022; Sobol and Kaster, 2023]. This is why standard SCG and metagenomics can become very expensive in regard to time and cost when one is interested in a specific organism with a certain taxonomy or metabolism, and especially when looking for minority members of microbial communities.

That is why other methods have been developed in the recent past. “Mini-metagenomics” uses small pools of typically 5–100 sorted cells from which the DNA is then amplified via MDA and sequenced [McLean et al., 2013; Alteio et al., 2020]. While the higher amount of cells is considered to reduce amplification bias, amplification is still necessary to obtain enough DNA for sequencing. To overcome the problem of low DNA input amounts for sequencing, “midi-metagenomics” was developed. In this approach, cells are fractionated into larger pools of several million cells in regard to different properties which can be distinguished by flow cytometers (e.g., different shape and size, metabolic, or physiological markers). From these pools, DNA can be directly isolated and sequenced resulting in assemblies that are suitable for co-variance-based binning approaches, yielding MAGs of significantly higher quality compared to conventional metagenomic concepts and avoiding additional costs for amplification [Vollmers et al., 2023].

Instead of sequencing nucleic acids of single cells, one can also attempt to cultivate from a single organism [Hu et al., 2020; Täuber et al., 2020; Anggraini et al., 2022]. This technique has the advantage of having a certain microorganism already separated from its community, hence avoiding competition for nutrients or the danger of being destroyed by competitors. Currently, there are different methods available for targeted cell detection and isolation. These methods can be either rather unspecific, label-free approaches, or aim at selectively targeting cells with a specific trait, taxonomy, or metabolism using radioactive or stable isotopes, or – in the majority of cases – fluorescent labels. However, since the mere detection of cells is not the ultimate goal, labelling methods have to be compatible with downstream processes for sequencing or cultivation. The same holds true for cell sorting techniques such as flow cytometry (FC), microfluidic devices, or cell printing. In this review, we aim to describe challenges as well as recent advances in prokaryotic cell labeling and sorting allowing downstream omics as well as live recovery of cells for cell cultivation.

Cultivation-dependent and cultivation-independent processes are the two approaches mainly applied for characterizing prokaryotic species. However, it has been demonstrated frequently that communities characterized using cultivation-independent methodologies often show greater complexity compared to traditional cultivation-based methods. In general, minorities in prokaryotic communities, but also anaerobes or cells with tough cell walls are usually neglected and understudied. This also holds true for the domain of Archaea, which is predominated by extremophiles [Rampelotto, 2013; Reed et al., 2013; Kaushik et al., 2021]. Low-abundance microbes in complex communities might, however, still play important roles in biogeochemical processes (e.g., due to high enzyme affinities to certain substrates) or might have biotechnological or environmental relevance [Frias-Lopez et al., 2008; Pratscher et al., 2018]. Consequently, targeted separation and sorting approaches are needed to overcome the limitations of conventional methods to allow the enrichment or isolation of formerly unidentified or underrepresented species.

Label-Free Approaches and Detection

Label-free approaches focus on natural and intrinsic cellular features that provide useful information about a cell’s phenotype [He et al., 2019; Jayan et al., 2022]. Autofluorescence, e.g., caused by a variety of internal fluorescent biomolecules (e.g., cofactors) can be utilized to identify and categorize prokaryotic species [Ammor, 2007; Kang et al., 2020] (Fig. 1a). These compounds excite in the UV/VIS spectral range usually between 250 nm and 450 nm and emit fluorescence in the spectral range of 280 nm–540 nm [Ammor, 2007]. It was demonstrated that even different serotypes of species from reference genera (Xanthomonas, Stenotrophomonas, Burkholderia, and Pseudomonas) can show different autofluorescence signals [Tourkya et al., 2009]. However, this approach’s efficacy is constrained for environmental samples since many different species might exhibit the same autofluorescence.

Fig. 1.

Label-free detection methods. a Autofluorescence. Cells exhibiting autofluorescence are excited by light and fluorescence can be monitored via detectors or microscopy. b Fourier transform infrared spectroscopy (FT-IR). The split beam generates an interferogram that is transformed into an absorption spectrum. c Secondary ion mass spectrometry (SIMS). Bombardment of the sample with primary ions causes sputtering of secondary ions that are analyzed via mass spectrometry. d Raman spectroscopy (RS). The sample is excited via a laser and generated Raman-shifted photons are collected. For detailed explanations, we refer to the text. h, Planck’s constant; ν, frequency; RFU, relative fluorescence units; IR, Infrared radiation; m/z, mass/charge ratio; a.u., absorption units. Created with Biorender.

Fig. 1.

Label-free detection methods. a Autofluorescence. Cells exhibiting autofluorescence are excited by light and fluorescence can be monitored via detectors or microscopy. b Fourier transform infrared spectroscopy (FT-IR). The split beam generates an interferogram that is transformed into an absorption spectrum. c Secondary ion mass spectrometry (SIMS). Bombardment of the sample with primary ions causes sputtering of secondary ions that are analyzed via mass spectrometry. d Raman spectroscopy (RS). The sample is excited via a laser and generated Raman-shifted photons are collected. For detailed explanations, we refer to the text. h, Planck’s constant; ν, frequency; RFU, relative fluorescence units; IR, Infrared radiation; m/z, mass/charge ratio; a.u., absorption units. Created with Biorender.

Close modal

Fourier Transform Infrared Spectroscopy

An organsisms’ biochemical fingerprint can be identified using Fourier Transform Infrared spectroscopy (FT-IR) since infrared radiation (IR) is absorbed by various biological components (e.g., nucleic acids, proteins, lipids, or carbohydrates) within the cell [Quintelas et al., 2018; Phelan et al., 2020; Novais and Peixe, 2021]. Here, a split IR radiation beam is used to determine a sample-specific interferogram that can be analyzed against reference databases to identify molecular configurations as well as whole cells (Fig. 1b). Only a few studies have examined single microbial cells in their native environments using FT-IR [Harrison and Berry, 2017], but the technique was used, e.g., to identify 14 different strains of Enterobacter cloacae from patients of a neonatal intensive care unit and the results were consistent with results from whole genome sequencing analyses [Vogt et al., 2019]. Despite its nondestructive nature, cost-effectiveness, and speed, the technology has a number of drawbacks, including a shortage of band frequencies for distinct chemical compounds, a lack of comprehensive commercial spectral libraries and water being a strong absorber of IR radiation. So far, these drawbacks have significantly limited successful application in identifying specific cells from natural habitats so far.

Secondary Ion Mass Spectrometry

Secondary ion mass spectrometry (SIMS) provides sufficient resolution to investigate individual microbial cells. In this approach, analyte ions (secondary ions) are formed by the bombardment of the sample surface with high-energy particles (primary ions). Secondary ions are then directed into a mass spectrometer for analysis [Watrous and Dorrestein, 2011] (Fig. 1c). Mass spectrometric imaging analyses such as SIMS are most commonly used on flat surfaces (depth resolution 1–2 nm). As a result, mass spectrometric imaging 3D mapping of complex microbiological forms (e.g., aggregates or biofilms) can be difficult. Additional measures are required for successful sample preparation, demanding further stabilization or freezing, which generally results in loss of live cells [Watrous and Dorrestein, 2011; Grujcic et al., 2022]. Moreover, SIMS imaging employs a harsh and destructive ionization technique leaving cells in a nonviable state after analysis [Watrous and Dorrestein, 2011]. Another issue with this method is its difficulty in precisely selecting a specific microorganism.

A further development is the so-called NanoSIMS, which has a greater lateral resolution [Gao et al., 2016; Nuñez et al., 2018] but requires expensive instrumentation. Briefly, this method enables the generation of nanoscale maps of elemental or isotopic distribution, and high mass resolution (10–50 times greater than conventional SIMS) [Kilburn and Wacey, 2014; Oehler and Cady, 2014]. This is achieved by using Cs+ and O as primary ion beams with narrow focal planes (50–200 nm for Cs+ primary ion beam) [Gates et al., 2018]. Additionally, a double-focusing design is used in NanoSIMS for secondary ion collection and detection. Most NanoSIMS analysis efforts utilize isotope and/or rare element labeling to enable the detection of processes or species of interest inside biological samples [Nuñez et al., 2018]. NanoSIMS is now frequently used in combination with other measures such as fluorescent in situ hybridization (FISH) or bioorthogonal noncanonical amino acid tagging (BONCAT) [Grujcic et al., 2022] (discussed below). Although the method is destructive and downstream applications such as cultivation and omics approaches are not possible, its application in a variety of studies shows a valuable information gain in the detection of MDM and description of prokaryotic communities on single-cell level (reviewed in, e.g., [Orphan and House, 2009; Gao et al., 2016; Musat et al., 2016; Nuñez et al., 2018]).

Raman Spectroscopy

Raman spectroscopy (RS) is a powerful tool to identify cellular components and/or individual cells based on their specific Raman spectrum. The Raman effect describes the inelastic scattering of photons interacting with molecules [Raman and Krishnan, 1928]. Inelastic scattering occurs on two energy levels: wavelengths of inelastically scattered photons can be shifted to either higher (stokes) or lower (anti-stokes) wavelengths. The Raman shift (difference of wavelengths between excitation and Raman signal) is specific for different molecular vibrational modes and can be used as “fingerprint” to identify certain components in a sample (Fig. 1d). Raman excitation is typically based on laser light, with 532 nm and 785 nm being the most frequent laser wavelengths for biological samples due to the lower fluorescence background [Tu and Chang, 2012; Wang et al., 2021]. RS is a point-based technique that only probes a distinct spatial region of the sample. In order to provide chemoselective images of larger biological structures (e.g., in µm scale), raster scanning can be performed [Lohumi et al., 2017]. Collecting spatially resolved Raman spectra with scanning devices results in Raman “maps” harboring information about the chemical fingerprints at distinct spots but are rather time-consuming depending on the signal strength [Le et al., 2010; Stewart et al., 2012]. RS has been applied to heterogeneous prokaryotic cell populations. Kanno et al. [2021] recently classified three bacterial and three archaeal species from various phyla by using single-cell Raman microspectroscopy combined with machine learning algorithms [Kanno et al., 2021]. Song and colleagues were able to identify (and sort) different bacterial strains from the Red Sea according to their carotinoid content and moreover to classify the identified carotenoids into subgroups [Song et al., 2017]. Despite the potential of the method, the main issue remains the lack of spectral databases to identify phenotypic differences among different prokaryotes and their physiological states [Spitsyn et al., 2021]. In addition, energy acquired during signal acquisition (i.e., heat) has proved to be a major issue when aiming for genomics approaches [Su et al., 2020b]. Raman signals are also often masked by fluorescence, frequently emitted by biological molecules, especially when probed in the UV or short-wavelength visible spectral region [Le et al., 2010; Spitsyn et al., 2021; Tang et al., 2021].

To address these issues, Surface Enhanced Raman spectroscopy (SERS) has been developed [Fleischmann et al., 1974; Jeanmaire and Duyne, 1977]. Here, Raman signals can be increased by SERS-active substrates such as roughened noble-metal surfaces or noble-metal nanoparticles. In this approach, the Raman signal intensity increases by orders of magnitude when microorganisms are brought in close vicinity of the substrate [Cui et al., 2019; Weiss et al., 2019]. However, since only cell components in close proximity to the SERS-active substrate contribute to the Raman spectrum, the spatial resolution is decreased. While a “conventional” Raman spectrum represents a cross section profile of all Raman-active molecules of a particular cell, a SERS spectrum represents just a subfraction of components that are closely oriented to the surface (e.g., membrane components). Since the spatial orientation of the sample on the surface has a great impact on the obtained Raman spectrum, one has to make sure that cells are evenly adhered to the surface following the surface selection rules [Le Ru et al., 2011; Xu et al., 2011]. Studies have been performed on the analysis of pathogenic and nonpathogenic microorganisms using label-free SERS methods [Bodelón et al., 2018; Akanny et al., 2020; Liu et al., 2021]. It is noteworthy that SERS signals still strongly rely on the operational conditions related to sample preparation (i.e., nanoparticle morphology, chemical composition, concentration as well as the chemical nature of the SERS substrate) [Witkowska et al., 2019; Sportelli et al., 2022]. In addition, some metals, especially silver, show antibacterial effects and might affect microorganisms’ viability. By conducting a comparative study on nontoxic gold with toxic silver nanoparticles, the effects of nanoparticle concentration and incubation time on Lactobacillus fermentum and Escherichia coli SERS spectral variance were investigated. The results demonstrated both nanoparticle concentration and incubation time to be closely related to the toxicity of silver particles [Cui et al., 2015]. A fundamental understanding of the origin of the Raman signal and the correlation between the SERS signals from single cells and their metabolic activity is essential for reliable analysis [Sportelli et al., 2022]. An analysis of six different prokaryotic species (one archaeon, two Gram-positive bacteria, and three Gram-negative bacteria) demonstrated that the metabolic activity of the studied cells has a significant impact on the corresponding SERS signal at the single-cell level [Weiss et al., 2019]. Therefore, a large database and standardized analysis methods are required to have broader applications for this method [Cui et al., 2019; Matanfack et al., 2020].

Coherent Anti-Stokes Raman scattering (CARS) is an additional method for amplifying Raman scattered light [Volkmer et al., 2002; Camp and Cicerone, 2015]. CARS imaging was frequently used for identification of mammalian brain tissue, tumors, and lipids in vivo [Cui et al., 2022] but was also applied for imaging of Bacillus subtilis spores in water, illustrating that limitations of conventional Raman signal acquisition with respect to signal strength could be overcome using CARS [Petrov et al., 2005]. For all Raman methods, photodamage of cells is, however, a major problem, limiting the applications for downstream processing [Fu et al., 2006; Tripathi et al., 2008; Pilo-Pais et al., 2014]. For detailed information about Raman-based techniques as well as profound background information, we refer to the excellent review written by Cialla-May and colleagues [Cialla-May et al., 2022].

Labeling Approaches

The abovementioned techniques use chemico-physical properties of certain molecules of cells or cell fractions/compounds to classify cells. However, target molecules (e.g., carotenoids or cofactors) might be expressed by different strains and show the same (or similar) detection behavior. This might lead to a collection of cells or strains showing the same signals but still belong to different clades of prokaryotes with different metabolic properties. A simple example is the well known live/dead staining, where dyes staining nucleic acids (e.g., propidium-iodide, SYBR® or SYTO® dyes) allow for discrimination of living from dead cells according to the cellular viability (i.e., membrane integrity). However, numerous research questions require more specific labeling of cells and high-resolution selection which can be accomplished using sophisticated labeling techniques. Here, subfractions of a given community are targeted according to different (characteristic) properties that allow discrimination. The following section introduces a set of methods to specifically label cells and discusses their possible downstream analyses.

