Open access peer-reviewed chapter

Advanced Microfluidic Assays for Caenorhabditis elegans

Written By

Natalia A. Bakhtina, Neil MacKinnon and Jan G. Korvink

Submitted: 22 May 2015 Reviewed: 18 May 2016 Published: 23 November 2016

DOI: 10.5772/64283

From the Edited Volume

Advances in Microfluidics - New Applications in Biology, Energy, and Materials Sciences

Edited by Xiao-Ying Yu

Chapter metrics overview

2,482 Chapter Downloads

View Full Metrics

Abstract

The in vivo analysis of a model organism, such as the nematode Caenorhabditis elegans, enables fundamental biomedical studies, including development, genetics, and neurobiology. In recent years, microfluidics technology has emerged as an attractive and enabling tool for the study of the multicellular organism. Advances in the application of microfluidics to C. elegans assays facilitate the manipulation of nematodes in high-throughput format and allow for the precise spatial and temporal control of their environment. In this chapter, we aim to illustrate the current microfluidic approaches for the investigation of behavior and neurobiology in C. elegans and discuss the trends of future development.

Keywords

  • C. elegans
  • chip-based
  • manipulation
  • microfluidics
  • model organism

1. Introduction

The invertebrate Caenorhabditis elegans, Drosophila melanogaster, and the vertebrate zebrafish (Danio rerio) are the most widely studied multicellular organisms. The in vivo analysis of these model organisms allows the understanding of many complex physiological processes, addressing many of the questions relevant to human biology. The choice of model organism depends on the biological question under investigation. For example, C. elegans is simple enough to be experimentally tractable. It has a short life cycle (3 days at 25 °C) and lifespan (15–17 days at 25 °C), passes through four larva (L1–L4) stages and an adult stage [1]. Its small size (1–1.2 mm long and 80 µm wide), transparent body at all life stages, and preferred food source (Escherichia coli) simplify its maintenance on agar plates or liquid cultures allowing visualization of individual cells and organs in intact animals. C. elegans possesses one of the simplest central nervous systems (the adult hermaphrodite has 302 neurons). Because it is so well studied, rapid identification of signaling pathways, for instance, in studies of aging, has become possible. About 40 % of human disease genes have an orthologue in the genome of C. elegans, including those genes associated with Alzheimer disease, Parkinson’s disease, Huntington’s disease (HD), and many other neurodegenerative disorders [13]. This astonishing degree of correspondence permits the modeling of human ailments in a simple invertebrate without involving actual human subjects and provides a meaningful insight into the pathogenesis of a complex disease phenotype.

Traditionally, behavioral genetics is employed as a prime method for neurobiological studies in C. elegans. It is based on manual worm manipulation on a Petri dish or a multiwell plate, and monitoring the effects on various biological processes, such as growth and fertility, by visual inspection. Refreshment of old buffer solutions by a fresh solution is invasive and causes stress both to the larvae and adults. For drug screening, the concentration of the active compounds in the exposing solution might not be precisely controlled because of evaporation and non-selective adsorption effects on the wall of the wells. Permanent immobilization of the worm for further neuronal analysis is performed by means of glue or anesthetics. These methods are time consuming, expensive, tedious, prone to human flaws, and frequently result in failure. To address these problems, novel technologies for the manipulation of multicellular organisms are needed.

Microfluidics has recently been adopted as an instrument both to expand and accelerate progress related to the treatment of human diseases and injuries. Due to precise and automated manipulation of fluids and samples (e.g., single cell, multicellular organism, etc.) in a system of channels (10 –150 µm), a microfluidic-based approach is able to open up aspects that would remain hidden from traditional laboratory techniques. The technology provides a junction between engineering and pure sciences with an immense potential for offering simple and practical solutions. The unique properties of this technology are highlighted by several aspects. First, the dimensions of microfluidic channels perfectly match to the size of samples, allowing precise manipulation. With moving parts, flowing fluids, or other passive mechanisms, microsystems can be used to align samples with a particular orientation with ease as compared to hand-manipulations. Second, the ability to manipulate small amounts of liquid makes it suitable for the precise delivery of small amounts of reagent. Due to the laminar nature of the flow at the micro scale, efficient mass and energy transfer can be controlled in a completely predictable manner (e.g., diffusion of dissolved gases across tens of microns through fluids or polymer membrane materials). Third, based on relatively inexpensive polymer-based fabrication techniques, such as polydimethylsiloxane (PDMS) replica molding, it has become feasible to realize disposable, economic, and biocompatible systems [4]. Complex structures, adapted to different applications, can be easily fabricated in a short time. Finally, the capability to realize large-scale integration makes it possible to handle a large population of samples in parallel or in series for high -throughput assays. For example, COPAS BIOSORT high-throughput analysis system from Union Biometrica, Inc. (USA), enables the performing of high-speed imaging and offers the possibility of studying a large quantity of individual worms, thereby providing detailed statistical information on the biological variance within the same population.

