Abstract
Technologies to assay single cells and their extracellular microenvironments are valuable in elucidating biological function, but there are challenges. Sample volumes are low, the physicochemical parameters of the analytes vary widely, and the cellular environment is chemically complex. In addition, the inherent difficulty of isolating individual cells and handling small volume samples complicates many experimental protocols. Here we highlight a number of mass spectrometry (MS)-based measurement approaches for characterizing the chemical content of small volume analytes, with a focus on methods used to detect intracellular and extracellular metabolites and peptides from samples as small as individual cells. MS has become one of the most effective means for analyzing small biological samples due to its high sensitivity, low analyte consumption, compatibility with a wide array of sampling approaches, and ability to detect a large number of analytes with different properties without preselection. Having access to a flexible portfolio of MS-based methods allows quantitative, qualitative, untargeted, targeted, multiplexed, spatially resolved investigations of single cells and their similarly scaled extracellular environments. Combining MS with on-line and off-line sample conditioning tools, such as microfluidic and capillary electrophoresis systems, significantly increases the analytical coverage of the sample’s metabolome and peptidome, and improves individual analyte characterization / identification. Small volume assays help to reveal the causes and manifestations of biological and pathological variability, as well as the functional heterogeneity of individual cells within their microenvironments and within cellular populations.
Keywords: Bioanalytical microanalysis, single cell, microenvironments, mass spectrometry
1. Introduction
Advancing our capability to analyze single cells and their microenvironments represents an exciting goal of bioanalytical measurement science. The cell is a central structural and functional unit of most unicellular and multicellular organisms. Methods to assay small volume samples from heterogeneous biological organisms, including the ability to select the most appropriate analytes for a specific investigation while ensuring reduced analyte consumption, can provide important information about cellular function. Depending on the analytical approach used and the specimen being examined, sample sizes may vary widely, from femtoliter to milliliter or larger. Here, we consider samples of less than a microliter as being small volume; these samples can be comprised of fundamental structures such as individual organelles and cells, as well as their extracellular environments.
Why are single cell assays important? One key factor is that the coordinated activities among individual cells contribute to the physiology and behavior of multicellular organisms, and also to the organization of ecological communities in unicellular organisms. Cellular variability, as well as pathological and functional heterogeneities, are integral to the development of individual biological and behavioral traits, and have important roles in the etiology of many diseases. The study of single cells and their microenvironments (i.e., extracellular spaces) aids our understanding of these phenomena, as well as the overall structure, organization and activity of organisms and cellular populations [1–6].
The initial achievements in small sample analysis were associated with the technological progress of microscopy, starting centuries ago with Robert Hooke and Anton van Leeuwenhoek [7]. Visual observations of individual cells led to the development of methods that were used to directly or indirectly determine cellular chemical composition, e.g., the Gram and Golgi cell staining methods. Today, chemical investigations of a diverse range of analyte types can be qualitatively and quantitatively characterized and localized using a plethora of new technologies that employ molecular probes, analyte separation approaches such as chromatography and capillary electrophoresis (CE), microfluidic devices, electrochemical methods, mass spectrometry (MS), and a number of optical and resonance methods [1,8].
When selecting the appropriate tool for a small volume analysis, MS is often the method of choice; it provides low detection limits, multiplexed detection, greater coverage for many analyte classes, and capabilities for nontargeted, targeted, qualitative, and quantitative investigations. Hyphenation of MS to sample conditioning and/or microseparation approaches, such as solid phase microextraction, CE, microfluidics, gas chromatography, and capillary-scale chromatography, can improve the coverage of the peptidome and metabolome and add useful information about the analytes, such as migration time and hydrophobicity.
In this review, we focus on small volume assays of two analyte classes—peptides and metabolites. These compounds influence cellular morphology, energy balance, cell-to-cell communication, and many other cellular functions. There is a caveat to consider when planning a single cell investigation; given the array of molecules and their large dynamic range, the current limits of detection of most available methods do not allow the characterization of an entire peptidome or metabolome of an individual cell. Nonetheless, one can measure the most abundant and/or detectable analytes present, many of which have important functional roles, e.g., signaling molecules. These include peptides as well as a large variety of small molecules such as lipids, carbohydrates, and amino acids, which often accumulate inside a cell and are released into the surrounding microenvironment under tightly regulated events.
We begin discussion with an overview of the single cell sample preparation process, followed by a section highlighting several commonly used mass spectrometric methods to analyze small volume samples. Next we review the hyphenation of analyte separation techniques to MS, and close with MS-based technologies used to measure cellular microenvironments and releasates.
2. Sample preparation
One issue with single cell measurements is that many typical sample conditioning steps, such as concentration, desalting, and fractionation, can lead to unacceptable sample / analyte loss; thus, choosing the appropriate sampling technique becomes critical to a successful MS investigation. Sampling an individual cell is challenging due to its small scale, chemical complexity, and the structural and chemical instability of the cell and surrounding spaces. Often, as in the case of single neurons and astrocytes, it is difficult to isolate cells such as a neuron or astrocyte in their entirety due to the extensive branching of cellular processes and their intertwining and mechanical interconnectivity with other cells. Additionally, removing a cell from its native environment triggers intracellular biochemical changes. Therefore, special care is required to stabilize the sample by lowering its temperature, adding stabilization agents such as glycerol and trehalose, and decreasing sample collection time. Stringent control experiments are often performed to determine the extent of possible alteration due to sample preparation.
An assortment of MS-compatible sample collection and manipulation approaches are available, including whole cell mechanical dissection, direct cytoplasmic sampling, flow cytometry, optical trapping, and microenvironment collection using a variety of microfluidic devices [9]. Furthermore, some of the approaches, such as optical trapping, can be used to isolate and profile analytes at the subcellular level [10,11]. Among these, microfluidics provides a powerful platform on which all steps of the sample preparation process can be performed, including cell lysis and content conditioning, prior to MS analysis [12]; these devices are also used to collect cell releasate (see Sections 4 and 5 for an expanded discussion of how microfluidics is applied in MS-based investigations).
3. MS-based approaches
A variety of MS-based methodologies have been developed and applied to direct assays of small volume samples: matrix-assisted laser desorption / ionization (MALDI) [13–16], desorption electrospray ionization [17], laser desorption ionization [18], secondary ion MS [19– 22], nanostructure-initiator [23], electrospray ionization (ESI) [24,25], inductively coupled plasma (ICP) [26–29], and laser ablation electrospray ionization [30–32]. Here we highlight methods that are particularly well suited to such small volumes assays: direct single cell profiling using MALDI MS, followed by two high-throughput methods, microarray mass spectrometry (MAMS) and mass cytometry, and MALDI MS imaging (MSI). For more information on these and other technologies for single cell and other small volume investigations, the interested reader is referred to several recent reviews [6,12,33,34].
3.1. Direct MALDI MS
Single cell analysis generates data that can document an individual cell’s actual chemical state, rather than a deriving an “average” state based on assay of a cell population. So why aren’t single cell measurements performed more routinely? A major challenge is the small sample size and the issues this creates for its chemical characterization. One approach that has been successfully used to assay small volume samples is MALDI MS [14,35,36]. Because of its low detection limits, high salt tolerance and wide molecular mass range of detection, it is one of the most commonly used methods for single cell peptide and metabolite measurements. A MALDI MS investigation generally starts with the identification and removal of the target cell from a population. The isolated cell is exposed to a MALDI matrix solution and the combined sample dried onto a sample plate. Next, the sample is irradiated with a laser to generate a plume of neutral particles and ions, which are subsequently analyzed by a mass analyzer. As each laser shot only consumes a small portion of the sample, one can average the signal from hundreds to thousands of laser shots. Moreover, multiple experiments can be conducted on just one cell. MALDI MS has been used as a complementary technique to confirm known peptide localization, and as a tool for new molecular discovery. Many classes of analytes from a variety of organisms have been assayed with MALDI MS, including DNA, RNA, proteins, peptides, and metabolites.