Isotope-Based Labeling

Stable Isotope Probing. Stable isotope probing (SIP) can identify metabolic fluxes and functional microorganisms using different stable isotopes (13C, 15N, 2H, or 18O) [Coyotzi et al., 2016; Berry and Loy, 2018; Jayan et al., 2022] (Fig. 2a). Here, samples are incubated with a substrate labeled with a heavy stable isotope, and the labeled molecules, such as proteins, DNA, or RNA, are traced [Campana et al., 2021]. Using isopycnic centrifugation, the heavy-labeled molecules form different fractions in a cesium chloride density gradient matrix. Molecules with higher incorporation rates of the heavy isotopes migrate deeper in the applied gravity field and can then be isolated and sequenced. Thus, the identity of the organism refers to the uptake of certain substrates or metabolic processes [Campana et al., 2021]. SIP has been extensively used in microbial ecology, e.g., applying 13C SIP revealed that alternate carbon sources, such as acetate, play a substantial role in the metabolism of potential atmospheric methane oxidants in highland soils [Pratscher et al., 2011]. DNA-SIP combined with metagenomics (metagenomic-SIP) could link specific functions to distinct genomes of complex communities, as well as help assemble target genomes [Wilhelm et al., 2019; Barnett and Buckley, 2020]. Nine MAGs belonging to Proteobacteria, Gemmatimonadetes, and Chloroflexota were retrieved from SIP-13C-labeled DNA in a project investigating active bacteria involved in methane metabolism in high Arctic cryosols and the authors were able to identify non-culturable active methane oxidizers [Altshuler et al., 2022]. Starr et al. traced the carbon flux from plants to a microbial community of the rhizosphere via quantitative SIP and were able to identify different fractions of SIP-labeled prokaryotes [Starr et al., 2021]. However, since the resolution of SIP is low, detection of stable isotope incorporation on single-cell level is not possible to date.

Fig. 2.

Labeling methods for prokaryotic cells. a Stable isotope probing (SIP). Samples are incubated with stable isotope-labeled substrates. Substrate metabolization results in uptake of stable isotopes. b Click chemistry. Samples are incubated with a synthetic substrate suitable for click chemistry followed by incubation with a clickable fluorophore. Click reaction results in fluorescent labeling of cells that metabolized clickable substrates. c Antibody labeling. After sample preparation, samples are incubated with a fluorescently labeled antibody. After recognition of the target epitope by the antibody , cells exhibit fluorescence. d Fluorescence in situ hybridization (FISH). Samples are incubated with a fluorescent probe. After successful probe hybridization cells exhibit fluorescence (for details, see Fig. 3). e Microautoradiography (MAR). Like SIP, the radioactively labeled substrate is metabolized by the cells. β-decay causes black amorphous silver-precipitation by reduction of silver ions which can then be monitored. See text for further details. Created with Biorender.

Fig. 2.

Labeling methods for prokaryotic cells. a Stable isotope probing (SIP). Samples are incubated with stable isotope-labeled substrates. Substrate metabolization results in uptake of stable isotopes. b Click chemistry. Samples are incubated with a synthetic substrate suitable for click chemistry followed by incubation with a clickable fluorophore. Click reaction results in fluorescent labeling of cells that metabolized clickable substrates. c Antibody labeling. After sample preparation, samples are incubated with a fluorescently labeled antibody. After recognition of the target epitope by the antibody , cells exhibit fluorescence. d Fluorescence in situ hybridization (FISH). Samples are incubated with a fluorescent probe. After successful probe hybridization cells exhibit fluorescence (for details, see Fig. 3). e Microautoradiography (MAR). Like SIP, the radioactively labeled substrate is metabolized by the cells. β-decay causes black amorphous silver-precipitation by reduction of silver ions which can then be monitored. See text for further details. Created with Biorender.

Close modal

The combination of SIP with NanoSIMS (NanoSIP) can be employed to identify and target species of interest in a community of microorganisms. One important advantage of NanoSIP is the capacity to image multiple isotopes at the same time [Nuñez et al., 2018]. Many microbiological studies involving NanoSIP focus on microbial metabolism, where primarily 13C and 15N are used to follow interspecies transfer of metabolites [McGlynn et al., 2015; Berry and Loy, 2018; He et al., 2021]. In addition, catabolic degradation and subsequent uptake of extracellular DNA in anoxic marine sediments could be monitored using NanoSIP [Wasmund et al., 2021]. Furthermore, specific carbon assimilation rates from glucose in Pseudomonas putida as well as the cellular density of certain elements like nitrogen, phosphor, and oxygen could be determined on the single-cell level [Stryhanyuk et al., 2018]. In a recent approach, Kitzinger and colleagues [2019] were able to determine nitrite production from urea and cyanate by Thaumarchaeota on single-cell level [Kitzinger et al., 2019]. Using a combination of 15N2-based NanoSIP, CARD-FISH (see below), 16S rRNA and nifH gene sequencing, Woebken and colleagues were able to correlate metabolic function (i.e., N2-incorporation) to a distinct Chloroflexota species of a complex microbial consortium from a photosynthetic microbial mat derived from Monterey Bay (CA, USA) [Woebken et al., 2012]. In a similar approach, sulfate-reducing bacteria could be excluded as potential diazotrophs, whereas Lyngbia sp. could be clearly identified as major N2-fixing organisms in microbial mats from Laguna Ojo de Liebre (Mexico) [Woebken et al., 2015]. However, as mentioned above, the underlying mechanism of SIMS is highly destructive and cells cannot be subjected to further growth experiments or to sequencing approaches. Still, as a supplementary tool combined with other techniques, NanoSIP has shown to be useful in community description, albeit it requires expensive instrumentation.

SIP can also be combined with Raman microscopy and other imaging and spectroscopic techniques to explore the phenotypic characteristics and activities of microbial communities at single-cell resolution [Matanfack et al., 2020; Lee et al., 2021; Jayan et al., 2022]. Muhamadali and co-workers used a simultaneous combination of Raman and FT-IR for distinguishing E. coli cells grown under different carbon and nitrogen isotope concentrations. Based on the different isotope ratios, the combined methods allowed a distinct discrimination of the cells [Muhamadali et al., 2015]. Deuterium in heavy water (D2O) could also be used to identify metabolically active cells via its incorporation into newly synthesized biomolecules. It was shown that incorporation rates highly depend on complexity of the substrate metabolized by the cells. Rather low complex substrates like sugars, short-chain fatty acids, and aromatics yielded higher deuterium incorporation rates compared to peptides or complex medium. Although Raman-based signal identification of deuterium incorporation is comparably easy to detect (due to the signal occurrence in the Raman silent region, 2,040–2,300 cm−1), organism-specific metabolism as well as substrate incorporation have to be taken into account [Matanfack et al., 2020]. Raman-SIP is also substantially less sensitive compared to NanoSIP [Wang et al., 2016].

Single-Cell Capturing via Microautoradiography and Genome Sequencing

Microautoradiography (MAR) uses radioactively labeled compounds such as, e.g.,14C aromatics to label prokaryotic cells based on their metabolic activity. Based on 14C uptake, radioactively labeled cells trigger the conversion of silver cations in radiation-sensitive silver halide emulsion to amorphous, black silver atoms (Fig. 2e) [Brock and Brock, 1966; Nielsen and Nielsen, 2005], which can then be identified as black spots. A major drawback is that MAR can only detect metabolic activity and does not provide any species information. To overcome this issue, MAR-FISH (FISH is discussed below) was established [Lee et al., 1999; Ouverney and Fuhrman, 1999; Wagner et al., 2006]. However, since FISH requires prior knowledge about the sequences to label, this combination excludes unknown taxa from further analysis. Recently, a new technology named single-cell capturing via microautoradiography and genome sequencing was established to close this knowledge gap (Lo et al. under review). Here, radioactively labeled single cells are encapsulated in photoemulsion-hydrogels using microfluidics technology. After conversion of silver ions to black silver atoms, the microcapsules turn black. The black emulsion microcapsules can then be easily detected and sorted via FC. The sorted capsules can be dissolved and the DNA of the cell can be subjected to sequencing. Lo and colleagues isolated and sequenced single benzene-degrading Pseudomonas veronii cells from a mock community. Although application to environmental samples is needed, this technique holds a lot of promise for determining the identity of microorganisms metabolizing certain compounds in a variety of habitats. In contrast to the aforementioned destructive NanoSIP, this technique is highly compatible with downstream applications like genome sequencing on the single-cell level.

Fluorescent Labeling

Due to its good visualization properties, fluorescent labeling is the most common labeling technique in microbiology where a fluorophore is coupled to a targeting molecule (e.g., oligonucleotide, antibody). Fluorescein-derivates (fluorescein–isothiocyanate, 5-[-6-] carboxyfluorescein-N-hydroxysuccinimide-ester) and rhodamine-derivates (tetramethyl-rhodamine-isothiocyanate [TRITC], Texas Red) are popular dyes [Moter and Göbel, 2000]. Cyanine dyes such as Cy3 or Cy5 enhance fluorescence intensity since they show improved resistance to photobleaching. Alexa Fluor® 488, 555, 594, and 647 coumarin dyes are also widely used and have, e.g., enabled the simultaneous localization of numerous sequence-specific RNA species [Frickmann et al., 2017].

Substrate Analog Probing

An alternative to the SIP approach is the so-called substrate analog probing. Here, synthetic substrate ana-logs are used to identify specific cells based on metabolic characteristics. Artificial substrates are either labeled with a fluorophore or are suitable for azide-alkyne click chemistry (see below). Once taken up by the cells these molecules can be tracked [Hatzenpichler et al., 2020]. As opposed to most SIP and some label-free techniques, substrate analog probing has the benefit of using infrastructure that is more easily available, such as fluorescence microscopes and fluorescence-activated cell sorting (FACS) devices. For example, Doud and colleagues were able to identify a new cellulose degrader from the rare biosphere using this function-driven approach [Doud et al., 2020].

Azide–Alkyne Click Chemistry

“Click chemistry” refers to chemical reactions that produce stable heteroatom bonds that can be used for specific identification of particular compounds [Kolb et al., 2001; Moses and Moorhouse, 2007; Hatzenpichler et al., 2020].). Here, a specific cellular molecule is combined with a molecule that is easy to detect in a simple and fast reaction. The tagged cells can then be subjected to identification and sorting methods (Fig. 2b). There are various reactions in nature forming heteroatom bonds. However, click reactions must be modular, stereospecific, broad in scope, provide large yields, and produce only non-toxic byproducts [Kolb et al., 2001]. In more detail, click reaction refers to two types of labeling reactions yielding triazole conjugates: a Cu(I)-catalyzed and a metal-free version. Copper-free click chemistry has been widely used in a variety of biological projects [Baskin et al., 2007; Fugier et al., 2015; Cañeque et al., 2018; Zhang et al., 2021]. Synthetic substrates used in click chemistry are commercially available in several different forms: synthetic substitutes of l-methionine (l-azidohomoalanine [AHA] and l-homopropargylglycine [HPG]), modifiable N-acetylmuramic acid derivatives (MurNAc, NAM [DeMeester et al., 2018; DeMeester et al., 2019]), lipopolysaccharide component analog 8-azido-8-deoxy-Kdo (8AzKdo [Wang et al., 2017a]) and thymidine analogs (5-bromo-2′-deoxyuridine [BrdU] and 5-ethynyl-2′-deoxyuridine [EdU] [Borneman, 1999; Tada and Grossart, 2014; Taniguchi and Eguchi, 2020]) are some examples. Natural substances, such as d-glutamic acid and d-alanine [Liang et al., 2017], as well as numerous forms of glycans [Li et al., 2020; Han et al., 2021], have also been used as labels due to their uniqueness in bacteria. The click reaction offers several benefits, the most notable of which is that it is an enzyme-free reaction that is straightforward to use. In context of targeted labeling of cells, click chemistry reactions are often visualized using fluorescent labels [Ishizuka et al., 2016; Cañeque et al., 2018; Hatzenpichler et al., 2020] (Fig. 2b). New generations of clickable fluorophores, such as picolyl dyes and fluorogenic “turn-on” azide probes (that only react with a fluorescent alkene), are especially well-suited to a wide range of complex samples [Leizeaga et al., 2017; Müller et al., 2019; Hatzenpichler et al., 2020]. Alkyne-conjugated Alexa Fluor®488 in combination with HPG was used to assess the activity of marine bacterial communities in both fast- and slow-growing cells [Samo et al., 2014]. It was possible to discriminate single Alteromonas cells actively synthesizing protein from the majority of resting cells, which remained uncovered by MAR approaches. Still, most clickable substrates have to be incorporated actively by the cells (e.g., during cell wall synthesis and protein biosynthesis). Therefore, click chemistry is rather unsuitable for identification of resting cells or discrimination of slowly growing cells from cells with similar metabolisms with respect to substrate incorporation.

Alternatively, BONCAT can be used, a method of tracking newly synthesized peptides/proteins by the use of amino acid surrogates (e.g., AHA and HPG) [Dieterich et al., 2006]. Since proteins account for the majority of cellular dry weight (50–65%) [Beck et al., 2018], incorporation of amino acid surrogates into newly translated proteins can be used to trace metabolic activity of certain cells in microbial communities [Kiick et al., 2002; Dieterich et al., 2006]. It was corroborated that incorporation of HPG and AHA has only minor impacts on the activity or community structure of bacteria [Hatzenpichler et al., 2016; Steward et al., 2020], as well as on protein tertiary structure [Lehner et al., 2017]. Researchers also used BONCAT to tag the active fraction of anammox microbiota [Chen et al., 2021], discover process-specific microbial features [Du and Behrens, 2021], investigate their in situ activity [Lindivat et al., 2020], and monitor the metabolic response of microbial community members to changing environmental conditions [Reichart et al., 2020]. It was also demonstrated that BONCAT-based labeling can be comparably very fast using AHA (down to 2% of the respective doubling time in E. coli). When BONCAT is combined with rRNA-targeted FISH or CARD-FISH (described below), active cells can be monitored and identified [Hatzenpichler et al., 2016; Pereira et al., 2022]. Furthermore, this technique was applied for the isolation of new strains [Han et al., 2021], classification of bacteria based on their structure (such as Gram-negatives and Gram-positives) [Wang et al., 2017b], studying the composition of the actively growing bacteria at the taxon level [Tada and Grossart, 2014], and examination of extracellular DNA from bacteria [Alhede et al., 2020]. In summary, this technique harbors the potential to label, isolate, and cultivate prokaryotic cells in a rather ambient way for further sequencing as well as cultivation approaches. However, it harbors similar drawbacks like the aforementioned click reactions of only targeting active cells due to the need for active substrate import. Additionally, stability of the substrates was reported to be an issue, especially during long-term exposures in alkaline and/or sulfide-rich environments [Hatzenpichler et al., 2014].