Several early MALDI MS-based investigations profiled isolated neurons from the central nervous system of the pond snail Lymnaea stagnalis [37–39] and isolated dense core vesicles from the sea hare Aplysia californica [40]. Molluskan neurons are well suited for single cell analysis and method development because they are relatively large and are involved in welldefined physiological functions. These cells are variable in size, with cell body diameters ranging from 10–1000 µm. The larger cells are appropriate targets for individual isolation and manipulation; not only do they contain easier to measure volumes, but they are similar to smaller cells in terms of their mechanisms of cell functioning. As the difference in analyte amounts from the smaller to larger cells is a million fold, this increased volume enables many measurements. Beyond mollusks, there are many examples of peptide-content studies using smaller cells. For example, individual rat pituitary cells were isolated and profiled for the presence of cell-to-cell signaling peptides [8,41]. Peptides from the pro-opiomelanocortin (POMC) and cocaine-and-amphetamine-regulated transcript prohormones were detected, consistent with the known peptide content of the pituitary.
While typically used to identify compounds, MALDI MS can also generate quantitative information from an individual cell. Several quantitative single cell MALDI MS approaches were described using A. californica neurons [42]. In these approaches stable isotope analyte labeling allowed measurement of relative peptide content, and standard addition was conducted for absolute quantitation. Quantitative analysis of small volume samples with standard addition is possible because MALDI MS consumes only a fraction of the sample per measurement, allowing sequential data collection after each addition of standard and sample reconstitution.
Another interesting application of MALDI MS involves using tandem MS (MS/MS) for the structural characterization of peptides, which can be challenging because typically MS/MS requires larger analyte amounts than direct MS profiling. In one notable example, Bai and colleagues [43] identified several D-amino acid-containing peptides in abdominal ganglia neurons from A. californica. These peptides exhibit the unusual post-translational modification in which one amino acid residue is isomerized from the L-form to the D-form. While isomerization does not introduce a mass shift, the reported results showed that the MS/MS spectra contained differences in the intensity of the fragment ions sufficient for the identification and quantification of the studied D-amino acid-containing peptides.
Several MALDI MS-based studies have incorporated other analytical approaches to confirm analyte identification and localization. Jarecki et al. [44] demonstrated the efficacy of single cell MALDI MS for the detection of low abundance peptides in rare cells. They compared peptides detected in ALA and RID neurons from the dorsal ganglion of the nematode Ascaris suum with known peptides previously found using immunocytochemical and in situ hybridization approaches; the results from all three methods were in good agreement. Six novel peptides were also observed and sequenced via MS. The complementary nature of different analytical approaches is further illustrated in two studies on insect neurons [15,45]. The neuropeptidome of the antennal lobe of the American cockroach Periplaneta americana was investigated with MALDI MS by systematically reducing the sample size from whole tissue to single cells [15]. Dissected tissues were first profiled via MALDI MS to get an overview of their neuropeptide content. Single glomeruli, cell clusters, and neurons were then progressively interrogated in combination with immunolabeling to obtain data on the peptide profiles of different regions of the specimen. However, immunolabeling typically requires a fixation step that negatively affects the MS detection of many compounds. To resolve this issue, Neupert et al. [15] created a workflow to facilitate MALDI MS analysis after immunostaining. By adding a mixture containing 2,4-dinitrophenylhydrazine and α-cyano-4-hydroxycinnamic acid and heating the sample, the crosslinks formed by fixation can be destroyed, freeing analytes for MS detection. Together these studies show that immunocytochemical staining is useful for localizing relevant cells for MALDI MS analysis [15,44,45]. Advances like this broaden the applicability of MALDI MS single cell profiling to more biological systems.
3.2. MALDI MSI
MALDI MSI is an information-rich technique that adds spatial information to MS data sets. It is commonly performed in the microprobe mode in which the laser beam is rastered across the sample to collect position-resolved mass spectrometric information [46–48]. Although the spatial resolution of MALDI MSI is often limited by laser dimensions (typically 15–75 µm), and by having a sufficient amount of analyte to detect within the laser spot, the investigation of cells at subcellular scales is becoming possible. Currently, MALDI MSI spatial resolution pales in comparison with the resolution attained by optical microscopy; however, the molecular information acquired makes single cell MALDI MSI an important approach to have in one’s analytical armamentarium.
There are many examples showing the capability of MALDI MSI to reveal the subcellular localization of analytes and cell-to-cell chemical heterogeneity within a tissue section. One early study used single cell MALDI MSI to investigate a culture of A. californica bag cell neurons [49]. Even at 50 µm lateral resolution, the cell body and neuronal processes could be resolved based on peptide signal. More recently, MSI experiments have been performed at <10 µm spatial resolution using a laboratory-built instrument [50–53]. Mouse pituitary sections were imaged at 5 µm resolution [51] and a distribution of the POMC peptide, α-melanocyte-stimulating hormone, between two mixed cell types within the intermediate pituitary was observed, illustrating how single cell multiplexed peptide information can be obtained without cell isolation from tissue. In another study, HeLa cells were imaged at 7 µm spatial resolution, demonstrating that an instrument with high mass resolution and high mass accuracy can improve molecular image contrast and specificity by reducing interference from compounds with similar m/z ratios [52]. High spatial resolution using a custom built mass spectrometer was also achieved by Zavalin et al. [54]. They imaged the distribution of lipids in human embryonic kidney 293 (HEK-293) and RKO cells at nearly 1 µm spatial resolution. They also showed the potential for subcellular peptide and protein MSI in a profiling experiment that showed insulin localization to the cytoplasm of human pancreas islet cells.
It is possible to obtain a higher spatial resolution than the diameter of the laser beam footprint. In one example, a network of A. californica neurons was cultured on a stretchable substrate [55]. Prior to MSI, the network was broken apart into individual pieces containing single cells by stretching the substrate. Spatially targeted acquisition of mass spectra and implementation of a molecular image reconstruction step allowed for the determination of the original distribution of analytes in the network of single cells. Both MS and MS/MS experiments were performed using the approach, thus aiding neuropeptide identification.
Oversampling is another method that improves spatial resolution [56]. By depleting the sample with the laser at the starting point and stepping the laser by less than the beam diameter, analyte signal will only come from the sample region that had not previously been exposed to laser irradiation, thus improving spatial resolution. Jurchen et al. [56] developed this method using individual neurons deposited on an electron microscopy calibration grid and traditional UV laser microprobe MALDI MSI. Features of the grid cannot be resolved when the raster step is 100 µm and laser beam footprint is ~100×200 µm. However, both the grid’s wires (40 µm wide) and the holes (60 µm wide) become visible when the laser is rastered at 25 µm and oversampling is used. A ~100 µm neuron from A. californica was imaged using this approach. Changes in peptide signal intensity between adjacent pixels was observed, demonstrating that distribution of analytes at a spatial resolution smaller than the laser beam diameter can be imaged.
While impressive advances have been made, single cell laser microprobe MSI tends to have a low throughput because hundreds or even thousands of locations have to be examined one at a time. Stigmatic imaging presents an alternative that can bypass some of these problems. The sample is irradiated with a diffused laser and specialized ion optics preserve the generated ions’ relative two-dimensional spatial localization before their detection by a positionsensitive detector. The Heeren group has developed—and is continually improving—mass microscopes that allow this detection modality [57–61]. Several types of samples have been imaged using the mass microscopes [58,62–64]. In one study, pituitary gland sections from rats, mice, and humans were imaged with 4 µm resolving power [62]; known pituitary peptides were detected and localized for all three sample types. Stigmatic imaging also can be done with infrared lasers, permitting the use of water that is endogenous to the sample as a substitute for classical MALDI matrixes [64]; the boundary between the liver and the stomach of a bait fish was also investigated with 10 µm lateral resolution using this approach. The tradeoff with mass microscopy is that ions in a relatively narrow m/z range are detected in each image so that the chemical information is reduced although the spatial resolution and imaging speed are increased.
3.3. MAMS
The typical single cell MS experiment reveals important information on the chemical composition of a few cells within a large population. However, how representative these cells are for the population is not always known. Even in a supposedly homogeneous population, individual cells differ from each other based on a variety of factors, e.g., different stages of the cell cycle. Many biological phenomena, such as drug resistance or malignant cell transformations, may begin with rare events that can be missed when only a small number of cells in a population are interrogated. A thorough understanding of a biological system’s functioning requires high-throughput single cell analysis, which can generate population level data from individual cells. In fact, the cellular-scale MSI methods discussed in the previous section can generate this type of output because hundreds to thousands of cells can be analyzed in one experiment. However, intact tissues or their sections are difficult targets, in part, because it is challenging to align each cell with the laser microprobe automatically. As a result many of the acquired mass spectra will contain signals from multiple neighboring cells and extracellular regions. Analyte migration between cells during MALDI matrix application can also be a confounding factor. Therefore, methods that attain spatial separation between many individual cells and offer highly parallel sample preparation prior to MALDI MS analysis have been developed.