Antibody Labeling

Bio-recognition elements like antibodies and aptamers are excellent tools for specifically labeling particular cellular characteristics. Due to their three-dimensional recognition quality, it is possible to take advantage of structural components of cells in situ as binding motifs to specifically target cells (e.g., cell wall components, outer membrane proteins) (Fig. 2c). Antibodies can be coupled to a variety of reporters (e.g., enzymes, metals, and fluorophores) and be used for labelling also in combination with other methods such as FISH (discussed below) and SERS (see above) [Pal et al., 2017; Lee et al., 2019b; Wang et al., 2021]. Since binding of antibodies happens on the outside of the cell this technique circumvents drawbacks like fixation and permeabilization steps coming along with, e.g., FISH-based labeling methods (described below). Using polyclonal antibodies designed against reference strains, Bellais et al. were able to isolate the gut bacterium Faecalibacterium prausnitzii from human fecal samples. The approach was also successfully applied for isolating the rarely present Christensenella minuta from the same samples using FC [Bellais et al., 2022]. A promising approach to target cells from MDM where no isolates or closely related reference strains are available is to apply reverse genetics. A proof of principle was recently shown by Cross and colleagues. Based on SAG sequence information for protein domain prediction extracellularly exposed peptide fragments (epitopes) were synthesized and served as antigens. Human oral samples labeled with these antibodies showed 25–100% recovery of the target organism Saccharibacterium sp. TM7 based on 16S rRNA sequencing [Cross et al., 2019]. Although the underlying genetic information was SAG-derived, this approach might also be useful if only MAG information is available when targeting specific (outer) membrane components. Although antibodies harbor a great potential for specifically labeling distinct organisms due to unique epitopes they, however, usually exhibit a certain cross-reactivity resulting in reduced selectivity and resolution [Kramer et al., 1997; Holm et al., 2012; Landegren et al., 2012].

Fluorescent in situ Hybridization

Fluorescently labeled oligonucleotides complementary to certain DNA or RNA sequences are widely used for the targeted labelling of cells (Fig. 2d, 3) [Antón et al., 1999; Pernthaler et al., 2001; Wagner et al., 2003a; Sekar et al., 2004; Fazi et al., 2007]. In general, short oligonucleotide probes (15–25 nucleotides) attach to defined sequence motifs of nucleic acids (i.e., genes, mRNA, lncRNA). FISH methods usually comprise permeabilization of the cell wall allowing the probe to enter the cell, binding of the labeled probe to the target sequence, and fluorescence-based detection of the bound probe either via microscopy or other fluorescence-based detection methods (e.g., FC, see below). However, in many cases the fluorescence signal from bound probes is too faint for proper detection for manifold reasons. For example, low cellular activity can be an issue when probes are designed for mRNA or rRNA targets and low genomic copy numbers cause weak signals when probes are designed to match specific marker genes causing signal strengths close to or even below the detection limit [Albertson et al., 1995; Wagner et al., 2003b].

Fig. 3.

Fluorescence in situ hybridization (FISH). The basic principle is probe penetration into the cell followed by hybridization against a target sequence. Hybridization is carried out with either DNA or RNA as target (blue box). Important subclasses of FISH techniques are shown. a DOPE-FISH uses double-labeled probes for higher signal strength. b MiL-FISH increases signal intensity by adding more fluorophores to the particular probe. c CLASI-FISH uses combinations of different fluorescently labeled probes. d HCR-FISH takes advantage of fluorescently labeled amplifier probes that hybridize against a special nonbinding part of the target-specific initiator probe. e CARD-FISH is conducted via a horseradish peroxidase (HRP) coupled to the probe. Tyramides attached to the fluorophore are converted to a radical state that causes immediate binding to aromatic compounds in the close vicinity (Trp, Phe) thereby increasing the signal intensity. f RING-FISH, recognition of individual genes is carried out using very long polynucleotides that form secondary structures. These molecules bind partially to each other thereby increasing fluorescence signal intensity. Created with Biorender. DOPE-FISH, double-labeled oligonucleotide probe FISH; MiL-FISH, multi-labeled oligonucleotide probe FISH; CLASI-FISH, combinatorial labeling and spectral imaging-FISH; HCR-FISH, hybridization chain reaction-FISH; CARD-FISH, catalyzed reporter deposition-FISH.

Fig. 3.

Fluorescence in situ hybridization (FISH). The basic principle is probe penetration into the cell followed by hybridization against a target sequence. Hybridization is carried out with either DNA or RNA as target (blue box). Important subclasses of FISH techniques are shown. a DOPE-FISH uses double-labeled probes for higher signal strength. b MiL-FISH increases signal intensity by adding more fluorophores to the particular probe. c CLASI-FISH uses combinations of different fluorescently labeled probes. d HCR-FISH takes advantage of fluorescently labeled amplifier probes that hybridize against a special nonbinding part of the target-specific initiator probe. e CARD-FISH is conducted via a horseradish peroxidase (HRP) coupled to the probe. Tyramides attached to the fluorophore are converted to a radical state that causes immediate binding to aromatic compounds in the close vicinity (Trp, Phe) thereby increasing the signal intensity. f RING-FISH, recognition of individual genes is carried out using very long polynucleotides that form secondary structures. These molecules bind partially to each other thereby increasing fluorescence signal intensity. Created with Biorender. DOPE-FISH, double-labeled oligonucleotide probe FISH; MiL-FISH, multi-labeled oligonucleotide probe FISH; CLASI-FISH, combinatorial labeling and spectral imaging-FISH; HCR-FISH, hybridization chain reaction-FISH; CARD-FISH, catalyzed reporter deposition-FISH.

Close modal

To increase the fluorescence signal several improved methods have been developed, such as double-labeled oligonucleotide probes FISH (5′ and 3′ ends, Fig. 3a) or multi-labeled oligonucleotide probes FISH (Fig. 3b) which rely on the use of fluorescently labeled probes that have been labeled multiple times [Stoecker et al., 2010; Schimak et al., 2016]. In addition one or more types of fluorescent dyes can be used [Behnam et al., 2012]. Using the double-labeled oligonucleotide probes FISH technology, a novel species within the Beijerinckiaceae, closely related to Methylocapsa, was discovered in a study to detect atmospheric methane-oxidizing bacteria in the USCα clade from forest soil [Pratscher et al., 2018]. The first finding of an archaeal symbiont-host relationship was made by using multi-labeled oligonucleotide probes FISH [Schwank et al., 2019]. An alternative strategy for increasing the fluorescence signal is to label an organism with two or more probes targeting different genes [Azimi et al., 2022] which is called combinatorial labeling and spectral imaging-FISH (Fig. 3c) [Valm et al., 2011]. For example, a study revealed the microbial complexity in the human tongue using this technique [Wilbert et al., 2020].

Fluorescent signals can also be amplified based on an isothermal amplification technology. In situ DNA-hybridization chain reaction (DNA-HCR) uses an initiator probe partially hybridizing against the target sequence. The unbound part serves as a target to trigger the polymerization of two fluorescently labeled amplifier probes [Yamaguchi et al., 2015; Jia et al., 2021] (Fig. 3d). DNA-HCR is described to be a reasonable labeling technique for environmental samples reaching the same or even higher labeling efficiency compared to CARD-FISH (described below) and is suitable for downstream applications like SIP, FC, and next-generation sequencing [Yamaguchi et al., 2015]. DNA-HCR in combination with FACS was used for subsequent whole-genome sequencing and retrieval of MAGs for genomic annotation and characterization of uncultured phytoplankton bloom-associated flavobacterial clade Vis6 [Grieb et al., 2020].

Catalyzed reporter deposition FISH (CARD-FISH) is a FISH variation approach that can boost the fluorescence signal up to 41 times compared to standard FISH [Pereira et al., 2022]. Here, peroxidase activity leads to the deposition of a considerable number of labeled tyramine molecules [Eickhorst and Tippkötter, 2008]. Tyramines are phenolic substances that can be dimerized by the enzyme horseradish peroxidase (HRP). Many fluorescent molecules can be incorporated at the hybridization site in situ if fluorochrome-labeled tyramides are used [Pernthaler et al., 2002]. Here, HRP is used to label the probes. The signal from fluorescently labeled tyramides is then amplified by catalyzed reporter deposition (CARD) [Kubota, 2013] (Fig. 3e). CARD-FISH was first applied to environmental microbes in 1997 in attempts to identify prokaryotes from natural environments which feature a low ribosome content compared to cultured microbes [Lebaron et al., 1997; Schönhuber et al., 1997]. The first visual evidence for Heimdallarchaeota was observed using specifically tailored CARD-FISH probes [Salcher et al., 2020]. The same study revealed a different cell shape of Lokiarchaeota in environmental samples compared to cultured strains and indicated a condensed DNA localization atypical for prokaryotes. Still, when it comes to FC-based sorting signal intensities might be too low for proper detection of very small cells with low ribosome content (e.g., ultramicrobacteria). To solve this problem, a second signal amplification step can be performed, which is mediated by HRP-labeled antibodies targeted to the fluorophores already deposited by CARD-FISH. This further development is called double CARD-FISH (2C-FISH), where two rounds of signal amplification are conducted [Neuenschwander et al., 2015]. The first round couples HRP-attached anti-fluorophore antibodies to the hybridized fluorophore, and the second is used for signal amplification with fluorophore-labeled tyramides. This technique was successfully used to sort LD12 ultramicrobacteria by FC from samples obtained from an oligo-mesotrophic lake. In a proof of principle study the authors were also able to show successful amplification of the 16S rRNA gene from whole-genome amplification (WGA) products (from multiple cells) after 2C-FISH treatment using ethanol as well as formaldehyde as initial fixative [Neuenschwander et al., 2015].

While traditional FISH methods target the 16S rRNA gene and its product, recognition of individual genes (RING)-FISH is used to target metabolic marker genes. The polynucleotide probe is labeled with a reporter molecule during probe synthesis via in vitro transcription. During hybridization, the probe subsequently anchors further probes due to secondary structures formed by incomplete binding (secondary structure formation) of the large probes, thereby forming a network around the cell periphery resulting in a halo-shaped fluorescent signal [Zwirglmaier et al., 2004; Pereira et al., 2022] (Fig. 3f). RING-FISH was, e.g., used to identify nonredundant type III secretion systems (T3SS) in Vibrio parahaemolyticus strains isolated from the Gulf of Mexico [Noriea et al., 2010].

A combination of different approaches can also be an effective way to improve cell labelling efficiency and increasing targeting specificity and thereby enhancing downstream sorting selectivity. The simultaneous application of FISH and BONCAT (see above) is a successful example of combining techniques. This approach links the cells’ metabolic/translational active state (BONCAT) of the cell with species-specific markers (FISH), thus enabling the discrimination of metabolically active cells of a certain type from a whole community in situ. In a mock community of bacteria, the successful combination of FISH and BONCAT was able to detect Methylococcaceae sp. WF1 [Hatzenpichler et al., 2014], however, no sequencing or cultivation was performed.

Although FISH-based techniques have many advantages such as specificity, adaptability, and low costs, they also harbor some drawbacks, especially when it comes to sensitivity. Depending on the probe’s target, the organism’s physiological state and cell wall status as well as the detection method the intracellular fluorophore concentration might not exceed the detection limit. One major problem is the low penetration of the probe into the cell due to insufficient permeabilization of the cell wall. Common fixatives such as paraformaldehyde (for Gram-negatives) and ethanol (for Gram-positives) help permeabilizing the cell wall; however, they also inevitably destabilize the cell wall and damage the DNA thereby limiting downstream research possibilities such as sorting and/or single-cell sequencing approaches [Douglas and Rogers, 1998; Clingenpeel et al., 2014; Doud and Woyke, 2017; Dam et al., 2020]. There have been several studies published in the past years to overcome this problem, e.g., by eliminating the fixation steps from the conventional FISH protocols [Yilmaz et al., 2010; Haroon et al., 2013]. However, removing these chemical fixation substances also significantly reduces the probe’s penetration rate and results in weak or no signals. To enhance the probe’s penetration into the cell ethanol dehydration was shown to result in almost the same hybridization efficiency (E. coli) or an even slightly higher (Bacillus megaterium) when using the in-solution protocol omitting paraformaldehyde fixation [Haroon et al., 2013]. The live-FISH protocol developed by Batani and colleagues (based on the aforementioned protocol by Haroon et al. [2013]) was able to increase the outcome of viable cells by a factor of 10, resulting in >1% viable cells at the end of the protocol by replacing ethanol dehydration series with PBS washing and introducing an additional heat shock. The live-FISH protocol was shown to keep cells alive allowing cultivation even after undergoing a FACS-based sorting process [Batani et al., 2019]. Dam et al. simply used higher probe concentrations and longer hybridization times to gain sufficient fluorescent intensities in environmental samples. Following phylogenetic labeling of target cells and sorting, so far unknown Chloroflexota species from wastewater treatment plants could be isolated and subjected to genome sequencing [Dam et al., 2020].

Recently, a new fixation-free labeling approach for archaea and bacteria has been developed called FISH of transcript-annealing molecular beacons (FISH-TAMB) [Harris et al., 2022]. Here, a molecular beacon (MB) is designed to target the mRNA of bacteria and archaea. The MB comprises an oligonucleotide forming a hairpin structure connected to a fluorophore and a fluorescence quencher at the opposite ends of the hairpin structure so that fluorescence is quenched in the unbound state. The hairpin structure is dissolved in the presence of the target sequence thereby releasing the quencher from the fluorophore’s proximity and resulting in a fluorescence signal. MBs are transferred into the cells by cell-penetrating peptides which in experiments did not affect cell viability. FISH-TAMB is a promising approach to specifically target metabolically active cells since it targets mRNA and can potentially also discriminate between different levels of transcription. It has to be mentioned that so far only model organisms (E. coli and Methanosarcina barkeri) were tested under laboratory conditions and that FISH-TAMB still has to prove its value when it comes to environmental samples and in situ conditions.

The opportunity to selectively identify or target cells from a community facilitates enrichment of certain subfractions of populations or even isolation at the single cell level. The downstream goal of the specific research interest ultimately specifies which sorting techniques are reasonable. The following section summarizes prominent techniques with respect to accuracy, throughput and downstream cell viability.

Flow Cytometry

FC is a technique for assessing large heterogeneous cell populations that uses a laser to detect light scattered or emitted by the cells [Givan, 2011; McKinnon, 2018]. FC coupled with specific fluorescence sensors allows for subclass cell sorting as well as sorting of single cells [Moor et al., 2016; Hiramatsu et al., 2019; Luo et al., 2020; Pereira et al., 2022]. FACS is frequently used to sort prokaryotic cells and sorters are commercially available from many companies in different price classes.

Fluorescence-Activated Cell Sorting

FACS technology relies on the detection of fluorescence signals caused by the cells themselves (autofluorescence), by fluorescently labeled antibodies or ligands that bind to certain cell-associated molecules. Besides fluorescence, simple scattering of light can also be detected and used for characterization of cells in terms of size and complexity/granularity as detected by forward scatter (FSC) and side scatter (SSC) [Veal et al., 2000; Luo et al., 2020; Vollmers et al., 2023]. Using a droplet deflection technique, FACS sorters are capable of sorting tens of thousands of cells per second enabling high throughput analyses to detect, count as well as sort subfractions of a complex heterogenous sample [Veal et al., 2000] (Fig. 4a). For example, Tada and Grossart identified a shift in community composition of actively growing bacteria in response to N-acetyl-glucosamine feeding at the taxon level by using a combination of BrdU-labeling and FACS-based cell sorting [Tada and Grossart, 2014]. Although FACS is a well-established, easy-to-use, fast, and precise option to sort cells, it has some limitations. The high pressure applied to cells during the sorting process might harm fragile cells. If researchers aim to cultivate from single-cells, these sorting conditions can be too harsh for some cell types [Pereira et al., 2022]. Low stability of the used fluorescence probes, bleaching and especially intrinsic background fluorescence from cells can lead to low sensitivity and “cross-talk” resulting in decreased purity and/or sorting efficiency. A major drawback is that obligately anaerobic microorganisms require strict anoxic conditions as long as cells should be kept alive after sorting. Due to large buffer volumes, these conditions are difficult to maintain, although a successful attempt has been reported [Thompson et al., 2015].

Fig. 4.

Sorting techniques for prokaryotic cells. Cells can be subjected to labelling procedures prior to analysis or analyzed directly based on intrinsic sorting factors. a Flow-cytometry. Cells are diluted to a concentration, where single-cell droplets statistically occur. Cells in droplets are analyzed based on size, granularity, and/or fluorescence. Target cells can be chosen from the scattergrams based on the respective parameters and droplets are being charged accordingly. Deflectors guide the charged droplets to corresponding collection vessels (e.g., Eppendorf tubes or well plates). b Microfluidics platform. Cells are loaded onto the device in aqueous phase. Following a laminar flow droplet formation is accomplished by orthogonal oil application resulting in distinct droplets containing a single cell. Between droplet formation and ultimate sorting decision droplets can be analyzed, e.g., via fluorescence detectors, photodiodes, or Raman detectors (not shown). c Cell printer. Cells are diluted to a concentration, at which single-cell events occur at the tip of the nozzle where the droplet is formed by a piezo-actuated deflection of a silicon membrane. A fluorescence detector decides the fate of the droplet. Non-fluorescent droplets are eliminated from collection by a vacuum tube. Target cells are collected in batch (Eppendorf tube) or as single cells by use of microtiter plates (left inset). FSC, forward scatter; SSC, side scatter; MAR, microautoradiography. Created with Biorender.

Fig. 4.

Sorting techniques for prokaryotic cells. Cells can be subjected to labelling procedures prior to analysis or analyzed directly based on intrinsic sorting factors. a Flow-cytometry. Cells are diluted to a concentration, where single-cell droplets statistically occur. Cells in droplets are analyzed based on size, granularity, and/or fluorescence. Target cells can be chosen from the scattergrams based on the respective parameters and droplets are being charged accordingly. Deflectors guide the charged droplets to corresponding collection vessels (e.g., Eppendorf tubes or well plates). b Microfluidics platform. Cells are loaded onto the device in aqueous phase. Following a laminar flow droplet formation is accomplished by orthogonal oil application resulting in distinct droplets containing a single cell. Between droplet formation and ultimate sorting decision droplets can be analyzed, e.g., via fluorescence detectors, photodiodes, or Raman detectors (not shown). c Cell printer. Cells are diluted to a concentration, at which single-cell events occur at the tip of the nozzle where the droplet is formed by a piezo-actuated deflection of a silicon membrane. A fluorescence detector decides the fate of the droplet. Non-fluorescent droplets are eliminated from collection by a vacuum tube. Target cells are collected in batch (Eppendorf tube) or as single cells by use of microtiter plates (left inset). FSC, forward scatter; SSC, side scatter; MAR, microautoradiography. Created with Biorender.

Close modal

Microfluidic Technologies

Microfluidic technologies offer solutions for sample preparation, isolation, and identification of microorganisms. These techniques use very distinct configurations of microchannels embedded in a bio-compatible matrix (Fig. 4b). Many studies use polydimethylsiloxane (PDMS) since it is cheap, inert, easy to handle, and allows gas penetration [Merkel et al., 2000; Lee et al., 2019a]. Other polymer-based materials are chosen primarily for their simplicity of surface modification and superior biocompatibility. Meanwhile, new materials for quick microbial detection, such as hydrogels and paper, have been incorporated into microfluidic circuits [Zhou et al., 2019]. Antibodies [Abafogi et al., 2020], antimicrobial peptides (AMPs) [Dao et al., 2018], aptamers [Su et al., 2020a], bacteriophages [Hussain et al., 2021], and molecularly imprinted polymers (MIPs) [Buensuceso et al., 2022] have all been exploited as functional coatings on the surface of microchannels.

Raman-Activated Cell Sorting

Raman-activated cell sorting spans the boundary between FC and microfluidics. Although the throughput of the RACS systems described below is higher than conventional microfluidic devices, a microfluidic chip or at least channel is the core component for separation of cells. Obtaining a full single-cell Raman spectrum (SCRS) is quite time-consuming (∼1 s). Using coherent Raman scattering and coherent anti-Stokes Raman scattering, the enhanced signal reduces acquisition time down to ms or even µs [Hiramatsu et al., 2019, 2020; Suzuki et al., 2019]. Still, even when using SERS or CARS the throughput remains orders of magnitude lower compared to FACS [Hiramatsu et al., 2019; Gala De Pablo et al., 2021]. Although attempts were made to sort microalgae, yeast, and bacteria, the reported throughput and efficiency were rather low compared to “classical” FACS sorting [Camp et al., 2011; Huang et al., 2017; Hiramatsu et al., 2019]. However, RACS-based sorting is a useful tool in cases where label-free characterization is essential or the application of specific labels is ambiguous or not feasible. RACS has been shown to be capable of in situ single-cell detection, identification, and sorting of active cells in complex environments and was successfully used to sort carotenoid-containing cells (Pantoea spp., Legionella spp., Massilia spp., Pseudomonas spp., and Pedobacter spp.) from mock bacterial communities [Jing et al., 2021; Li et al., 2022]. Although Raman-based characterization performs well in the absence of markers, combining it with SIP by the preceding addition of, e.g., D2O substantially expands the technique’s potential. In a study by Jing and co-workers, metabolically active cells (Corynebacterium spp., Clostridium spp., Moraxella spp., Pantoea spp., and Pseudomonas spp.) were sorted by RACS after the bacterium was D2O-marked [Jing et al., 2021]. Metabolic activity of single bacteria from mouse cecal microbiomes could be followed combining D2O-labelling and downstream RACS sorting followed by MDA and 16S rRNA-based identification [Berry et al., 2015].

There are several combinations of Raman-based signal acquisition and sorting mechanisms described resulting in a multitude of RACS subclasses where Raman-activated microfluidic sorting (RAMS), Raman-activated droplet sorting (RADS), and Raman-activated cell ejection (RACE) are the most relevant [Yan et al., 2021]. RACE is a RACS variant allowing sorting of single cells. Here, Raman signals are used to characterize cells mounted on a coated slide attached to a collection vessel (e.g., PDMS chip containing cavities). Collection of single cells is conducted by selectively releasing selected cells from the mounting slide to the collection vessel by, e.g., using laser-induced forward transfer [Hopp et al., 2005]. Here, the coating material in the proximity of the cell is evaporated by a laser, thereby pushing or releasing the cell into the collection vessel underneath the mounting slide. RACE was also performed to classify five oral bacteria (Streptococcus sanguinis, S. mutans, Porphyromonas gingivalis, Enterococcus faecalis, Actinomyces viscosus). Although it was possible to obtain and amplify DNA from single cells, the isolated cells were quickly destroyed as a result of the procedure [Wang et al., 2013]. In a study conducted by Wang et al. [2013], D2O-marked single antibiotic-resistant bacteria (ARBs) were characterized via SCRS, sorted using RACE, and the sorted cells were subjected to single-cell genome amplification downstream [Wang et al., 2020]. Although Raman-based signal acquisition is basically noninvasive, nondestructive and leaves the cells in a viable state after sorting, results of genome amplification after RACE are rather limited. Su and colleagues were able to show that laser energy utilized for Raman signal acquisition has a dramatic impact on downstream sequencing or WGA efficiency, most probably due to heat-generation caused by energy consumption during signal acquisition. The overall impact was determined as a reduction from 80.5% to 36.3% sequence coverage by acquiring SCRS before the ejection step. They were also able to determine the impact of laser energy absorbed by the cells where they showed that applying 3 mW for 10 s is a good compromise between a clear enough SCRS (signal-to-noise ratio above 3) and sufficient DNA quality for downstream WGA attempts, whereas signal acquisition at 3 mW for 20 s as well as 60 s decreased downstream genome amplification down to 33% and 0% success rate, respectively [Su et al., 2020b]. Of note, obtained WGA results based on MDA were carried out using at least 5 sorted cells per reaction since all single cell attempts failed at the level of WGA.

Zhang et al. [2015] developed a RAMS system that was able to discriminate carotenoid-producing yeast cells from non-producing strains with a subsecond event rate and an accuracy of around 73% [Zhang et al., 2015]. The combination of applying optical tweezers into a RAMS platform enabled automatization of the process. In this setup, cells enter a microfluidic device and are ordered via two sheath flows to form a row of cells. Optical tweezers capture a single cell and move it to the detection area where the SCRS is acquired. If the SCRS matches sorting criteria (e.g., presence of carotenes, incorporation of isotopes) the tweezers release the cell to the collection port, whereas they will move the cells back into waste stream if cells do not match sorting criteria [Lee et al., 2019a]. Using this technique combined with SIP, it was possible to analyze and collect up to 8 cells per minute. Since laser energies applied in this approach are low due to the short Raman signal acquisition (300 ms at 532 nm) cells sorted with this approach are generally viable and can be subjected to genomic analyses as well as cultivation attempts [Lee et al., 2021]. Studies aiming at sorting of specific cells also utilize oil to form small droplets wherein cells were trapped. Since droplet formation prior to Raman signal acquisition causes perturbations, droplets in RADS attempts were generated after spectral profiling and subsequently subjected to downstream applications. Using droplet-based microfluidic technologies Wang and colleagues were able to drastically increase screening rate of RAMS systems to 260 cells per minute [Wang et al., 2017b]. Moreover, the RADS setup used preserved sorted cells in a viable state with a proportion similar to untreated cells (>90%). In this setup, all investigated cells are incorporated into oil droplets. Depending on the experimental criteria droplets of cells considered positive are separated from the rest via an electric field. In addition, application of an electric field could replace optical tweezers in the setup, thereby decreasing chances of photodamage. Besides throughput, the major drawback of RACS compared to FACS is that there is just one system commercially available. All setups described here are custom-build.

Cell Printing

Labeling and cell sorting techniques, as previously indicated, can exert stress on cells and can cause cell rupture. Comparatively mild conditions can be maintained by so-called cell printers. While there are different direct and indirect printing attempts described (starting from modified inkjet printers, reviewed in [Gross et al., 2013; Zhou et al., 2022]), we focus on the recent advances in contact-free cell dispensers enabling WGA, cultivation an automatization. The basic principle behind this printing technique is generation of very small (pico liter scale) droplets harboring a single cells that are then deposited to a micro-well or agar plates by means of a dispenser chip [Gross et al., 2015] (Fig. 4c). Droplets are generated via piezo-actuated deflection of a silicon-membrane and formation is monitored via a bright-field video microscope, allowing automatic single-cell sorting of different cell sizes. However, the sensitivity is considerably reduced compared to, e.g., FACS, making it challenging when sorting is based on weak signal strengths. Further improvements facilitate smaller droplet volumes (down to 35 pL) as well as higher resolution (>1 µm) to reliably identify and print single bacteria with efficiencies >90% with a collection rate of 96 wells (i.e., cells) in approximately 10 min. However, it should be noted that this rate is significantly lower if only a rare portion of the cells in a community are to be sorted since the appearance of targeted cells within the nozzle happens randomly. A now commercially available setup enabled the shape-based sorting (drop-on-demand) of a mixture of E. coli (rods) and E. faecalis (cocci) followed by successful cultivation of the sorted single cells [Riba et al., 2016]. Successful label-free single-cell printing was also conducted to isolate cells from a complex wastewater treatment plant sample. Subsequent WGA and sequencing revealed 80% success rate for genome amplification and finally resulted in 27 SAGs from novel members of Patescibacteria, indicating that the gentle sorting conditions are beneficial for downstream WGA-based applications [Wiegand et al., 2021]. Although throughput of this technique is rather low compared to microfluidics and FC, this technique has the benefit of requiring only very small sample volumes ranging from 5 to 100 µL [Gross et al., 2015]. Moreover, one of its benefits is that the whole setup can be easily placed in an anaerobic glove box facilitating anoxic conditions throughout the sorting process. A recent study showed the successful isolation of 21 taxa of the human gut microbiome using anoxic single-cell dispensing. This approach shows the applicability of cell printing with respect to cultivation attempts since the (untargeted) cells were printed directly onto agar plates and colony formation from single-cell growth was followed [Afrizal et al., 2022]. Moreover, it shows that single-cell sorting is possible also under anoxic conditions, thus allowing sorting of strictly anaerobic microorganisms that are often excluded using “classic” sorting methods like FACS as mentioned above. Unfortunately, the design of the instrument prevents high sensitivity in the detection of fluorescence signals at the moment, which is why weakly fluorescing cells cannot be reliably detected. Since the microscopical observation of the cells is maintained via mirrors it seems feasible to couple cell printers to fluorescence and also Raman detectors in the future, thereby extending applications to targeted sorting, resulting in a compromise of throughput and viable cells.

Workflows for targeted labelling and sorting of prokaryotic cells are still lagging behind methods for eukaryotic research. The obvious reason is the predicament of size and – for sequencing attempts – the reduced amount and stability of nucleic acids. Reduced cell size results in comparatively low signal strength and technical issues when it comes to sorting attempts. In addition, the manifold cell wall structures and complexities of prokaryotic organisms impede a precise prediction of cell lysis conditions – some cells might lyse during the sorting process while others do not lyse at all when it would be needed, even when applying lysis treatment after sorting. In addition, many of the techniques and their respective combinations (identification and sorting) described in this review are custom laboratory setups tailor-made for a particular research focus and only used in the specific laboratories that published the respective research. The considerable advances in the last years regarding prokaryotic cell labelling and sorting raise hope for further commercially available devices in the future. Finally, one has to individually decide which techniques are suitable for a certain research question and must carefully take all the pros and cons into account before selecting a suitable method. Researchers not aiming at cultivation might select rather invasive identification methods (e.g., NanoSIMS). When higher cell viability or integrity is needed the smoother the targeting or identification techniques have to be. The same holds true for the particular question being addressed – methods aiming at specific metabolic processes significantly differ from methods targeting specific cell components. Likewise, complexity and origin of the sample have to be taken into account. The multitude of factors that must be considered when selecting a specific method makes targeted cell sorting an exciting and versatile tool for answering a wide variety of research questions.

The authors declare that they have no conflict of interest.

This work has been supported by the German Bundesministerium für Bildung und Forschung (BMBF) under project number 161 L0284A (Verbundprojekt MicroMatrix – Entwicklung einer Pipeline zur Kultivierung und Analyse bisher nicht kultivierbarer Mikroorganismen mit biotechnologischer Relevanz aus komplexen Umweltproben) and project number 03F0905A (MAPMAR: Marine Plasmids Driving the Spread of Antibiotic Resistances).

G.S., M.M., and A.-K.K. wrote the manuscript and approved the final version.