One such approach is MAMS, which is conducted on an array of hydrophilic wells surrounded by an omniphobic material [65]. With the proper cell suspension density, it is possible to deposit single cells into wells by pipetting the suspension onto the array, allowing them to sediment, and removing excess liquid. As result, many wells can be quickly and repeatedly loaded for high throughput and quantitative experiments [66]. The number of cells deposited in a well follows a Poisson distribution and can be visualized using a microscope [67]. Once optimized, each well should contain only a few cells, with a subset containing none or individual cells.
Using the MAMS approach, the metabolite profiles of a population of the yeast Saccharomyces cerevisiae were investigated [67]. Visual inspection showed that the MAMS wells contained 1–15 cells, with ~50% containing three cells or fewer. Two subpopulations with different levels of fructose-1,6-biphosphate were found, reflecting differences in glycolytic fluxes in the cells of the subpopulations.
MAMS investigation of cellular populations can be enhanced by additional bioanalytical examination of the same cells using Raman and fluorescence measurements. By employing these multi-method approaches, metabolomic characteristics of the transition of the algal cells Haematococcus pluvialis from the motile to the dormant state were revealed [68]. The dormant stage of H. pluvialis was characterized by an accumulation of the red pigment astaxanthin and a decrease of the adenosine triphosphate to adenosine diphosphate ratio, indicating a lowering of the energetic state of dormant cells. A metabolically unique cell was also detected, illustrating the capability to find rare cells with high throughput methods.
3.4. Mass cytometry
Another high throughput method is mass cytometry, which combines flow cytometry with a specially designed ICP-MS system [26,27,69]. This approach is distinct from those discussed so far in two ways; first, it presorts the cells using flow cytometry, and second, the mass spectrometer detects immuno-labeled analytes that contain differentiable labels. The approach is similar to fluorescent flow cytometry and requires one to preselect the analytes of interest before the experiment. While flow cytometry with fluorescent detection is capable of analyzing thousands of cells per minute, the spectral overlap of common fluorescence probes restricts the number of separate antibodies that can be used. Mass cytometry addresses this problem by using antibodies labeled with distinct stable isotopes. Thirty or more antibodies labeled with different isotopes can be used simultaneously to detect this number of analytes, including peptides, proteins, and some metabolites in a single cell, providing more analytical channels than fluorescence-based flow cytometry. This technique allows for the design and implementation of extensive barcoding schemes to enable the assay of a range of analytes, practically limited by the availability of specific antibodies and analyte concentration. Thousands of labeled cells from cell populations can be analyzed in one experiment. Importantly, the approach can determine absolute quantities of the stable isotope labels to provide high throughput quantitative data.
Mass cytometry has been applied to a variety of biological and analytical sciences [26,70–76]. One of the most extensive applications has been to study immune cells. For example, lymphocytes, the cells of the immune system, possess a wide range of antigen specificities [76]. Recently, Bendall et al. [26] analyzed the immune cell population in bone marrow during hematopoiesis. A total of 24 distinct cellular populations were detected using 13 surface expression markers. Additionally, differences in responses to stimuli between these cell types were investigated using 18 intracellular markers. This study was later expanded to include measurements of cell cycle-related parameters [70].
Not surprisingly, specific major immune cell types can be also distinguished and characterized using mass cytometry. Over 200 functional phenotypes of CD8+ T cells were identified by measuring 28 parameters representing surface markers and cytokines, among others [75]. Horowitz et al. [71] analyzed natural killer cells using 37 parameters. These results indicated the presence of between 6,000 to 30,000 cellular phenotypes within an individual. These studies also show that with high throughput, multiplexed analytical approaches, single cell analysis can generate deep insight into the organization of the cellular subpopulations of complex biological systems.
4. Analyte separation techniques hyphenated to MS
The chemical complexity of single cells makes comprehensive characterization of their peptide and metabolite content in one MS measurement challenging. The competitive nature of ionization, limited dynamic range of MS instruments, and relatively low mass resolution of the most sensitive mass spectrometers, restrict the number of analytes that can be directly analyzed in a sample. To improve analyte coverage and identification, a variety of separation and preconcentration methods can be hyphenated to MS detection. CE, capillary liquid chromatography, and microfluidics are frequently incorporated into the sample separation and conditioning process. Providing the capability to analyze small volumes while minimizing analyte loss are key features of these approaches. Some methods that include injection of the entire cell lysate or the extracellular environment into a capillary or a microfluidic device increase the ability to collect a greater fraction of the analyte content. The comparable internal volumes of the analytical devices and the analyzed cells or cellular environments result in minimal sample dilution [9,77–79].
Capillary zone electrophoresis is a useful approach for analyte separation and introduction of the sample into the mass spectrometer via nanoESI ion source [80]. Many endogenous peptides and metabolites exhibit different electrophoretic mobilities in MS-compatible CE buffers such as 1% formic acid. CE not only separates analytes but also provides information on their migration times, which in combination with fragmentation patterns obtained using MS/MS, allows endogenous analyte identification with high levels of confidence. Analyte fragmentation in MS/MS is achieved using a variety of techniques such as collision-induced dissociation, electron capture dissociation, electron transfer dissociation, and postsource decay. These techniques are often used to obtain key structural information on detected metabolites and peptides, including post-translational modifications [5], and can be applied in small volume analysis. CE-MS and CE-MS/MS demonstrated good performance in the investigation of a single cell metabolome [81], with more than 300 molecular features detected in a single neuron and 30 metabolite identities confirmed. The approach allows different types of neurons and their functional conditions/states to be distinguished [82].
There are numerous analytes present in single cells at levels below the limits of detection of many MS systems. On-column and on- or off-line analyte preconcentration methods can be employed to make the analysis feasible. On-line methods are preferable as they avoid analyte loss due to liquid handling or time-dependent degradation. Sample stacking is a commonly used on-column analyte preconcentration method that relies on changes in the electrophoretic mobility of the analytes at the boundary of two buffers prepared at varying concentrations or pH levels. The sensitivity of detection can be improved by 5- to 10-fold using the approach, but its extent is limited by the ratio of the conductivities of the buffers used. Sample stacking was applied in a CE-MS investigation of endogenous nucleotide content of individual A. californica neurons [83].
CE-MS technology has reached a level of maturity that allows for investigation of subcellular samples. Aerts et al. [84] examined the metabolite content of different single functioning rat neurons using a patch clamp pipet to collect ~3 pL of cytoplasm from electrophysiologically and morphologically identified thalamocortical neurons, a region known to contain both gamma-aminobutyric acidergic as well as glutamatergic cells. Using CE-MS, approximately 60 metabolites were detected in the cytoplasm of the cells studied. Demonstrating the ability to characterize cells, glutamate was detected in physiologically identified glutamatergic cells, and gamma-aminobutyric acid (GABA) in the cytoplasm of physiologically-characterized GABAergic neurons, while GABA was not detected within the glutamatergic cells. This study demonstrates the usefulness of combining electrophysiological, morphological, and single cell CE-MS approaches for determination of neurochemical correlates of functional activities in neuronal networks.
Microfluidic devices have physical dimensions that are well suited for small volume sample collection, processing, and injection into mass spectrometers. The application of microfluidics in biological research is becoming widespread. Microfluidic platforms allow manipulation of picoliter to nanoliter sample volumes and offer a wide variety of designs and features permitting individual cell capture, culturing, and stimulation, as well as direct sampling of cellular content and/or extracellular microenvironments, analyte separation, and on-line or offline hyphenation to different detection modalities such as MS [33,85]. Preconcentration of analytes can be done directly within devices, e.g., using incorporated solid phase extraction (SPE) materials [86]. As discussed further in the next section, such devices are especially useful for collection, preconcentration, and assaying of analytes released from cells.
5. Measuring cellular microenvironments, including cellular release, via MS
Thus far we have discussed MS approaches and their applications that provide information on the chemical contents of cells [12]. The results of these experiments are useful for the phenotypical classification of cells, but, typically, do not directly reveal functional roles of the detected analytes. In contrast, the investigation of stimulated cellular release provides an opportunity to determine a subset of cellular metabolites and peptides that are likely to be involved in cell-to-cell signaling. Secretomics, the study of the compounds released from a cell, uses a wide range of methodologies, including MS, microfluidics, Western blotting, enzyme-linked immunosorbent assays, gel electrophoresis, radioimmunoassay / labeling with radioactive isotopes, and amperometric detection [79,87–94]. Released compounds vary from small diatomic gases to large proteoglycans and peptides; thus, it is not surprising that a range of characterization approaches are required to assay this large variety of analyte types.
MS is an important tool for studying cellular release as it provides multiplexed information about the protein, peptide, and metabolite content of the releasate without the need for analyte pre-selection. The focus of cellular release studies range from biomarker detection and identification to the determination of protein concentrations [87]. We have chosen to highlight work that has involved the detection and identification of cell-to-cell signaling molecules in the nervous system. Uncovering the peptide composition of cellular releasate helps to elucidate the functional roles of observed peptides by relating their release parameters to specific stimulation paradigms. For peptides with known activities and targets, it may help to define the physiological role of the studied cell. MALDI-time-of-flight (TOF) MS is often used to study peptide secretion due to its salt tolerance, low detection limits, and ability to be hyphenated with a variety of sample conditioning approaches [14]. As with all single cell studies, secretomics is complicated by the chemical complexity of extracellular environments, low analyte amounts, and the need to remove unwanted chemical components from the sample prior to MS analysis. Microfluidic technologies enable efficient single cell culture, direct cell stimulation, and more specific temporal and spatial capture of released analytes.
Many microfluidic systems have been designed for the capture, treatment, and analysis of single cells, some of which have been applied to investigate analyte release from individual cells or small cellular structures [33,77–79]. Wei et al. [95] developed a series of microfluidic chips that incorporate cell culture and on-chip pretreatment coupled with ESI-quadrupole-TOF MS for analysis. In order to detect glutamate secreted from stimulated PC12 cells, the cells were cultured and stimulated in microchannels on one chip and then the extracellular fluid was moved through tubing to a miniature extraction chip in which the sample was flowed over polymer SPE beads for pretreatment. After pretreatment, the sample was once again moved through tubing to the ion source. This series of chips could easily be used for the analysis of peptides released upon stimulation. Microfluidic devices designed for the detection of peptides are used in biomarker detection assays. Yang et al. [96] presented a device in which specific peptides are affinity captured in microchannels lined with antibodies. This particular system allows detection of as few as 300 peptide molecules and can be when applied in targeted analysis of released peptides.
Collection of the extracellular media and its direct analysis via MALDI-TOF MS is a powerful approach for the analysis of normal and pathological peptide secretion [97]. However, even though MALDI-TOF MS has a high inorganic salt tolerance, the high levels of salts present in the physiological extracellular media hinders effective molecular characterization with MS. Therefore, SPE is often used for sample conditioning as it improves detection, minimizes sample loss, and decreases the salt content before the MS measurement. An analytical system developed by Croushore et al. [98] allowed for the selective stimulation of cultured A. californica bag cell neurons through a device in which fluid flow was controlled via applied external pressures and pneumatically controlled microvalves. Once the cells were stimulated, microliter volumes of the cellular microenvironment were collected, purified using SPE, and analyzed by MALDI-TOF MS analysis. This approach allowed low temporal resolution studies of analyte release from low-density cultures and single cells.
Mao et al. [86] studied intercellular communication through the integration of stimulation, release, and pretreatment steps all on one chip. A microfluidic device capable of culturing two distinct cell populations was used for stimulation of one population and a surface tension plug control was used to control the chemical signaling between the populations. Once the desired stimulation occurred, samples of the extracellular media were moved through the chip and into micro-SPE columns for analyte capture and sample conditioning. After pretreatment, the analysis was performed using ESI-Q-TOF-MS (Fig. 3A).
Fig. 3.
A variety of SPE-based analyte collection approaches have been developed and applied to small volume sample analysis. (A) An integrated microfluidic device allowing cell stimulation, cell-cell communication, and sample pretreatment. (Reprinted with permission from reference [86]. Copyright 2012 American Chemical Society). (B) Schematic of sample collection from A. californica bag cell neurons and its SPE pre-treatment prior to MALDI MS analysis. (Adapted with permission from reference [97]. Copyright 2001 WILEY-VCH Verlag.) (C) Placement of SPE beads (arrows) directly on neurites of cultured neurons. (Reprinted with permission from reference [99]. Copyright 2005 National Academy of Sciences of the United States of America.) (D) A two-capillary system allowing secretagogue application, released compound collection, and sample processing. (Reprinted with permission from reference [100]. Copyright 2011 American Chemical Society.) (E) A concentric dual capillary system used for temporal and spatial investigation of release. Reduced analyte dilution and flexible sample conditioning using octadecyl-modified silica nanoparticles are important properties. (Reproduced from reference [101] with permission from the Royal Society of Chemistry.)
Quantitative analyses of released analytes can also be performed using a microfluidic platform hyphenated to MS. Zhong et al. [102] developed a device for the quantification of peptides released from A. californica bag cell neurons. The device consisted of one serpentine channel treated with octadecyltrichlorosilane (OTS), used to collect the peptides at specific locations. A continuous flow of extracellular fluid brought the released peptides into contact with the OTS. This OTS layer with captured molecules was then interrogated using MALDI MSI. The distance that the peptide was detected along the length of the channel correlated with the amount of the peptide in the sample.
Commercially available, as well as home built SPE devices, are widely used to desalt and concentrate samples prior to analysis via MS. C18 reversed phase media-packed pipet tips have been used for the conditioning of releasates collected from clusters of bag cell neurons of A. californica [97], as well as from low-density bag cell neuron cultures [98]. Due to the large volume of C18 material used, this approach can be utilized when release from multiple cells is investigated. However, reducing the volume of SPE material, therefore, leading to higher analyte preconcentration, aids in the detection of the small amounts of analytes released from a single cell. In one example, a small number of SPE particles were mounted onto the surface of Parafilm M. The ~30 µL volume of collected extracellular media containing cell releasate was deposited onto the particles through a fused silica capillary (Fig. 3B) [97].
Further improvement in the spatial detection of peptide release was achieved by the use of SPE beads placed directly on the processes of cultured A. californica bag cell neurons (Fig. 3C) [99]. Alternatively, SPE packed pipette tips or cartridges, as demonstrated by Hatcher et al., can be used to collect peptide release [103]. Finally, to improve both spatial detection and sampling efficiency, lauryl methacrylate-ethylene glycol dimethacrylate porous polymer monolithic columns have been used to investigate both A. californica and mammalian systems, attaining sub-picomolar limits of detection with sample volumes of ~ 4 µL [89].
Additional improvements in analyte retention can be realized by acidification, alkylation, or addition of ion-pairing reagents to the sample. However, in most cases these modifications are not easily applied to the extracellular environment without interfering with the physiological activity of the studied cells. Therefore, to increase SPE analyte extraction efficiency in released peptide analysis, Fan et al. [100] customized particle-embedded monolithic capillaries with pyrrolidone or ethylenediamine in poly(stearyl methacrylate-co-ethylene glycol dimethacrylate) and used them to collect A. californica bag cell neuron releasates. The complete system consisted of two capillaries connected to individual syringe pumps, one for application of the secretagogue to the individual neurons at a rate of 0.25 µL/min and the other for collection and pretreatment of the releasate at the same flow rate for 30 minutes, allowing for low femtomole limits of analyte detection and specific analyte targeting (Figure 3D). This SPE approach was improved in a follow up study [101] with the use of a concentric dual capillary system consisting of an outer capillary that surrounds the stimulation and sampling area, and an inner octadecylmodified silica nanoparticle-filled capillary to collect and pretreat the ~10 µL sample. The outer capillary, which delivers the secretagogue, either surrounds the neuron or is positioned above the targeted area of the tissue and remains in that position while the inner sampling capillary can be removed and replaced if multiple collections of releasates are necessary (Fig. 3E). This set-up allows relatively high temporal and spatial resolution of cellular release investigation as well as reduced sample dilution.
6. Concluding remarks
MS-based analysis of small volume samples is a rapidly advancing analytical area. The ability to characterize individual cells and cellular release from selected cells is facilitating several areas of study, including cancer biology, stem cell research, neuroscience and even studies of interactions between specific cells in complex ecological communities. Within the United States alone, there has been a recent emphasis by government funding agencies such as the National Institutes of Health and the National Science Foundation to characterize the cellular makeup of the brain and other organs with single cell resolution.