1.
Abafogi
AT
,
Kim
J
,
Lee
J
,
Mohammed
MO
,
van Noort
D
,
Park
S
.
3D-printed modular microfluidic device enabling preconcentrating bacteria and purifying bacterial DNA in blood for improving the sensitivity of molecular diagnostics
.
Sensors
.
2020
;
20
(
4
):
1202
.
2.
Afrizal
A
,
Hitch
TCA
,
Viehof
A
,
Treichel
N
,
Riedel
T
,
Abt
B
.
Anaerobic single-cell dispensing facilitates the cultivation of human gut bacteria
.
Environ Microbiol
.
2022
;
24
(
9
):
3861
81
.
3.
Akanny
E
,
Bonhommé
A
,
Commun
C
,
Doleans-Jordheim
A
,
Farre
C
,
Bessueille
F
.
Surface-enhanced Raman spectroscopy using uncoated gold nanoparticles for bacteria discrimination
.
J Raman Spectrosc
.
2020
;
51
(
4
):
619
29
.
4.
Albertson
DG
,
Fishpool
RM
,
Birchall
PS
.
Fluorescence in situ hybridization for the detection of DNA and RNA
.
Methods Cell Biol
.
1995
;
48
:
339
64
.
5.
Alhede
M
,
Alhede
M
,
Qvortrup
K
,
Kragh
KN
,
Jensen
,
Stewart
PS
.
The origin of extracellular DNA in bacterial biofilm infections in vivo
.
Pathog Dis
.
2020
78
2
ftaa018
.
6.
Alteio
LV
,
Schulz
F
,
Seshadri
R
,
Varghese
N
,
Rodriguez-Reillo
W
,
Ryan
E
.
Complementary metagenomic approaches improve reconstruction of microbial diversity in a forest soil
.
mSystems
.
2020
5
2
e00768-19
.
7.
Altshuler
I
,
Raymond-Bouchard
I
,
Magnuson
E
,
Tremblay
J
,
Greer
CW
,
Whyte
LG
.
Unique high Arctic methane metabolizing community revealed through in situ 134-DNA-SIP enrichment in concert with genome binning
.
Sci Rep
.
2022
;
12
(
1
):
1160
.
8.
Amarasinghe
SL
,
Su
S
,
Dong
X
,
Zappia
L
,
Ritchie
ME
,
Gouil
Q
.
Opportunities and challenges in long-read sequencing data analysis
.
Genome Biol
.
2020
;
21
:
30
16
.
9.
Ammor
MS
.
Recent advances in the use of intrinsic fluorescence for bacterial identification and characterization
.
J Fluoresc
.
2007
;
17
(
5
):
455
9
.
10.
Anggraini
D
,
Ota
N
,
Shen
Y
,
Tang
T
,
Tanaka
Y
,
Hosokawa
Y
.
Recent advances in microfluidic devices for single-cell cultivation: methods and applications
.
Lab Chip
.
2022
;
22
(
8
):
1438
68
.
11.
Antón
J
,
Llobet-Brossa
E
,
Rodríguez-Valera
F
,
Amann
R
.
Fluorescence in situ hybridization analysis of the prokaryotic community inhabiting crystallizer ponds
.
Environ Microbiol
.
1999
;
1
(
6
):
517
23
.
12.
Arisdakessian
CG
,
Nigro
OD
,
Steward
GF
,
Poisson
G
,
Belcaid
M
.
CoCoNet: an efficient deep learning tool for viral metagenome binning
.
Bioinformatics
.
2021
;
37
(
18
):
2803
10
.
13.
Azimi
S
,
Lewin
GR
,
Whiteley
M
.
The biogeography of infection revisited
.
Nat Rev Microbiol
.
2022
;
20
:
579
92
.
14.
Barnett
SE
,
Buckley
DH
.
Simulating metagenomic stable isotope probing datasets with MetaSIPSim
.
BMC Bioinformatics
.
2020
;
21
(
1
):
37
.
15.
Baskin
JM
,
Prescher
JA
,
Laughlin
ST
,
Agard
NJ
,
Chang
PV
,
Miller
IA
.
Copper-free click chemistry for dynamic in vivo imaging
.
Proc Natl Acad Sci U S A
.
2007
;
104
(
43
):
16793
7
.
16.
Batani
G
,
Bayer
K
,
Böge
J
,
Hentschel
U
,
Thomas
T
.
Fluorescence In Situ Hybridization (FISH) and cell sorting of living bacteria
.
Sci Rep
.
2019
;
9
:
18618
3
.
17.
Beck
AE
,
Hunt
KA
,
Carlson
RP
.
Measuring cellular biomass composition for computational biology applications
.
Process
.
2018
;
6
(
5
):
38
.
18.
Behnam
F
,
Vilcinskas
A
,
Wagner
M
,
Stoecker
K
.
A straightforward DOPE (double labeling of oligonucleotide probes)-FISH (fluorescence in situ hybridization) method for simultaneous multicolor detection of six microbial populations
.
Appl Environ Microbiol
.
2012
;
78
(
15
):
5138
42
.
19.
Bellais
S
,
Nehlich
M
,
Ania
M
,
Duquenoy
A
,
Mazier
W
,
van den Engh
G
.
Species-targeted sorting and cultivation of commensal bacteria from the gut microbiome using flow cytometry under anaerobic conditions
.
Microbiome
.
2022
;
10
:
24
17
.
20.
Berdy
B
,
Spoering
AL
,
Ling
LL
,
Epstein
SS
.
In situ cultivation of previously uncultivable microorganisms using the ichip
.
Nat Protoc
.
2017
;
12
(
10
):
2232
42
.
21.
Berry
D
,
Loy
A
.
Stable-isotope probing of human and animal microbiome function
.
Trends Microbiol
.
2018
;
26
(
12
):
999
1007
.
22.
Berry
D
,
Mader
E
,
Lee
TK
,
Woebken
D
,
Wang
Y
,
Zhu
D
.
Tracking heavy water (D2O) incorporation for identifying and sorting active microbial cells
.
Proc Natl Acad Sci U S A
.
2015
112
2
E194
203
.
23.
Bickhart
DM
,
Kolmogorov
M
,
Tseng
E
,
Portik
DM
,
Korobeynikov
A
,
Tolstoganov
I
.
Generating lineage-resolved, complete metagenome-assembled genomes from complex microbial communities
.
Nat Biotechnol
.
2022
;
40
(
5
):
711
9
.
24.
Bodelón
G
,
Montes-García
V
,
Pérez-Juste
J
,
Pastoriza-Santos
I
.
Surface-enhanced Raman scattering spectroscopy for label-free analysis of P. aeruginosa Quorum Sensing
.
Front Cell Infect Microbiol
.
2018
;
8
:
143
.
25.
Borneman
J
.
Culture-independent identification of microorganisms that respond to specified stimuli
.
Appl Environ Microbiol
.
1999
;
65
(
8
):
3398
400
.
26.
Brock
TD
,
Brock
ML
.
Autoradiography as a tool in microbial ecology
.
Nature
.
1966
;
209
(
5024
):
734
6
.
27.
Buensuceso
CE
,
Tiu
BDB
,
Lee
LP
,
Sabido
PMG
,
Nuesca
GM
,
Caldona
EB
.
Electropolymerized-Molecularly Imprinted Polymers (E-MIPS) as sensing elements for the detection of dengue infection
.
Anal Bioanal Chem
.
2022
;
414
(
3
):
1347
57
.
28.
Camp
CH
,
Yegnanarayanan
S
,
Eftekhar
AA
,
Adibi
A
.
Label-free flow cytometry using Multiplex Coherent Anti-Stokes Raman scattering (MCARS) for the analysis of biological specimens
.
Opt Lett
.
2011
;
36
:
2309
11
.
29.
Camp
CH
,
Cicerone
MT
.
Chemically sensitive bioimaging with coherent Raman scattering
.
Nat Photon
.
2015
;
9
(
5
):
295
305
.
30.
Campana
S
,
Busch
K
,
Hentschel
U
,
Muyzer
G
,
de Goeij
JM
.
DNA-Stable Isotope Probing (DNA-SIP) identifies marine sponge-associated bacteria actively utilizing Dissolved Organic Matter (DOM)
.
Environ Microbiol
.
2021
;
23
(
8
):
4489
504
.
31.
Cañeque
T
,
Müller
S
,
Rodriguez
R
.
Visualizing biologically active small molecules in cells using click chemistry
.
Nat Rev Chem
.
2018
;
2
(
9
):
202
15
.
32.
Chen
L
,
Zhao
B
,
Li
X
,
Cheng
Z
,
Wu
R
,
Xia
Y
.
Isolating and characterizing translationally active fraction of anammox microbiota using bioorthogonal non-canonical amino acid tagging
.
Chem Eng J
.
2021
;
418
:
129411
.
33.
Cho
JC
.
Omics-based microbiome analysis in microbial ecology: from sequences to information
.
J Microbiol
.
2021
;
59
(
3
):
229
32
.
34.
Cialla-May
D
,
Krafft
C
,
Rösch
P
,
Deckert-Gaudig
T
,
Frosch
T
,
Jahn
IJ
.
Raman spectroscopy and imaging in bioanalytics
.
Anal Chem
.
2022
;
94
(
1
):
86
119
.
35.
Clingenpeel
S
,
Schwientek
P
,
Hugenholtz
P
,
Woyke
T
.
Effects of sample treatments on genome recovery via single-cell genomics
.
ISME J
.
2014
;
8
(
12
):
2546
9
.
36.
Coyotzi
S
,
Pratscher
J
,
Murrell
JC
,
Neufeld
JD
.
Targeted metagenomics of active microbial populations with stable-isotope probing
.
Curr Opin Biotechnol
.
2016
;
41
:
1
8
.
37.
Cross
KL
,
Campbell
JH
,
Balachandran
M
,
Campbell
AG
,
Cooper
CJ
,
Griffen
A
.
Targeted isolation and cultivation of uncultivated bacteria by reverse genomics
.
Nat Biotechnol
.
2019
;
37
(
11
):
1314
21
.
38.
Cui
L
,
Chen
S
,
Zhang
K
.
Effect of toxicity of Ag nanoparticles on SERS spectral variance of bacteria
.
Spectrochim Acta A Mol Biomol Spectrosc
.
2015
;
137
:
1061
6
.
39.
Cui
L
,
Li
R
,
Mu
T
,
Wang
J
,
Zhang
W
,
Sun
M
.
In situ Plasmon-Enhanced CARS and TPEF for Gram staining identification of non-fluorescent bacteria
.
Spectrochim Acta A Mol Biomol Spectrosc
.
2022
;
264
:
120283
.
40.
Cui
L
,
Zhang
DD
,
Yang
K
,
Zhang
X
,
Zhu
YG
.
Perspective on surface-enhanced Raman spectroscopic investigation of microbial world
.
Anal Chem
.
2019
;
91
(
24
):
15345
54
.
41.
Dam
HT
,
Vollmers
J
,
Sobol
MS
,
Cabezas
A
,
Kaster
AK
.
Targeted cell sorting combined with single cell genomics captures low abundant microbial dark matter with higher sensitivity than metagenomics
.
Front Microbiol
.
2020
;
11
:
1377
.
42.
Dao
TNT
,
Lee
EY
,
Koo
B
,
Jin
CE
,
Lee
TY
,
Shin
Y
.
A microfluidic enrichment platform with a recombinase polymerase amplification sensor for pathogen diagnosis
.
Anal Biochem
.
2018
;
544
:
87
92
.
43.
DeMeester
KE
,
Liang
H
,
Jensen
MR
,
Jones
ZS
,
D’Ambrosio
EA
,
Scinto
SL
.
Synthesis of functionalized N-acetyl muramic acids to probe bacterial cell wall recycling and biosynthesis
.
J Am Chem Soc
.
2018
;
140
(
30
):
9458
65
.
44.
DeMeester
KE
,
Liang
H
,
Zhou
J
,
Wodzanowski
KA
,
Prather
BL
,
Santiago
CC
.
Metabolic incorporation of N-acetyl muramic acid probes into bacterial peptidoglycan
.
Curr Protoc Chem Biol
.
2019
;
11
(
4
):
e74
.
45.
Dieterich
DC
,
Link
AJ
,
Graumann
J
,
Tirrell
DA
,
Schuman
EM
.
Selective identification of newly synthesized proteins in mammalian cells using Bioorthogonal Noncanonical Amino Acid Tagging (BONCAT)
.
Proc Natl Acad Sci U S A
.
2006
;
103
(
25
):
9482
7
.
46.
Doud
DFR
,
Bowers
RM
,
Schulz
F
,
De Raad
M
,
Deng
K
,
Tarver
A
.
Function-driven single-cell genomics uncovers cellulose-degrading bacteria from the rare biosphere
.
ISME J
.
2020
;
14
(
3
):
659
75
.
47.
Doud
DFR
,
Woyke
T
.
Novel approaches in function-driven single-cell genomics
.
FEMS Microbiol Rev
.
2017
;
41
(
4
):
538
48
.
48.
Douglas
MP
,
Rogers
SO
.
DNA damage caused by common cytological fixatives
.
Mutat Res
.
1998
401
1-2
77
88
.
49.
Du
Z
,
Behrens
SF
.
Tracking de novo protein synthesis in the activated sludge microbiome using BONCAT-FACS
.
Water Res
.
2021
;
205
:
117696
.
50.
Eickhorst
T
,
Tippkötter
R
.
Improved detection of soil microorganisms using Fluorescence In Situ Hybridization (FISH) and catalyzed reporter deposition (CARD-FISH)
.
Soil Biol Biochem
.
2008
;
40
(
7
):
1883
91
.
51.
Fazi
S
,
Amalfitano
S
,
Pizzetti
I
,
Pernthaler
J
.
Efficiency of fluorescence in situ hybridization for bacterial cell identification in temporary river sediments with contrasting water content
.
Syst Appl Microbiol
.
2007
;
30
(
6
):
463
70
.
52.
Fleischmann
M
,
Hendra
PJ
,
McQuillan
AJ
.
Raman spectra of pyridine adsorbed at a silver electrode
.
Chem Phys Lett
.
1974
;
26
(
2
):
163
6
.
53.
Frias-Lopez
J
,
Shi
Y
,
Tyson
GW
,
Coleman
ML
,
Schuster
SC
,
Chisholm
SW
.
Microbial community gene expression in ocean surface waters
.
Proc Natl Acad Sci U S A
.
2008
;
105
(
10
):
3805
10
.
54.
Frickmann
H
,
Zautner
AE
,
Moter
A
,
Kikhney
J
,
Hagen
RM
,
Stender
H
.
Fluorescence In Situ Hybridization (FISH) in the microbiological diagnostic routine laboratory: a review
.
Crit Rev Microbiol
.
2017
;
43
(
3
):
263
93
.
55.
Fu
Y
,
Wang
H
,
Shi
R
,
Cheng
J-X
.
Characterization of photodamage in coherent anti-Stokes Raman scattering microscopy
.
Opt Express
.
2006
;
14
(
9
):
3942
51
.
56.
Fugier
E
,
Dumont
A
,
Malleron
A
,
Poquet
E
,
Mas Pons
J
,
Baron
A
.
Rapid and specific enrichment of culturable gram negative bacteria using non-lethal copper-free click chemistry coupled with magnetic beads separation
.
PLoS One
.
2015
;
10
(
6
):
e0127700
.
57.
Gala De Pablo
J
,
Lindley
M
,
Hiramatsu
K
,
Goda
K
.
High-throughput Raman flow cytometry and beyond
.
Acc Chem Res
.
2021
;
54
(
9
):
2132
43
.
58.
Gao
D
,
Huang
X
,
Tao
Y
.
A critical review of NanoSIMS in analysis of microbial metabolic activities at single-cell level
.
Crit Rev Biotechnol
.
2016
;
36
(
5
):
884
90
.
59.
Gates
SD
,
Condit
RC
,
Moussatche
N
,
Stewart
BJ
,
Malkin
AJ
,
Weber
PK
.
High initial sputter rate found for vaccinia virions using isotopic labeling, NanoSIMS, and AFM
.
Anal Chem
.
2018
;
90
(
3
):
1613
20
.
60.
Givan
AL
.
Flow cytometry: an introduction
.
Methods Mol Biol
.
2011
;
699
:
1
29
.
61.
Grieb
A
,
Bowers
RM
,
Oggerin
M
,
Goudeau
D
,
Lee
J
,
Malmstrom
RR
.
A pipeline for targeted metagenomics of environmental bacteria
.
Microbiome
.
2020
;
8
:
21
17
.
62.
Gross
A
,
Schoendube
J
,
Zimmermann
S
,
Steeb
M
,
Zengerle
R
,
Koltay
P
.
Technologies for single-cell isolation
.
Int J Mol Sci
.
2015
;
16
(
8
):
16897
919
.
63.
Gross
A
,
Schöndube
J
,
Niekrawitz
S
,
Streule
W
,
Riegger
L
,
Zengerle
R
.
Single-cell printer: automated, on demand, and label free
.
J Lab Autom
.
2013
;
18
(
6
):
504
18
.
64.
Grujcic
V
,
Taylor
GT
,
Foster
RA
.
One cell at a time: advances in single-cell methods and instrumentation for discovery in aquatic microbiology
.
Front Microbiol
.
2022
;
13
:
1747
.
65.
Han
Z
,
Thuy-Boun
PS
,
Pfeiffer
W
,
Vartabedian
VF
,
Torkamani
A
,
Teijaro
JR
.
Identification of an N-acetylneuraminic acid-presenting bacteria isolated from a human microbiome
.
Sci Rep
.
2021
;
11
:
4763
12
.
66.
Haroon
MF
,
Skennerton
CT
,
Steen
JA
,
Lachner
N
,
Hugenholtz
P
,
Tyson
GW
.
In-solution fluorescence in situ hybridization and Fluorescence-activated cell sorting for single cell and population genome recovery
.
Methods in Enzymology
Academic Press
2013
. p.
3
19
.
67.
Harris
RL
,
Vetter
MCYL
,
van Heerden
E
,
Cason
E
,
Vermeulen
JG
,
Taneja
A
.
FISH-TAMB, a fixation-free mRNA fluorescent labeling technique to target transcriptionally active members in microbial communities
.
Microb Ecol
.
2022
;
84
(
1
):
182
97
.
68.
Harrison
JP
,
Berry
D
.
Vibrational spectroscopy for imaging single microbial cells in complex biological samples
.
Front Microbiol
.
2017
;
8
:
675
.
69.
Hatzenpichler
R
,
Connon
SA
,
Goudeau
D
,
Malmstrom
RR
,
Woyke
T
,
Orphan
VJ
.
Visualizing in situ translational activity for identifying and sorting slow-growing archaeal - bacterial consortia
.
Proc Natl Acad Sci U S A
.
2016
113
28
E4069
E4078
.
70.
Hatzenpichler
R
,
Krukenberg
V
,
Spietz
RL
,
Jay
ZJ
.
Next-generation physiology approaches to study microbiome function at single cell level
.
Nat Rev Microbiol
.
2020
;
18
(
4
):
241
56
.
71.
Hatzenpichler
R
,
Scheller
S
,
Tavormina
PL
,
Babin
BM
,
Tirrell
DA
,
Orphan
VJ
.
In situ visualization of newly synthesized proteins in environmental microbes using amino acid tagging and click chemistry
.
Environ Microbiol
.
2014
;
16
(
8
):
2568
90
.
72.
He
X
,
Chadwick
GL
,
Kempes
CP
,
Orphan
VJ
,
Meile
C
.
Controls on interspecies electron transport and size limitation of anaerobically methane-oxidizing microbial consortia
.
mBio
.
2021
12
3
e03620-20
.
73.
He
Y
,
Wang
X
,
Ma
B
,
Xu
J
.
Ramanome technology platform for label-free screening and sorting of microbial cell factories at single-cell resolution
.
Biotechnol Adv
.
2019
;
37
(
6
):
107388
.
74.
Hiramatsu
K
,
Ideguchi
T
,
Yonamine
Y
,
Lee
SW
,
Luo
Y
,
Hashimoto
K
.
High-throughput label-free molecular fingerprinting flow cytometry
.
Sci Adv
.
2019
5
1
eaau0241
257
.
75.
Hiramatsu
K
,
Yamada
K
,
Lindley
M
,
Suzuki
K
,
Goda
K
.
Large-scale label-free single-cell analysis of paramylon in Euglena gracilis by high-throughput broadband Raman flow cytometry
.
Biomed Opt Express
.
2020
;
11
(
4
):
1752
9
.
76.
Holm
A
,
Wu
W
,
Lund-Johansen
F
.
Antibody array analysis of labelled proteomes: how should we control specificity
.
N Biotechnol
.
2012
;
29
(
5
):
578
85
.
77.
Hopp
B
,
Smausz
T
,
Kresz
N
,
Barna
N
,
Bor
Z
,
Kolozsvári
L
.
Survival and proliferative ability of various living cell types after laser-induced forward transfer
.
Tissue Eng
.
2005
11
11-12
1817
23
.
78.
Hu
B
,
Xu
B
,
Yun
J
,
Wang
J
,
Xie
B
,
Li
C
.
High-throughput single-cell cultivation reveals the underexplored rare biosphere in deep-sea sediments along the Southwest Indian Ridge
.
Lab Chip
.
2020
;
20
(
2
):
363
72
.
79.
Huang
K-C
,
Kuang
S
,
Robinson
JP
,
Cheng
J-X
,
Liao
C
,
Zhang
C
.
Stimulated Raman scattering flow cytometry for label-free single-particle analysis
.
Opt
.
2017
4
Issue 1
103
9
.
80.
Hussain
W
,
Ullah
MW
,
Farooq
U
,
Aziz
A
,
Wang
S
.
Bacteriophage-based advanced bacterial detection: concept, mechanisms, and applications
.
Biosens Bioelectron
.
2021
;
177
:
112973
.
81.
Ishizuka
T
,
Liu
HS
,
Ito
K
,
Xu
Y
.
Fluorescence imaging of chromosomal DNA using click chemistry
.
Sci Rep
.
2016
;
6
:
33217
0
.
82.
Jayan
H
,
Pu
H
,
Sun
DW
.
Recent developments in Raman spectral analysis of microbial single cells: techniques and applications
.
Crit Rev Food Sci Nutr
.
2022
;
62
(
16
):
4294
308
.
83.
Jeanmaire
DL
,
Van Duyne
RP
.
Surface Raman spectroelectrochemistry
.
J Electroanal Chem
.
1977
;
84
:
1
20
.
84.
Jeyavani
J
,
Sibiya
A
,
Shanthini
S
,
Ravi
C
,
Vijayakumar
S
,
Rajan
DK
.
A review on aquatic impacts of microplastics and its bioremediation aspects
.
Curr Pollut Rep
.
2021
;
7
(
3
):
286
99
.
85.
Jia
Z
,
Dong
Y
,
Xu
H
,
Wang
F
.
Optimizing the hybridization chain reaction-fluorescence in situ hybridization (HCR-FISH) protocol for detection of microbes in sediments
.
Mar Life Sci Technol
.
2021
;
3
(
4
):
529
41
.
86.
Jiao
JY
,
Liu
L
,
Hua
ZS
,
Fang
BZ
,
Zhou
EM
,
Salam
N
.
Microbial dark matter coming to light: challenges and opportunities
.
Natl Sci Rev
.
2021
8
3
nwaa280
.
87.
Jing
X
,
Gong
Y
,
Xu
T
,
Meng
Y
,
Han
X
,
Su
X
.
One-cell metabolic phenotyping and sequencing of soil microbiome by Raman-activated gravity-driven encapsulation (RAGE)
.
mSystems
.
2021
;
6
(
3
):
e0018121
21
.
88.
Kang
SM
,
de Josselin de Jong
E
,
Higham
SM
,
Hope
CK
,
Kim
BI
.
Fluorescence fingerprints of oral bacteria
.
J Biophotonics
.
2020
;
13
(
1
):
e201900190
.
89.
Kanno
N
,
Kato
S
,
Ohkuma
M
,
Matsui
M
,
Iwasaki
W
,
Shigeto
S
.
Machine learning-assisted single-cell Raman fingerprinting for in situ and nondestructive classification of prokaryotes
.
iScience
.
2021
;
24
(
9
):
102975
.
90.
Kaster
AK
,
Sobol
MS
.
Microbial single-cell omics: the crux of the matter
.
Appl Microbiol Biotechnol
.
2020
;
104
(
19
):
8209
20
.
91.
Kaushik
S
,
Alatawi
A
,
Djiwanti
SR
,
Pande
A
,
Skotti
E
,
Soni
V
.
Potential of extremophiles for bioremediation
. In:
Panpatte
DG
,
Jhala
YK
, editors.
Microbial rejuvenation of polluted environment
Singapore
Springer
2021
Vol. 1
. p.
293
328
.
92.
Kiick
KL
,
Saxon
E
,
Tirrell
DA
,
Bertozzi
CR
.
Incorporation of azides into recombinant proteins for chemoselective modification by the Staudinger ligation
.
Proc Natl Acad Sci U S A
.
2002
;
99
(
1
):
19
24
.
93.
Kilburn
MR
,
Wacey
D
.
Nanoscale Secondary Ion Mass Spectrometry (NanoSIMS) as an analytical tool in the geosciences
.
RSC Detect Sci
.
2014
2015-January
1
34
.
94.
Kitzinger
K
,
Padilla
CC
,
Marchant
HK
,
Hach
PF
,
Herbold
CW
,
Kidane
AT
.
Cyanate and urea are substrates for nitrification by Thaumarchaeota in the marine environment
.
Nat Microbiol
.
2019
;
4
(
2
):
234
43
.
95.
Kolb
HC
,
Finn
MG
,
Sharpless
KB
.
Click chemistry: diverse chemical function from a few good reactions
.
Angew Chem Int Ed
.
2001
;
40
(
11
):
2004
21
.
96.
Kramer
A
,
Keitel
T
,
Winkler
K
,
Stöcklein
W
,
Höhne
W
,
Schneider-Mergener
J
.
Molecular basis for the binding promiscuity of an anti-p24 (HIV-1) monoclonal antibody
.
Cell
.
1997
;
91
(
6
):
799
809
.
97.
Kubota
K
.
CARD-FISH for environmental microorganisms: technical advancement and future applications
.
Microbes Environ
.
2013
;
28
(
1
):
3
12
.
98.
Kumar
PS
,
Dabdoub
SM
,
Ganesan
SM
.
Probing periodontal microbial dark matter using metataxonomics and metagenomics
.
Periodontol
.
2021
;
85
(
1
):
12
27
.
99.
Landegren
U
,
Vänelid
J
,
Hammond
M
,
Nong
RY
,
Wu
D
,
Ulleräs
E
.
Opportunities for sensitive plasma proteome analysis
.
Anal Chem
.
2012
;
84
(
4
):
1824
30
.
100.
Lawrence
D
,
Campbell
DE
,
Schriefer
LA
,
Rodgers
R
,
Walker
FC
,
Turkin
M
.
Single-cell genomics for resolution of conserved bacterial genes and mobile genetic elements of the human intestinal microbiota using flow cytometry
.
Gut Microbes
.
2022
;
14
(
1
):
e2029673
.
101.
Le
TT
,
Yue
S
,
Cheng
JX
.
Shedding new light on lipid biology with coherent anti-Stokes Raman scattering microscopy
.
J Lipid Res
.
2010
;
51
(
11
):
3091
102
.
102.
Le Ru
ECL
,
Meyer
SA
,
Artur
C
,
Etchegoin
PG
,
Grand
J
,
Lang
P
.
Experimental demonstration of surface selection rules for SERS on flat metallic surfaces
.
Chem Commun
.
2011
;
47
(
13
):
3903
5
.
103.
Lebaron
P
,
Catala
P
,
Fajon
C
,
Joux
F
,
Baudart
J
,
Bernard
L
.
A new sensitive, whole-cell hybridization technique for detection of bacteria involving a biotinylated oligonucleotide probe targeting rRNA and tyramide signal amplification
.
Appl Environ Microbiol
.
1997
;
63
(
8
):
3274
8
.
104.
Lee
KS
,
Palatinszky
M
,
Pereira
FC
,
Nguyen
J
,
Fernandez
VI
,
Mueller
AJ
.
An automated Raman-based platform for the sorting of live cells by functional properties
.
Nat Microbiol
.
2019a
4
6
1035
48
.
105.
Lee
KS
,
Pereira
FC
,
Palatinszky
M
,
Behrendt
L
,
Alcolombri
U
,
Berry
D
.
Optofluidic Raman-activated cell sorting for targeted genome retrieval or cultivation of microbial cells with specific functions
.
Nat Protoc
.
2021
;
16
(
2
):
634
76
.
106.
Lee
N
,
Nielsen
PH
,
Andreasen
KH
,
Juretschko
S
,
Nielsen
JL
,
Schleifer
KH
.
Combination of fluorescent in situ hybridization and microautoradiography: a new tool for structure-function analyses in microbial ecology
.
Appl Environ Microbiol
.
1999
;
65
(
3
):
1289
97
.
107.
Lee
SW
,
Kim
BW
,
Shin
HS
,
Go
A
,
Lee
MH
,
Lee
DK
.
Aptamer affinity-bead mediated capture and displacement of gram-negative bacteria using acoustophoresis
.
Micromachines
.
2019b
10
11
770
.
108.
Lehner
F
,
Kudlinzki
D
,
Richter
C
,
Müller-Werkmeister
HM
,
Eberl
KB
,
Bredenbeck
J
.
Impact of azidohomoalanine incorporation on protein structure and ligand binding
.
Chembiochem
.
2017
;
18
(
23
):
2340
50
.
109.
Leizeaga
A
,
Estrany
M
,
Forn
I
,
Sebastián
M
.
Using click-chemistry for visualizing in situ changes of translational activity in planktonic marine bacteria
.
Front Microbiol
.
2017
;
8
:
2360
.
110.
Li
J
,
Zhang
D
,
Li
B
,
Luo
C
,
Zhang
G
.
Identifying the active phenanthrene degraders and characterizing their metabolic activities at the single-cell level by the combination of magnetic-nanoparticle-mediated isolation, stable-isotope probing, and Raman-activated cell sorting (MMI-SIP-RACS)