Of course, more research is needed. After all, genomic and transcriptomic approaches allow the characterization of nearly the complete genome and transcriptome of a selected cell. This is not the case, however, for the metabolomes, proteomes, and peptidomes of single cells, where the fraction of analyte coverage for these limited volume samples is likely only a few percent (or less), especially when examining small cells such as mammalian glia. Small volume sample analysis is rapidly improving its figures of merit with the advent of new sampling protocols, sample conditioning and MS-based approaches, as well as bioinformatics tools to manage the large data sets obtained from these studies. The continued advancement of technologies to monitor individual cells within a complex microenvironment offers great promise for understanding cellular behavior and detecting changes caused by illness, physical damage or exposure to drugs, as well as designing improved therapies for a broad range of diseases.
Fig. 1.
Images of the distribution of an analyte at m/z 754.536 from a HeLa cell sample with (A and C) 49 µm and (B and D) 7 µm pixel sizes. The high mass accuracy in (C) and (D) improves image contrast by reducing interference from neighboring peaks (Reprinted with permission from ref [52]. Copyright 2012 American Chemical Society).
Fig. 2.
Microarray MS chips allow high throughput sample preparation in single cell analysis. A MAMS chip containing (A) 24 or (B) 12 spots per lane can be automatically aliquoted by dragging the sample over each lane (D–F). (Reprinted with permission from reference [66]. Copyright 2013 American Chemical Society).
Highlights.
Mass spectrometry (MS) is a versatile method for characterizing small volume samples.
Qualitative, quantitative, targeted, and discovery investigations are enabled by MS.
Samples as small as individual cells can be assayed for their small molecule content.
New approaches enable MS-based high throughput small volume sample analyses.
Acknowledgements
The authors acknowledge the support of the National Institutes of Health by Award No. R01 NS031609 from the National Institute of Neurological Disorders and Stroke and Award No. P30 DA018310 from the National Institute on Drug Abuse. The content is solely the responsibility of the authors and does not necessarily represent the official views of the funding agencies.
Abbreviations
- MS
mass spectrometry
- MALDI
matrix assisted laser desorption / ionization
- MSI
mass spectrometry imaging
- MAMS
microarray mass spectrometry
- MS/MS
tandem mass spectrometry
- ICP-MS
inductively coupled plasma mass spectrometry
- ESI
electrospray ionization
- TOF
time-of-flight
- CE
capillary electrophoresis
- SPE
solid phase extraction
- OTS
octadecyltrichlorosilane
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Conflict of interest
All authors state that they have no conflicts of interest.
References
- 1.Weaver WM, Tseng P, Kunze A, Masaeli M, Chung AJ, Dudani JS, Kittur H, Kulkarni RP, Di Carlo D. Advances in high-throughput single-cell microtechnologies. Curr. Opin. Biotechnol. 2014;25:114–123. doi: 10.1016/j.copbio.2013.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Liu N, Liu L, Pan X. Single-cell analysis of the transcriptome and its application in the characterization of stem cells and early embryos. Cell. Mol. Life Sci. 2014;71:2707–2715. doi: 10.1007/s00018-014-1601-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Irish JM, Doxie DB. High-dimensional single-cell cancer biology. Curr. Top. Microbiol. Immunol. 2014;377:1–21. doi: 10.1007/82_2014_367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Junker JP, van Oudenaarden A. Every cell is special: genome-wide studies add a new dimension to single-cell biology. Cell. 2014;157:8–11. doi: 10.1016/j.cell.2014.02.010. [DOI] [PubMed] [Google Scholar]
- 5.Li L, Sweedler JV. Peptides in the brain: mass spectrometry-based measurement approaches and challenges. Annu. Rev. Anal. Chem. 2008;1:451–483. doi: 10.1146/annurev.anchem.1.031207.113053. [DOI] [PubMed] [Google Scholar]
- 6.Zenobi R. Single-cell metabolomics: analytical and biological perspectives. Science. 2013;342 doi: 10.1126/science.1243259. [DOI] [PubMed] [Google Scholar]
- 7.Gest H. The discovery of microorganisms by Robert Hooke and Antoni Van Leeuwenhoek, fellows of the Royal Society. Notes Rec. R. Soc. Lond. 2004;58:187–201. doi: 10.1098/rsnr.2004.0055. [DOI] [PubMed] [Google Scholar]
- 8.Rubakhin SS, Churchill JD, Greenough WT, Sweedler JV. Profiling signaling peptides in single mammalian cells using mass spectrometry. Anal. Chem. 2006;78:7267–7272. doi: 10.1021/ac0607010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Cecala C, Sweedler JV. Sampling techniques for single-cell electrophoresis. Analyst. 2012;137:2922–2929. doi: 10.1039/c2an16211c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Lapainis T, Sweedler JV. Contributions of capillary electrophoresis to neuroscience. J. Chromatogr. A. 2008;1184:144–158. doi: 10.1016/j.chroma.2007.10.098. [DOI] [PubMed] [Google Scholar]
- 11.Cecala C, Rubakhin SS, Mitchell JW, Gillette MU, Sweedler JV. A hyphenated optical trap capillary electrophoresis laser induced native fluorescence system for single-cell chemical analysis. Analyst. 2012;137:2965–2972. doi: 10.1039/c2an35198f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Rubakhin SS, Romanova EV, Nemes P, Sweedler JV. Profiling metabolites and peptides in single cells. Nat. Methods. 2011;8:S20–S29. doi: 10.1038/nmeth.1549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Amantonico A, Urban PL, Fagerer SR, Balabin RM, Zenobi R. Single-cell MALDI-MS as an analytical tool for studying intrapopulation metabolic heterogeneity of unicellular organisms. Anal. Chem. 2010;82:7394–7400. doi: 10.1021/ac1015326. [DOI] [PubMed] [Google Scholar]
- 14.Li L, Garden RW, Sweedler JV. Single-cell MALDI: a new tool for direct peptide profiling. Trends Biotechnol. 2000;18:151–160. doi: 10.1016/s0167-7799(00)01427-x. [DOI] [PubMed] [Google Scholar]
- 15.Neupert S, Rubakhin SS, Sweedler JV. Targeted single-cell microchemical analysis: MS-based peptidomics of individual paraformaldehyde-fixed and immunolabeled neurons. Chem. Biol. 2012;19:1010–1019. doi: 10.1016/j.chembiol.2012.05.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Neupert S, Johard HA, Nassel DR, Predel R. Single-cell peptidomics of Drosophila melanogaster neurons identified by Gal4-driven fluorescence. Anal. Chem. 2007;79:3690–3694. doi: 10.1021/ac062411p. [DOI] [PubMed] [Google Scholar]
- 17.Bodzon-Kulakowska A, Drabik A, Marszalek M, Kotlinska JH, Suder P. DESI analysis of mammalian cell cultures--preparation and method optimisation. J. Mass Spectrom. 2014;49:613–621. doi: 10.1002/jms.3381. [DOI] [PubMed] [Google Scholar]
- 18.Holscher D, Shroff R, Knop K, Gottschaldt M, Crecelius A, Schneider B, Heckel DG, Schubert US, Svatos A. Matrix-free UV-laser desorption/ionization (LDI) mass spectrometric imaging at the single-cell level: distribution of secondary metabolites of Arabidopsis thaliana and Hypericum species. Plant J. 2009;60:907–918. doi: 10.1111/j.1365-313X.2009.04012.x. [DOI] [PubMed] [Google Scholar]
- 19.Lanni EJ, Dunham SJ, Nemes P, Rubakhin SS, Sweedler JV. Biomolecular imaging with a C60-SIMS/MALDI dual Ion source hybrid mass spectrometer: instrumentation, matrix enhancement, and single cell analysis. J. Am. Soc. Mass Spectrom. 2014;25:1897–1907. doi: 10.1007/s13361-014-0978-9. [DOI] [PubMed] [Google Scholar]