.
Environ Sci Technol
.
2022
;
56
(
4
):
2289
99
.
111.
Li
R
,
Yu
H
,
Chen
X
.
Recent progress in chemical synthesis of bacterial surface glycans
.
Curr Opin Chem Biol
.
2020
;
58
:
121
36
.
112.
Liang
H
,
DeMeester
KE
,
Hou
CW
,
Parent
MA
,
Caplan
JL
,
Grimes
CL
.
Metabolic labelling of the carbohydrate core in bacterial peptidoglycan and its applications
.
Nat Commun
.
2017
;
8
:
15015
1
.
113.
Lindivat
M
,
Larsen
A
,
Hess-Erga
OK
,
Bratbak
G
,
Hoell
IA
.
Bioorthogonal non-canonical amino acid tagging combined with flow cytometry for determination of activity in aquatic microorganisms
.
Front Microbiol
.
2020
;
11
:
1929
.
114.
Liu
S
,
Li
H
,
Hassan
MM
,
Ali
S
,
Chen
Q
.
SERS based artificial peroxidase enzyme regulated multiple signal amplified system for quantitative detection of foodborne pathogens
.
Food Control
.
2021
;
123
:
107733
.
115.
Lohumi
S
,
Kim
MS
,
Qin
J
,
Cho
BK
.
Raman imaging from microscopy to macroscopy: quality and safety control of biological materials
.
Trac Trends Anal Chem
.
2017
;
93
:
183
98
.
116.
Lok
C
.
Mining the microbial dark matter
.
Nature
.
2015
;
522
(
7556
):
270
3
.
117.
Luo
Y
,
Liu
F
,
Li
E
,
Fang
Y
,
Zhao
G
,
Dai
X
.
FRET-based fluorescent nanoprobe platform for sorting of active microorganisms by functional properties
.
Biosens Bioelectron
.
2020
;
148
:
111832
.
118.
Matanfack
GA
,
Taubert
M
,
Guo
S
,
Houhou
R
,
Bocklitz
T
,
Küsel
K
.
Influence of carbon sources on quantification of deuterium incorporation in heterotrophic bacteria: a Raman-stable isotope labeling approach
.
Anal Chem
.
2020
;
92
(
16
):
11429
37
.
119.
McGlynn
SE
,
Chadwick
GL
,
Kempes
CP
,
Orphan
VJ
.
Single cell activity reveals direct electron transfer in methanotrophic consortia
.
Nature
.
2015
;
526
(
7574
):
531
5
.
120.
McKinnon
KM
.
Flow cytometry: an overview
.
Curr Protoc Immunol
.
2018
120
1
5.1.1
11
.
121.
McLean
JS
,
Lombardo
MJ
,
Badger
JH
,
Edlund
A
,
Novotny
M
,
Yee-Greenbaum
J
.
Candidate phylum TM6 genome recovered from a hospital sink biofilm provides genomic insights into this uncultivated phylum
.
Proc Natl Acad Sci U S A
.
2013
110
26
E2390
9
.
122.
Merkel
TC
,
Bondar
VI
,
Nagai
K
,
Freeman
BD
,
Pinnau
I
.
Gas sorption, diffusion, and permeation in poly(dimethylsiloxane)
.
J Polym Sci B Polym Phys
.
2000
;
38
(
3
):
415
34
.
123.
Moor
K
,
Fadlallah
J
,
Toska
A
,
Sterlin
D
,
Balmer
ML
,
Macpherson
AJ
.
Analysis of bacterial-surface-specific antibodies in body fluids using bacterial flow cytometry
.
Nat Protoc
.
2016
;
11
(
8
):
1531
53
.
124.
Moses
JE
,
Moorhouse
AD
.
The growing applications of click chemistry
.
Chem Soc Rev
.
2007
;
36
(
8
):
1249
62
.
125.
Moter
A
,
Göbel
UB
.
Fluorescence in situ hybridization (FISH) for direct visualization of microorganisms
.
J Microbiol Methods
.
2000
;
41
(
2
):
85
112
.
126.
Muhamadali
H
,
Chisanga
M
,
Subaihi
A
,
Goodacre
R
.
Combining Raman and FT-IR spectroscopy with quantitative isotopic labeling for differentiation of E. Coli cells at community and single cell levels
.
Anal Chem
.
2015
;
87
(
8
):
4578
86
.
127.
Müller
TG
,
Sakin
V
,
Müller
B
.
A spotlight on viruses—application of click chemistry to visualize virus-cell interactions
.
Molecules
.
2019
;
24
(
3
):
481
.
128.
Mullis
MM
,
Rambo
IM
,
Baker
BJ
,
Reese
BK
.
Diversity, ecology, and prevalence of antimicrobials in nature
.
Front Microbiol
.
2019
;
10
:
2518
.
129.
Münch
JM
,
Sobol
MS
,
Brors
B
,
Kaster
A-K
Single-cell Transcriptomics and Data Analyses for Prokaryotes—Past
present and future conceptsAcademic Press
2023
.
130.
Musat
N
,
Musat
F
,
Weber
PK
,
Pett-Ridge
J
.
Pett-Ridge J: tracking microbial interactions with NanoSIMS
.
Curr Opin Biotechnol
.
2016
;
41
:
114
21
.
131.
Neuenschwander
SM
,
Salcher
MM
,
Pernthaler
J
.
Fluorescence in situ hybridization and sequential catalyzed reporter deposition (2C-FISH) for the flow cytometric sorting of freshwater ultramicrobacteria
.
Front Microbiol
.
2015
;
6
:
247
.
132.
Nielsen
JL
,
Nielsen
PH
.
Advances in microscopy: microautoradiography of single cells
.
Methods Enzymol
.
2005
;
397
:
237
56
.
133.
Nissen
JN
,
Johansen
J
,
Allesøe
RL
,
Sønderby
CK
,
Armenteros
JJA
,
Grønbech
CH
.
Improved metagenome binning and assembly using deep variational autoencoders
.
Nat Biotechnol
.
2021
;
39
(
5
):
555
60
.
134.
Noriea
NF
,
Johnson
CN
,
Griffitt
KJ
,
Grimes
DJ
.
Distribution of type III secretion systems in Vibrio parahaemolyticus from the northern Gulf of Mexico
.
J Appl Microbiol
.
2010
;
109
(
3
):
953
62
.
135.
Novais
Â
,
Peixe
L
.
Fourier transform infrared spectroscopy (FT-IR) for food and water microbiology
.
Appl Integr Omi Diagn Clin Public Heal Microbiol
.
2021
191
217
.
136.
Nuñez
J
,
Renslow
R
,
Cliff
JB
,
Anderton
CR
.
NanoSIMS for biological applications: current practices and analyses
.
Biointerphases
.
2018
13
3
03B301
.
137.
Oehler
D
,
Cady
S
.
Biogenicity and syngeneity of organic matter in ancient sedimentary rocks: recent advances in the search for evidence of past life
.
Challenges
.
2014
;
5
(
2
):
260
83
.
138.
Orphan
VJ
,
House
CH
.
Geobiological investigations using secondary ion mass spectrometry: microanalysis of extant and paleo-microbial processes
.
Geobiology
.
2009
;
7
(
3
):
360
72
.
139.
Ouverney
CC
,
Fuhrman
JA
.
Combined microautoradiography-16S rRNA probe technique for determination of radioisotope uptake by specific microbial cell types in situ
.
Appl Environ Microbiol
.
1999
;
65
(
4
):
1746
52
.
140.
Pal
M
,
Lee
S
,
Kwon
D
,
Hwang
J
,
Lee
H
,
Hwang
S
.
Direct immobilization of antibodies on Zn-doped Fe34 nanoclusters for detection of pathogenic bacteria
.
Anal Chim Acta
.
2017
;
952
:
81
7
.
141.
Parras-Moltó
M
,
Rodríguez-Galet
A
,
Suárez-Rodríguez
P
,
López-Bueno
A
.
Evaluation of bias induced by viral enrichment and random amplification protocols in metagenomic surveys of saliva DNA viruses
.
Microbiome
.
2018
;
6
:
119
8
.
142.
Pereira
AC
,
Tenreiro
A
,
Cunha
MV
.
When FLOW-FISH met FACS: combining multiparametric, dynamic approaches for microbial single-cell research in the total environment
.
Sci Total Environ
.
2022
806
Pt 2
150682
.
143.
Pernthaler
A
,
Pernthaler
J
,
Amann
R
.
Fluorescence in situ hybridization and catalyzed reporter deposition for the identification of marine bacteria
.
Appl Environ Microbiol
.
2002
;
68
(
6
):
3094
101
.
144.
Pernthaler
J
,
Glöckner
FO
,
Schönhuber
W
,
Amann
R
.
Fluorescence In Situ Hybridization (FISH) with rRNA-targeted oligonucleotide probes
.
Methods Microbiol
.
2001
;
30
:
207
26
.
145.
Petrov
GI
,
Yakovlev
VV
,
Sokolov
AV
,
Scully
MO
.
Detection of Bacillus subtilis spores in water by means of broadband coherent anti-Stokes Raman spectroscopy
.
Opt Express
.
2005
;
13
(
23
):
9537
42
.
146.
Phelan
J
,
Altharawi
A
,
Chan
KLA
.
Tracking glycosylation in live cells using FTIR spectroscopy
.
Talanta
.
2020
;
211
:
120737
.
147.
Pilo-Pais
M
,
Watson
A
,
Demers
S
,
Labean
TH
,
Finkelstein
G
.
Surface-enhanced Raman scattering plasmonic enhancement using DNA origami-based complex metallic nanostructures
.
Nano Lett
.
2014
;
14
(
4
):
2099
104
.
148.
Pratscher
J
,
Dumont
MG
,
Conrad
R
.
Assimilation of acetate by the putative atmospheric methane oxidizers belonging to the USCα clade
.
Environ Microbiol
.
2011
;
13
(
10
):
2692
701
.
149.
Pratscher
J
,
Vollmers
J
,
Wiegand
S
,
Dumont
MG
,
Kaster
AK
.
Unravelling the identity, metabolic potential and global biogeography of the atmospheric methane-oxidizing upland soil cluster α
.
Environ Microbiol
.
2018
;
20
(
3
):
1016
29
.
150.
Quintelas
C
,
Ferreira
EC
,
Lopes
JA
,
Sousa
C
.
An overview of the evolution of infrared spectroscopy applied to bacterial typing
.
Biotechnol J
.
2018
;
13
(
1
):
1700449
.
151.
Raman
CV
,
Krishnan
KS
.
A new type of secondary radiation
.
Nature
.
1928
;
121
(
3048
):
501
2
.
152.
Rampelotto
PH
.
Extremophiles and extreme environments
.
Life
.
2013
;
3
:
482
5
.
153.
Reed
CJ
,
Lewis
H
,
Trejo
E
,
Winston
V
,
Evilia
C
.
Protein adaptations in archaeal extremophiles
.
Archaea
.
2013
;
2013
:
373275
.
154.
Reichart
NJ
,
Jay
ZJ
,
Krukenberg
V
,
Parker
AE
,
Spietz
RL
,
Hatzenpichler
R
.
Activity-based cell sorting reveals responses of uncultured archaea and bacteria to substrate amendment
.
ISME J
.
2020
;
14
(
11
):
2851
61
.
155.
Riba
J
,
Gleichmann
T
,
Zimmermann
S
,
Zengerle
R
,
Koltay
P
.
Label-free isolation and deposition of single bacterial cells from heterogeneous samples for clonal culturing
.
Sci Rep
.
2016
;
6
:
32837
9
.
156.
Rinke
C
,
Lee
J
,
Nath
N
,
Goudeau
D
,
Thompson
B
,
Poulton
N
.
Obtaining genomes from uncultivated environmental microorganisms using FACS-based single-cell genomics
.
Nat Protoc
.
2014
;
9
(
5
):
1038
48
.
157.
Salcher
MM
,
Andrei
,
Bulzu
P-A
,
Keresztes
ZG
,
Banciu
HL
,
Ghai
R
.
Visualization of lokiarchaeia and heimdallarchaeia (asgardarchaeota) by fluorescence in situ hybridization and Catalyzed Reporter Deposition (CARD-FISH)
.
mSphere
.
2020
5
4
e00686-20
.
158.
Samo
TJ
,
Smriga
S
,
Malfatti
F
,
Sherwood
BP
,
Azam
F
.
Broad distribution and high proportion of protein synthesis active marine bacteria revealed by click chemistry at the single cell level
.
Front Mar Sci
.
2014
;
1
:
48
.
159.
Schimak
MP
,
Kleiner
M
,
Wetzel
S
,
Liebeke
M
,
Dubilier
N
,
Fuchs
BM
.
MiL-FISH: multilabeled oligonucleotides for fluorescence in situ hybridization improve visualization of bacterial cells
.
Appl Environ Microbiol
.
2016
;
82
(
1
):
62
70
.
160.
Schönhuber
W
,
Fuchs
B
,
Juretschko
S
,
Amann
R
.
Improved sensitivity of whole-cell hybridization by the combination of horseradish peroxidase-labeled oligonucleotides and tyramide signal amplification
.
Appl Environ Microbiol
.
1997
;
63
(
8
):
3268
73
.
161.
Schwank
K
,
Bornemann
TLV
,
Dombrowski
N
,
Spang
A
,
Banfield
JF
,
Probst
AJ
.
An archaeal symbiont-host association from the deep terrestrial subsurface
.
ISME J
.
2019
;
13
(
8
):
2135
9
.
162.
Sekar
R
,
Fuchs
BM
,
Amann
R
,
Pernthaler
J
.
Flow sorting of marine bacterioplankton after fluorescence in situ hybridization
.
Appl Environ Microbiol
.
2004
;
70
(
10
):
6210
9
.
163.
Sobol
MS
,
Kaster
AK
.
Back to basics: a simplified improvement to multiple displacement amplification for microbial single-cell genomics
.
Int J Mol Sci
.
2023
;
24
(
5
):
4270
.
164.
Song
YY
,
Kaster
AK
,
Vollmers
J
,
Song
YY
,
Davison
PA
,
Frentrup
M
.
Single-cell genomics based on Raman sorting reveals novel carotenoid-containing bacteria in the Red Sea
.
Microb Biotechnol
.
2017
;
10
(
1
):
125
37
.
165.
Spitsyn
AN
,
Utkin
DV
,
Kuznetsov
OS
,
Erokhin
PS
,
Osina
NA
,
Kochubei
VI
.
Application of optical techniques to investigation and identification of microorganisms: a review
.
Opt Spectrosc
.
2021
;
129
(
1
):
135
48
.
166.
Sportelli
MC
,
Kranz
C
,
Mizaikoff
B
,
Cioffi
N
.
Recent advances on the spectroscopic characterization of microbial biofilms: a critical review
.
Anal Chim Acta
.
2022
;
1195
:
339433
.
167.
Staley
JT
,
Konopka
A
.
Measurement of in situ activities of nonphotosynthetic microorganisms in aquatic and terrestrial habitats
.
Annu Rev Microbiol
.
1985
;
39
:
321
46
.
168.
Starr
EP
,
Shi
S
,
Blazewicz
SJ
,
Koch
BJ
,
Probst
AJ
,
Hungate
BA
.
Stable-isotope-informed, genome-resolved metagenomics uncovers potential cross-kingdom interactions in rhizosphere soil
.
mSphere
.
2021
;
6
(
5
):
e0008521
.
169.
Steward
KF
,
Eilers
B
,
Tripet
B
,
Fuchs
A
,
Dorle
M
,
Rawle
R
.
Metabolic implications of using BioOrthogonal Non-Canonical Amino Acid Tagging (BONCAT) for tracking protein synthesis
.
Front Microbiol
.
2020
;
11
:
197
.
170.
Stewart
S
,
Priore
RJ
,
Nelson
MP
,
Treado
PJ
.
Raman imaging
.
Annu Rev Anal Chem
.
2012
;
5
:
337
60
.
171.
Stoecker
K
,
Dorninger
C
,
Daims
H
,
Wagner
M
.
Double labeling of Oligonucleotide Probes for Fluorescence In Situ Hybridization (DOPE-FISH) improves signal intensity and increases rRNA accessibility
.
Appl Environ Microbiol
.
2010
;
76
(
3
):
922
6
.
172.
Stryhanyuk
H
,
Calabrese
F
,
Kümmel
S
,
Musat
F
,
Richnow
HH
,
Musat
N
.
Calculation of single cell assimilation rates from sip-nanosims-derived isotope ratios: a comprehensive approach
.
Front Microbiol
.
2018
;
9
:
2342
15
.
173.
Su
CH
,
Tsai
MH
,
Lin
CY
,
Ma
YD
,
Wang
CH
,
Chung
YD
.
Dual aptamer assay for detection of Acinetobacter baumannii on an electromagnetically-driven microfluidic platform
.
Biosens Bioelectron
.
2020a
159
112148
.
174.
Su
X
,
Gong
Y
,
Gou
H
,
Jing
X
,
Xu
T
,
Zheng
X
.
Rational optimization of Raman-activated cell ejection and sequencing for bacteria
.
Anal Chem
.
2020b
92
12
8081
9
.
175.
Suzuki
Y
,
Kobayashi
K
,
Wakisaka
Y
,
Deng
D
,
Tanaka
S
,
Huang
CJ
.
Label-free chemical imaging flow cytometry by high-speed multicolor stimulated Raman scattering
.
Proc Natl Acad Sci U S A
.
2019
;
116
(
32
):
15842
8
.
176.
Tada
Y
,
Grossart
HP
.
Community shifts of actively growing lake bacteria after N-acetyl-glucosamine addition: improving the BrdU-FACS method
.
ISME J
.
2014
;
8
(
2
):
441
54
.
177.
Tang
JW
,
Liu
QH
,
Yin
XC
,
Pan
YC
,
Wen
PB
,
Liu
X
.
Comparative analysis of machine learning algorithms on surface enhanced Raman spectra of clinical Staphylococcus species
.
Front Microbiol
.
2021
;
12
:
696921
.
178.
Taniguchi
A
,
Eguchi
M
.
Community structure of actively growing bacteria in a coastal fish-farming area
.
PLoS One
.
2020
;
15
(
6
):
e0235336
.
179.
Täuber
S
,
Golze
C
,
Ho
P
,
Von Lieres
E
,
Grünberger
A
.
dMSCC: a microfluidic platform for microbial single-cell cultivation of: Corynebacterium glutamicum under dynamic environmental medium conditions
.
Lab Chip
.
2020
;
20
(
23
):
4442
55
.
180.
Thompson
AW
,
Crow
MJ
,
Wadey
B
,
Arens
C
,
Turkarslan
S
,
Stolyar
S
.
A method to analyze, sort, and retain viability of obligate anaerobic microorganisms from complex microbial communities
.
J Microbiol Methods
.
2015
;
117
:
74
7
.
181.
Tourkya
B
,
Boubellouta
T
,
Dufour
E
,
Leriche
F
.
Fluorescence spectroscopy as a promising tool for a polyphasic approach to pseudomonad taxonomy
.
Curr Microbiol
.
2009
;
58
(
1
):
39
46
.
182.
Tripathi
A
,
Jabbour
RE
,
Treado
PJ
,
Neiss
JH
,
Nelson
MP
,
Jensen
JL
.
Waterborne pathogen detection using Raman spectroscopy
.
Appl Spectrosc
.
2008
;
62
:
1
9
.
183.
Tu
J
,
Qiao
Y
,
Luo
Y
,
Long
N
,
Lu
Z
.
Quantifying genome DNA during whole-genome amplification: via quantitative real-time multiple displacement amplification
.
RSC Adv
.
2021
;
11
(
8
):
4617
21
.
184.
Tu
Q
,
Chang
C
.
Diagnostic applications of Raman spectroscopy
.
Nanomedicine
.
2012
;
8
(
5
):
545
58
.
185.
Valm
AM
,
Mark Welch
JL
,
Rieken
CW
,
Hasegawa
Y
,
Sogin
ML
,
Oldenbourg
R
.
Systems-level analysis of microbial community organization through combinatorial labeling and spectral imaging
.
Proc Natl Acad Sci U S A
.
2011
;
108
(
10
):
4152
7
.
186.
Veal
DA
,
Deere
D
,
Ferrari
B
,
Piper
J
,
Attfield
PV
.
Fluorescence staining and flow cytometry for monitoring microbial cells
.
J Immunol Methods
.
2000
243
1-2
191
210
.
187.
Vogt
S
,
Löffler
K
,
Dinkelacker
AG
,
Bader
B
,
Autenrieth
IB
,
Peter
S
.
Fourier-Transform Infrared (FTIR) spectroscopy for typing of clinical Enterobacter cloacae complex isolates
.
Front Microbiol
.
2019
;
10
:
2582
.
188.
Volkmer
A
,
Book
LD
,
Xie
XS
.
Time-resolved coherent anti-Stokes Raman scattering microscopy: imaging based on Raman free induction decay
.
Appl Phys Lett
.
2002
;
80
(
9
):
1505
7
.
189.
Vollmers
J
,
Cassal
MC
,
Kaster
A-K
.
Midi-metagenomics”: a novel approach for cultivation independent microbial genome reconstruction from environmental samples
.
bioRxiv
.
2023
;
2023
:
2023
.
190.
Vollmers
J
,
Wiegand
S
,
Lenk
F
,
Kaster
AK
.
How clear is our current view on microbial dark matter? (Re-)assessing public MAG & SAG datasets with MDMcleaner
.
Nucleic Acids Res
.
2022
50
13
E76
.
191.
Wagner
M
,
Horn
M
,
Daims
H
.
Fluorescence in situ hybridisation for the identification and characterisation of prokaryotes
.
Curr Opin Microbiol
.
2003a
6
3
302
9
.
192.
Wagner
M
,
Horn
M
,
Daims
H
.
Fluorescence in situ hybridisation for the identification and characterisation of prokaryotes
.
Curr Opin Microbiol
.
2003b
6
3
302
9
.
193.
Wagner
M
,
Nielsen
PH
,
Loy
A
,
Nielsen
JL
,
Daims
H
.
Linking microbial community structure with function: fluorescence in situ hybridization-microautoradiography and isotope arrays
.
Curr Opin Biotechnol
.
2006
;
17
(
1
):
83
91
.
194.
Wan
X
,
Zhou
XR
,
Moncalian
G
,
Su
L
,
Chen
WC
,
Zhu
HZ
.
Reprogramming microorganisms for the biosynthesis of astaxanthin via metabolic engineering
.
Prog Lipid Res
.
2021
;
81
:
101083
.
195.
Wang
L
,
Liu
W
,
Tang
JW
,
Wang
JJ
,
Liu
QH
,
Wen
PB
.
Applications of Raman spectroscopy in bacterial infections: principles, advantages, and shortcomings
.
Front Microbiol
.
2021
;
12
:
683580
.
196.
Wang
W
,
Zhu
Y
,
Chen
X
.
Selective imaging of gram-negative and gram-positive microbiotas in the mouse gut
.
Biochemistry
.
2017a
56
30
3889
93
.
197.
Wang
X
,
Ren
L
,
Su
Y
,
Ji
Y
,
Liu
Y
,
Li
C
.
Raman-Activated Droplet Sorting (RADS) for label-free high-throughput screening of microalgal single-cells
.
Anal Chem
.
2017b
89
22
12569
77
.
198.
Wang
Y
,
Huang
WE
,
Cui
L
,
Wagner
M
.
Single cell stable isotope probing in microbiology using Raman microspectroscopy
.
Curr Opin Biotechnol
.
2016
;
41
:
34
42
.
199.
Wang
Y
,
Ji
Y
,
Wharfe
ES
,
Meadows
RS
,
March
P
,
Goodacre
R
.
Raman activated cell ejection for isolation of single cells
.
Anal Chem
.
2013
;
85
(
22
):
10697
701
.
200.
Wang
Y
,
Xu
J
,
Kong
L
,
Li
B
,
Li
H
,
Huang
WE
.
Raman-activated sorting of antibiotic-resistant bacteria in human gut microbiota
.
Environ Microbiol
.
2020
;
22
(
7
):
2613
24
.
201.
Wasmund
K
,
Pelikan
C
,
Schintlmeister
A
,
Wagner
M
,
Watzka
M
,
Richter
A
.
Genomic insights into diverse bacterial taxa that degrade extracellular DNA in marine sediments
.
Nat Microbiol
.
2021
;
6
(
7
):
885
98
.
202.
Watrous
JD
,
Dorrestein
PC
.
Imaging mass spectrometry in microbiology
.
Nat Rev Microbiol
.
2011
;
9
:
683
94
.
203.
Weiss
R
,
Palatinszky
M
,
Wagner
M
,
Niessner
R
,
Elsner
M
,
Seidel
M
.
Surface-enhanced Raman spectroscopy of microorganisms: limitations and applicability on the single-cell level
.
Analyst
.
2019
;
144
(
3
):
943
53
.
204.
Wiegand
S
,
Dam
HT
,
Riba
J
,
Vollmers
J
,
Kaster
AK
.
Printing microbial dark matter: using single cell dispensing and genomics to investigate the patescibacteria/candidate phyla radiation
.
Front Microbiol
.
2021
;
12
:
635506
.
205.
Wilbert
SA
,
Mark Welch
JL
,
Borisy
GG
.
Spatial ecology of the human tongue dorsum microbiome
.
Cell Rep
.
2020
;
30
(
12
):
4003
15.e3
.
206.
Wilhelm
RC
,
Singh
R
,
Eltis
LD
,
Mohn
WW
.
Bacterial contributions to delignification and lignocellulose degradation in forest soils with metagenomic and quantitative stable isotope probing
.
ISME J
.
2019
;
13
(
2
):
413
29
.
207.
Witkowska
E
,
Niciński
K
,
Korsak
D
,
Szymborski
T
,
Kamińska
A
.
Sources of variability in SERS spectra of bacteria: comprehensive analysis of interactions between selected bacteria and plasmonic nanostructures
.
Anal Bioanal Chem
.
2019
;
411
(
10
):
2001
17
.
208.
Woebken
D
,
Burow
LC
,
Behnam
F
,
Mayali
X
,
Schintlmeister
A
,
Fleming
ED
.
Revisiting N2 fixation in Guerrero Negro intertidal microbial mats with a functional single-cell approach
.
ISME J
.
2015
;
9
(
2
):
485
96
.
209.
Woebken
D
,
Burow
LC
,
Prufert-Bebout
L
,
Bebout
BM
,
Hoehler
TM
,
Pett-Ridge
J
.
Identification of a novel cyanobacterial group as active diazotrophs in a coastal microbial mat using NanoSIMS analysis
.
ISME J
.
2012
;
6
(
7
):
1427
39
.
210.
Xu
J
,
Zhang
L
,
Gong
H
,
Homola
J
,
Yu
Q
,
Xu
JJ
.
Tailoring plasmonic nanostructures for optimal SERS sensing of small molecules and large microorganisms
.
Small
.
2011
;
7
(
3
):
371
6
.
211.
Yamaguchi
T
,
Kawakami
S
,
Hatamoto
M
,
Imachi
H
,
Takahashi
M
,
Araki
N
.
In situ DNA-hybridization chain reaction (HCR): a facilitated in situ HCR system for the detection of environmental microorganisms
.
Environ Microbiol
.
2015
;
17
(
7
):
2532
41
.
212.
Yan
S
,
Qiu
J
,
Guo
L
,
Li
D
,
Xu
D
,
Liu
Q
.
Development overview of Raman-activated cell sorting devoted to bacterial detection at single-cell level
.
Appl Microbiol Biotechnol
.
2021
;
105
(
4
):
1315
31
.
213.
Yilmaz
S
,
Haroon
MF
,
Rabkin
BA
,
Tyson
GW
,
Hugenholtz
P
.
Fixation-free fluorescence in situ hybridization for targeted enrichment of microbial populations
.
ISME J
.
2010
;
4
(
10
):
1352
6
.
214.
Zhang
H
,
Zhong
H
,
Wang
X
,
Zhang
S
,
Shao
X
,
Hu
H
.
Use of NAD tagSeq II to identify growth phase-dependent alterations in E. coli RNA NAD+ capping
.
Proc Natl Acad Sci U S A
.
2021
;
118
(
14
):
e2026183118
.
215.
Zhang
P
,
Ren
L
,
Zhang
X
,
Shan
Y
,
Wang
Y
,
Ji
Y
.
Raman-activated cell sorting based on dielectrophoretic single-cell trap and release
.
Anal Chem
.
2015
;
87
(
4
):
2282
9
.
216.
Zhang
Z
,
Zhang
L
.
METAMVGL: a multi-view graph-based metagenomic contig binning algorithm by integrating assembly and paired-end graphs
.
BMC Bioinformatics
.
2021
22
Suppl 10
378
14
.
217.
Zhou
W
,
Le
J
,
Chen
Y
,
Cai
Y
,
Hong
Z
,
Chai
Y
.
Recent advances in microfluidic devices for bacteria and fungus research
.
Trac Trends Anal Chem
.
2019
;
112
:
175
95
.
218.
Zhou
X
,
Wu
H
,
Wen
H
,
Zheng
B
.
Advances in single-cell printing
.
Micromachines
.
2022
;
13
(
1
):
80
.
219.
Zoheir
AE
,
Meisch
L
,
Martín
MV
,
Bickmann
C
,
Kiselev
A
,
Lenk
F
.
Macroporous silicone chips for decoding microbial dark matter in environmental microbiomes
.
ACS Appl Mater Inter
.
2022
;
14
(
44
):
49592
603
.
220.
Zwirglmaier
K
,
Ludwig
W
,
Schleifer
KH
.
Recognition of individual genes in a single bacterial cell by fluorescence in situ hybridization: ring-fish
.
Mol Microbiol
.
2004
;
51
(
1
):
89
96
.