- 20.Parry S, Winograd N. High-resolution TOF-SIMS imaging of eukaryotic cells preserved in a trehalose matrix. Anal. Chem. 2005;77:7950–7957. doi: 10.1021/ac051263k. [DOI] [PubMed] [Google Scholar]
- 21.Colliver TL, Brummel CL, Pacholski ML, Swanek FD, Ewing AG, Winograd N. Atomic and molecular imaging at the single-cell level with TOF-SIMS. Anal. Chem. 1997;69:2225–2231. doi: 10.1021/ac9701748. [DOI] [PubMed] [Google Scholar]
- 22.Heien ML, Piehowski PD, Winograd N, Ewing AG. Lipid detection, identification, and imaging single cells with SIMS. Methods Mol. Biol. 2010;656:85–97. doi: 10.1007/978-1-60761-746-4_4. [DOI] [PubMed] [Google Scholar]
- 23.O'Brien PJ, Lee M, Spilker ME, Zhang CC, Yan Z, Nichols TC, Li W, Johnson CH, Patti GJ, Siuzdak G. Monitoring metabolic responses to chemotherapy in single cells and tumors using nanostructure-initiator mass spectrometry (NIMS) imaging. Cancer Metab. 2013;1:1–4. doi: 10.1186/2049-3002-1-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Lorenzo Tejedor M, Mizuno H, Tsuyama N, Harada T, Masujima T. Direct single-cell molecular analysis of plant tissues by video mass spectrometry. Anal. Sci. 2009;25:1053–1055. doi: 10.2116/analsci.25.1053. [DOI] [PubMed] [Google Scholar]
- 25.Date S, Mizuno H, Tsuyama N, Harada T, Masujima T. Direct drug metabolism monitoring in a live single hepatic cell by video mass spectrometry. Anal. Sci. 2012;28:201–203. doi: 10.2116/analsci.28.201. [DOI] [PubMed] [Google Scholar]
- 26.Bendall SC, Simonds EF, Qiu P, Amir el AD, Krutzik PO, Finck R, Bruggner RV, Melamed R, Trejo A, Ornatsky OI, Balderas RS, Plevritis SK, Sachs K, Pe'er D, Tanner SD, Nolan GP. Single-cell mass cytometry of differential immune and drug responses across a human hematopoietic continuum. Science. 2011;332:687–696. doi: 10.1126/science.1198704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Bandura DR, Baranov VI, Ornatsky OI, Antonov A, Kinach R, Lou X, Pavlov S, Vorobiev S, Dick JE, Tanner SD. Mass cytometry: technique for real time single cell multitarget immunoassay based on inductively coupled plasma time-of-flight mass spectrometry. Anal. Chem. 2009;81:6813–6822. doi: 10.1021/ac901049w. [DOI] [PubMed] [Google Scholar]
- 28.Managh AJ, Edwards SL, Bushell A, Wood KJ, Geissler EK, Hutchinson JA, Hutchinson RW, Reid HJ, Sharp BL. Single cell tracking of gadolinium labeled CD4+ T cells by laser ablation inductively coupled plasma mass spectrometry. Anal. Chem. 2013;85:10627–10634. doi: 10.1021/ac4022715. [DOI] [PubMed] [Google Scholar]
- 29.Groombridge AS, Miyashita S, Fujii S, Nagasawa K, Okahashi T, Ohata M, Umemura T, Takatsu A, Inagaki K, Chiba K. High sensitive elemental analysis of single yeast cells (Saccharomyces cerevisiae) by time-resolved inductively-coupled plasma mass spectrometry using a high efficiency cell introduction system. Anal. Sci. 2013;29:597–603. doi: 10.2116/analsci.29.597. [DOI] [PubMed] [Google Scholar]
- 30.Shrestha B, Vertes A. High-throughput cell and tissue analysis with enhanced molecular coverage by laser ablation electrospray ionization mass spectrometry using ion mobility separation. Anal. Chem. 2014;86:4308–4315. doi: 10.1021/ac500007t. [DOI] [PubMed] [Google Scholar]
- 31.Stolee JA, Vertes A. Toward single-cell analysis by plume collimation in laser ablation electrospray ionization mass spectrometry. Anal. Chem. 2013;85:3592–3598. doi: 10.1021/ac303347n. [DOI] [PubMed] [Google Scholar]
- 32.Shrestha B, Vertes A. In situ metabolic profiling of single cells by laser ablation electrospray ionization mass spectrometry. Anal. Chem. 2009;81:8265–8271. doi: 10.1021/ac901525g. [DOI] [PubMed] [Google Scholar]
- 33.Trouillon R, Passarelli MK, Wang J, Kurczy ME, Ewing AG. Chemical analysis of single cells. Anal. Chem. 2012;85:522–542. doi: 10.1021/ac303290s. [DOI] [PubMed] [Google Scholar]
- 34.Romanova EV, Aerts JT, Croushore CA, Sweedler JV. Small-volume analysis of cell-cell signaling molecules in the brain. Neuropsychopharmacology. 2014;39:50–64. doi: 10.1038/npp.2013.145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Chen R, Li L. Mass spectral imaging and profiling of neuropeptides at the organ and cellular domains. Anal. Bioanal. Chem. 2010;397:3185–3193. doi: 10.1007/s00216-010-3723-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Hummon AB, Amare A, Sweedler JV. Discovering new invertebrate neuropeptides using mass spectrometry. Mass Spectrom. Rev. 2006;25:77–98. doi: 10.1002/mas.20055. [DOI] [PubMed] [Google Scholar]
- 37.Jiménez CR, Li KW, Dreisewerd K, Spijker S, Kingston R, Bateman RH, Burlingame AL, Smit AB, van Minnen J, Geraerts WPM. Direct mass spectrometric peptide profiling and sequencing of single neurons reveals differential peptide patterns in a small neuronal network. Biochemistry. 1998;37:2070–2076. doi: 10.1021/bi971848b. [DOI] [PubMed] [Google Scholar]
- 38.Li KW, Hoek RM, Smith F, Jiménez CR, van der Schors RC, van Veelen PA, Chen S, van der Greef J, Parish DC, Benjamin PR. Direct peptide profiling by mass spectrometry of single identified neurons reveals complex neuropeptide-processing pattern. J. Biol. Chem. 1994;269:30288–30292. [PubMed] [Google Scholar]
- 39.Jiménez CR, Veelen PAv, Li KW, Wildering WC, Geraerts WPM, Tjaden UR, Greef Jvd. Rapid communication: neuropeptide expression and processing as revealed by direct matrix-assisted laser desorption ionization mass spectrometry of single neurons. J. Neurochem. 1994;62:404–407. doi: 10.1046/j.1471-4159.1994.62010404.x. [DOI] [PubMed] [Google Scholar]
- 40.Rubakhin SS, Garden RW, Fuller RR, Sweedler JV. Measuring the peptides in individual organelles with mass spectrometry. Nat. Biotechnol. 2000;18:172–175. doi: 10.1038/72622. [DOI] [PubMed] [Google Scholar]
- 41.Rubakhin SS, Sweedler JV. Characterizing peptides in individual mammalian cells using mass spectrometry. Nat. Protoc. 2007;2:1987–1997. doi: 10.1038/nprot.2007.277. [DOI] [PubMed] [Google Scholar]
- 42.Rubakhin SS, Sweedler JV. Quantitative measurements of cell−cell signaling peptides with single-cell MALDI MS. Anal. Chem. 2008;80:7128–7136. doi: 10.1021/ac8010389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Bai L, Romanova EV, Sweedler JV. Distinguishing endogenous d-Amino acidcontaining neuropeptides in individual neurons using tandem mass spectrometry. Anal. Chem. 2011;83:2794–2800. doi: 10.1021/ac200142m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Jarecki JL, Andersen K, Konop CJ, Knickelbine JJ, Vestling MM, Stretton AO. Mapping neuropeptide expression by mass spectrometry in single dissected identified neurons from the dorsal ganglion of the nematode Ascaris suum. ACS Chem. Neurosci. 2010;1:505–519. doi: 10.1021/cn1000217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Neupert S, Schattschneider S, Predel R. Allatotropin-related peptide in cockroaches: Identification via mass spectrometric analysis of single identified neurons. Peptides. 2009;30:489–494. doi: 10.1016/j.peptides.2008.10.023. [DOI] [PubMed] [Google Scholar]
- 46.Lanni EJ, Rubakhin SS, Sweedler JV. Mass spectrometry imaging and profiling of single cells. J. Proteomics. 2012;75:5036–5051. doi: 10.1016/j.jprot.2012.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Gessel MM, Norris JL, Caprioli RM. MALDI imaging mass spectrometry: spatial molecular analysis to enable a new age of discovery. J. Proteomics. 2014;107:71–82. doi: 10.1016/j.jprot.2014.03.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Heeren RMA. Getting the picture: The coming of age of imaging MS. Int. J. Mass Spectrom. 2014 in press. [Google Scholar]
- 49.Rubakhin SS, Greenough WT, Sweedler JV. Spatial profiling with MALDI MS: distribution of neuropeptides within single neurons. Anal. Chem. 2003;75:5374–5380. doi: 10.1021/ac034498+. [DOI] [PubMed] [Google Scholar]
- 50.Bouschen W, Schulz O, Eikel D, Spengler B. Matrix vapor deposition/recrystallization and dedicated spray preparation for high-resolution scanning microprobe matrix-assisted laser desorption/ionization imaging mass spectrometry (SMALDI-MS) of tissue and single cells. Rapid Commun. Mass Spectrom. 2010;24:355–364. doi: 10.1002/rcm.4401. [DOI] [PubMed] [Google Scholar]
- 51.Guenther S, Rompp A, Kummer W, Spengler B. AP-MALDI imaging of neuropeptides in mouse pituitary gland with 5 mm spatial resolution and high mass accuracy. Int. J. Mass Spectrom. 2011;305:228–237. [Google Scholar]
- 52.Schober Y, Guenther S, Spengler B, Römpp A. Single cell matrix-assisted laser desorption/ionization mass spectrometry imaging. Anal. Chem. 2012;84:6293–6297. doi: 10.1021/ac301337h. [DOI] [PubMed] [Google Scholar]
- 53.Spengler B, Hubert M. Scanning microprobe matrix-assisted laser desorption ionization (SMALDI) mass spectrometry: instrumentation for sub-micrometer resolved LDI and MALDI surface analysis. J. Am. Soc. Mass Spectrom. 2002;13:735–748. doi: 10.1016/S1044-0305(02)00376-8. [DOI] [PubMed] [Google Scholar]
- 54.Zavalin A, Todd EM, Rawhouser PD, Yang J, Norris JL, Caprioli RM. Direct imaging of single cells and tissue at sub-cellular spatial resolution using transmission geometry MALDI MS. J. Mass Spectrom. 2012;47:1473–1481. doi: 10.1002/jms.3108. [DOI] [PubMed] [Google Scholar]
- 55.Zimmerman TA, Rubakhin SS, Sweedler JV. MALDI mass spectrometry imaging of neuronal cell cultures. J. Am. Soc. Mass Spectrom. 2011;22:828–836. doi: 10.1007/s13361-011-0111-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Jurchen J, Rubakhin S, Sweedler J. MALDI-MS imaging of features smaller than the size of the laser beam. J. Am. Soc. Mass Spectrom. 2005;16:1654–1659. doi: 10.1016/j.jasms.2005.06.006. [DOI] [PubMed] [Google Scholar]
- 57.Jungmann JH, MacAleese L, Buijs R, Giskes F, de Snaijer A, Visser J, Visschers J, Vrakking MJJ, Heeren RMA. Fast, high resolution mass spectrometry imaging using a medipix pixelated detector. J. Am. Soc. Mass Spectrom. 2010;21:2023–2030. doi: 10.1016/j.jasms.2010.08.014. [DOI] [PubMed] [Google Scholar]
- 58.Altelaar AFM, Luxembourg SL, McDonnell LA, Piersma SR, Heeren RMA. Imaging mass spectrometry at cellular length scales. Nat. Protoc. 2007;2:1185–1196. doi: 10.1038/nprot.2007.117. [DOI] [PubMed] [Google Scholar]
- 59.Luxembourg SL, Mize TH, McDonnell LA, Heeren RMA. High-spatial resolution mass spectrometric imaging of peptide and protein distributions on a surface. Anal. Chem. 2004;76:5339–5344. doi: 10.1021/ac049692q. [DOI] [PubMed] [Google Scholar]
- 60.Jungmann JH, MacAleese L, Visser J, Vrakking MJJ, Heeren RMA. High dynamic range bio-molecular ion microscopy with the timepix detector. Anal. Chem. 2011;83:7888–7894. doi: 10.1021/ac2017629. [DOI] [PubMed] [Google Scholar]
- 61.Jungmann JH, Smith DF, Kiss A, MacAleese L, Buijs R, Heeren RMA. An invacuum, pixelated detection system for mass spectrometric analysis and imaging of macromolecules. Int. J. Mass Spectrom. 2013;341–342:34–44. [Google Scholar]
- 62.Altelaar AFM, Taban IM, McDonnell LA, Verhaert PDEM, de Lange RPJ, Adan RAH, Mooi WJ, Heeren RMA, Piersma SR. High-resolution MALDI imaging mass spectrometry allows localization of peptide distributions at cellular length scales in pituitary tissue sections. Int. J. Mass Spectrom. 2007;260:203–211. [Google Scholar]
- 63.Jungmann J, Smith D, MacAleese L, Klinkert I, Visser J, Heeren RA. Biological tissue imaging with a position and time sensitive pixelated detector. J. Am. Soc. Mass Spectrom. 2012;23:1679–1688. doi: 10.1007/s13361-012-0444-5. [DOI] [PubMed] [Google Scholar]
- 64.Soltwisch J, Göritz G, Jungmann JH, Kiss A, Smith DF, Ellis SR, Heeren RMA. MALDI mass spectrometry imaging in microscope mode with infrared lasers: bypassing the diffraction limits. Anal. Chem. 2013;86:321–325. doi: 10.1021/ac403421v. [DOI] [PubMed] [Google Scholar]
- 65.Urban PL, Jefimovs K, Amantonico A, Fagerer SR, Schmid T, Madler S, Puigmarti-Luis J, Goedecke N, Zenobi R. High-density micro-arrays for mass spectrometry. Lab Chip. 2010;10:3206–3209. doi: 10.1039/c0lc00211a. [DOI] [PubMed] [Google Scholar]
- 66.Pabst M, Fagerer SR, Köhling R, Küster SK, Steinhoff R, Badertscher M, Wahl F, Dittrich PS, Jefimovs K, Zenobi R. Self-aliquoting microarray plates for accurate quantitative matrix-assisted laser desorption/ionization mass spectrometry. Anal. Chem. 2013;85:9771–9776. doi: 10.1021/ac4021775. [DOI] [PubMed] [Google Scholar]
- 67.Ibáñez AJ, Fagerer SR, Schmidt AM, Urban PL, Jefimovs K, Geiger P, Dechant R, Heinemann M, Zenobi R. Mass spectrometry-based metabolomics of single yeast cells. Proc. Natl. Acad. Sci. U.S.A. 2013;110:8790–8794. doi: 10.1073/pnas.1209302110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Fagerer SR, Schmid T, Ibanez AJ, Pabst M, Steinhoff R, Jefimovs K, Urban PL, Zenobi R. Analysis of single algal cells by combining mass spectrometry with Raman and fluorescence mapping. Analyst. 2013;138:6732–6736. doi: 10.1039/c3an01135f. [DOI] [PubMed] [Google Scholar]
- 69.Tanner SD, Baranov VI, Ornatsky OI, Bandura DR, George TC. An introduction to mass cytometry: fundamentals and applications. Cancer Immunol. Immunother. 2013;62:955–965. doi: 10.1007/s00262-013-1416-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Behbehani GK, Bendall SC, Clutter MR, Fantl WJ, Nolan GP. Single-cell mass cytometry adapted to measurements of the cell cycle. Cytometry Part A. 2012;81A:552–566. doi: 10.1002/cyto.a.22075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Horowitz A, Strauss-Albee DM, Leipold M, Kubo J, Nemat-Gorgani N, Dogan OC, Dekker CL, Mackey S, Maecker H, Swan GE, Davis MM, Norman PJ, Guethlein LA, Desai M, Parham P, Blish CA. Genetic and environmental determinants of human NK cell diversity revealed by mass cytometry. Sci. Transl. Med. 2013;5:208ra145. doi: 10.1126/scitranslmed.3006702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Leipold MD, Ornatsky O, Baranov V, Whitfield C, Nitz M. Development of mass cytometry methods for bacterial discrimination. Anal. Biochem. 2011;419:1–8. doi: 10.1016/j.ab.2011.07.035. [DOI] [PubMed] [Google Scholar]
- 73.Newell EW, Sigal N, Nair N, Kidd BA, Greenberg HB, Davis MM. Combinatorial tetramer staining and mass cytometry analysis facilitate T-cell epitope mapping and characterization. Nat. Biotechnol. 2013;31:623–629. doi: 10.1038/nbt.2593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Han A, Newell EW, Glanville J, Fernandez-Becker N, Khosla C, Chien Y-h, Davis MM. Dietary gluten triggers concomitant activation of CD4+ and CD8+ αβ T cells and γδ T cells in celiac disease. Proc. Natl. Acad. Sci. U.S.A. 2013;110:13073–13078. doi: 10.1073/pnas.1311861110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Newell Evan W, Sigal N, Bendall Sean C, Nolan Garry P, Davis Mark M. Cytometry by time-of-flight shows combinatorial cytokine expression and virus-specific cell niches within a continuum of CD8+ T cell phenotypes. Immunity. 2012;36:142–152. doi: 10.1016/j.immuni.2012.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Chen G, Weng N-p. Analyzing the phenotypic and functional complexity of lymphocytes using CyTOF (cytometry by time-of-flight) Cell. Mol. Immunol. 2012;9:322–323. doi: 10.1038/cmi.2012.16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Dugan CE, Kennedy RT. Measurement of lipolysis products secreted by 3T3-L1 adipocytes using microfluidics. Methods Enzymol. 2014;538:195–209. doi: 10.1016/B978-0-12-800280-3.00011-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Shackman JG, Dahlgren GM, Peters JL, Kennedy RT. Perfusion and chemical monitoring of living cells on a microfluidic chip. Lab Chip. 2005;5:56–63. doi: 10.1039/b404974h. [DOI] [PubMed] [Google Scholar]
- 79.Reid KR, Kennedy RT. Continuous operation of microfabricated electrophoresis devices for 24 hours and application to chemical monitoring of living cells. Anal. Chem. 2009;81:6837–6842. doi: 10.1021/ac901114k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Zhong X, Zhang Z, Jiang S, Li L. Recent advances in coupling capillary electrophoresis-based separation techniques to ESI and MALDI-MS. Electrophoresis. 2014;35:1214–1225. doi: 10.1002/elps.201300451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Nemes P, Knolhoff AM, Rubakhin SS, Sweedler JV. Metabolic differentiation of neuronal phenotypes by single-cell capillary electrophoresis-electrospray ionizationmass spectrometry. Anal. Chem. 2011;83:6810–6817. doi: 10.1021/ac2015855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Nemes P, Knolhoff AM, Rubakhin SS, Sweedler JV. Single-cell metabolomics: changes in the metabolome of freshly isolated and cultured neurons. ACS Chem. Neurosci. 2012;3:782–792. doi: 10.1021/cn300100u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Liu J, Aerts JT, Rubakhin SS, Zhang X, Sweedler J. Analysis of endogenous nucleotides by single cell capillary electrophoresis-mass spectrometry. Analyst. 2014;139:5835–5842. doi: 10.1039/c4an01133c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Aerts JT, Louis KR, Crandall SR, Govindaiah G, Cox CL, Sweedler JV. Patch clamp electrophysiology and capillary electrophoresis-mass spectrometry metabolomics for single cell characterization. Anal. Chem. 2014;86:3203–3208. doi: 10.1021/ac500168d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Mellors JS, Jorabchi K, Smith LM, Ramsey JM. Integrated microfluidic device for automated single cell analysis using electrophoretic separation and electrospray ionization mass spectrometry. Anal. Chem. 2010;82:967–973. doi: 10.1021/ac902218y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Mao S, Zhang J, Li H, Lin J-M. Strategy for signaling molecule detection by using an integrated microfluidic device coupled with mass spectrometry to study cell-to-cell communication. Anal. Chem. 2012;85:868–876. doi: 10.1021/ac303164b. [DOI] [PubMed] [Google Scholar]
- 87.Ngounou Wetie AG, Sokolowska I, Woods AG, Wormwood KL, Dao S, Patel S, Clarkson BD, Darie CC. Automated mass spectrometry–based functional assay for the routine analysis of the secretome. J. Lab. Autom. 2013;18:19–29. doi: 10.1177/2211068212454738. [DOI] [PubMed] [Google Scholar]
- 88.Skalnikova H, Motlik J, Gadher SJ, Kovarova H. Mapping of the secretome of primary isolates of mammalian cells, stem cells and derived cell lines. Proteomics. 2011;11:691–708. doi: 10.1002/pmic.201000402. [DOI] [PubMed] [Google Scholar]
- 89.Iannacone JM, Ren S, Hatcher NG, Sweedler JV. Collecting peptide release from the brain using porous polymer monolith-based solid phase extraction capillaries. Anal. Chem. 2009;81:5433–5438. doi: 10.1021/ac9005843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Vilim FS, Cropper EC, Price DA, Kupfermann I, Weiss KR. Release of peptide cotransmitters in Aplysia: regulation and functional implications. J. Neurosci. 1996;16:8105–8114. doi: 10.1523/JNEUROSCI.16-24-08105.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Lloyd PE, Schacher S, Kupfermann I, Weiss KR. Release of neuropeptides during intracellular stimulation of single identified Aplysia neurons in culture. Proc. Natl. Acad. Sci. U.S.A. 1986;83:9794–9798. doi: 10.1073/pnas.83.24.9794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Dishinger JF, Reid KR, Kennedy RT. Quantitative monitoring of insulin secretion from single islets of Langerhans in parallel on a microfluidic chip. Anal. Chem. 2009;81:3119–3127. doi: 10.1021/ac900109t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Dugan CE, Cawthorn WP, MacDougald OA, Kennedy RT. Multiplexed microfluidic enzyme assays for simultaneous detection of lipolysis products from adipocytes. Anal. Bioanal. Chem. 2014;406:4851–4859. doi: 10.1007/s00216-014-7894-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Wei D, Bailey MJ, Andrew P, Ryhanen T. Electrochemical biosensors at the nanoscale. Lab Chip. 2009;9:2123–2131. doi: 10.1039/b903118a. [DOI] [PubMed] [Google Scholar]
- 95.Wei H, Li H, Gao D, Lin J-M. Multi-channel microfluidic devices combined with electrospray ionization quadrupole time-of-flight mass spectrometry applied to the monitoring of glutamate release from neuronal cells. Analyst. 2010;135:2043–2050. doi: 10.1039/c0an00162g. [DOI] [PubMed] [Google Scholar]
- 96.Yang M, Chao T-C, Nelson R, Ros A. Direct detection of peptides and proteins on a microfluidic platform with MALDI mass spectrometry. Anal. Bioanal. Chem. 2012;404:1681–1689. doi: 10.1007/s00216-012-6257-3. [DOI] [PubMed] [Google Scholar]
- 97.Rubakhin SS, Page JS, Monroe BR, Sweedler JV. Analysis of cellular release using capillary electrophoresis and matrix assisted laser desorption/ionization-time of flightmass spectrometry. Electrophoresis. 2001;22:3752–3758. doi: 10.1002/1522-2683(200109)22:17<3752::AID-ELPS3752>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
- 98.Croushore CA, Supharoek S-a, Lee CY, Jakmunee J, Sweedler JV. Microfluidic device for the selective chemical stimulation of neurons and characterization of peptide release with mass spectrometry. Anal. Chem. 2012;84:9446–9452. doi: 10.1021/ac302283u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Hatcher NG, Richmond TA, Rubakhin SS, Sweedler JV. Monitoring activity-dependent peptide release from the CNS using single-bead solid-phase extraction and MALDI TOF MS detection. Anal. Chem. 2005;77:1580–1587. doi: 10.1021/ac0487909. [DOI] [PubMed] [Google Scholar]
- 100.Fan Y, Rubakhin SS, Sweedler JV. Collection of peptides released from single neurons with particle-embedded monolithic capillaries followed by detection with matrixassisted laser desorption/ionization mass spectrometry. Anal. Chem. 2011;83:9557–9563. doi: 10.1021/ac202338e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Fan Y, Lee CY, Rubakhin SS, Sweedler JV. Stimulation and release from neurons via a dual capillary collection device interfaced to mass spectrometry. Analyst. 2013;138:6337–6346. doi: 10.1039/c3an01010d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Zhong M, Lee CY, Croushore CA, Sweedler JV. Label-free quantitation of peptide release from neurons in a microfluidic device with mass spectrometry imaging. Lab Chip. 2012;12:2037–2045. doi: 10.1039/c2lc21085a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Hatcher NG, Atkins N, Annangudi SP, Forbes AJ, Kelleher NL, Gillette MU, Sweedler JV. Mass spectrometry-based discovery of circadian peptides. Proc. Natl. Acad. Sci. U.S.A. 2008;105:12527–12532. doi: 10.1073/pnas.0804340105. [DOI] [PMC free article] [PubMed] [Google Scholar]