Substantial advances in microfluidic techniques and particular research interest in C. elegans have driven the development of numerous microchip-based systems. They have been reviewed a number of times focusing on various aspects of miniaturized systems, their advantages, application challenges, and scientific potential [519]. A summary of microfluidic-based systems with respect to the organism, organ, or tissue studies was presented by Sivagnanam et al. [5]. All available on-chip approaches for C. elegans investigations were systemized by the authors [6]. A classification diagram for structuring of approximately 100 references that simplifies their search according to five evaluated aspects (measured output data, and method for sorting, immobilization, stimulation, and detection of C. elegans) is included. In addition, we listed the relevant sorting, immobilization, and imaging methods that have been reported in recent literature, and indicated the main qualitative and quantitative characteristics for each.

This chapter provides a comprehensive overview of recent microfluidic-based approaches for investigations of worm behavior and neurobiology (Figure 1). This includes a discussion on tools and approaches needed to ensure high-throughput manipulation (culturing, sorting, and immobilization) and assaying for behavioral and neuronal studies. In addition, a perspective of novel methods for studies of metabolic activity facilitated by microfluidics is presented.

Figure 1.

Schematic illustration of a microfluidic platform for C. elegans assaying. The structure of this chapter is as follows: in Section 2 we summarize the most frequently utilized imaging (A) and detection (B) methods; in Section 3 we concentrate our attention on different techniques for worm-on-a-chip manipulation, such as culturing, sorting, and immobilization (C, D, and E); Section 4 provides an overview of behavioral and neuronal phenotypes of C. elegans facilitated by different analysis techniques, including electrochemical impedance spectroscopy (F), microsurgery (G), and microinjection (H); additionally, Section 5 presents the methods utilized for metabolic activity studies (I).

Advertisement

2. Microfluidic approaches for C. elegans detection

Conventional optical imaging is an established detection technique for the observation of biological samples (e.g., cells, microorganisms, etc.). Microscopy-based (fluorescent, nonfluorescent, or their various combinations) systems can be used to extract valuable and unique data (e.g., image the activity of specific neurons) from biological samples. Combined with microfluidics, these systems offer several important advantages required for high-throughput screening [710]. Fully automated components, software control, and image processing tools make commercial confocal microscopes extremely versatile for real-time and high -resolution diagnosis. However, conventional optical imaging systems are quite expensive, bulky, and limit the miniaturization of chip-based systems. An overview of different optical imaging approaches in microfluidics (e.g., conventional optical imaging, lensless imaging, etc.) and their applications was recently presented by several research groups [2022].

To overcome limitations mentioned above, researchers utilize on-chip or lensless imaging technologies. On-chip imaging systems for C. elegans, including contact optofluidic imaging [23, 24], direct shadow imaging [25], holographic imaging [2628], in combination with automated data processing have enabled the observation and characterization of key behavioral parameters in vivo at micrometer and nanometer resolution (Figure 1A and B). Lensless imaging has the advantage of cointegration of microfluidics, microelectronics, and optical components into one platform. This has guaranteed an increase of the image quality, and has provided an ultimate spatial resolution of approximately 0.9 µm and a throughput approaching 40 worms per minute [23]. The combination with fluorescence imaging holds a great potential for screening of cellular processes [28].

Another promising approach is a lensless and sensor-less monitoring of the nematodes’ movement in various microenvironments [29]. In a micro-electro-fluidic (MEF) grid, a moving nematode is detected by change in the electrical impedance at the intersection regions of the microelectrode grid, formed by two identical orthogonally arranged arrays of metal lines (Figure 1B). The approach ensured the real-time readout of the crawling nematode with a spatial resolution of 30 µm (the distance between grid lines) of the reconstructed images at the frequency of 174 Hz per readout.

Usually, the use of fluorescence-based techniques, such as calcium imaging or green fluorescent protein (GFP) expression, and microfluidics to image the activity of specific neurons requires chemically or genetically labeled animals to be immobilized for imaging at a cellular level [3032]. An “immobilization-free” approach detection is achieved via two pairs of integrated optical fibers. Through the measurements of optical density and fluorescence, the fibers can detect and differentiate wild-type and green fluorescent protein (GFP)-type C. elegans even when they flow at high speeds (switching time of 1 s per worm) [33]. This has proven to be a well-controlled method for automated handling of worms in a high-throughput manner with a sorting accuracy of more than 96 %.

Advertisement

3. Microfluidic techniques for C. elegans manipulation

Environmental control and manipulation of whole animal poses significant challenges (e.g., animal’s body orientation, precise delivery of chemicals, etc.). Transferring traditional neurobiology and behavioral investigation techniques to the microfluidic platform has the potential to overcome these challenges. This is driven by substantial progress in integration of functional components (e.g., valves, detectors, etc.) that allow the monitoring of various steps, such as administration, distribution, metabolism, and toxicity during drug screening. The advanced microfluidic approach offers both qualitative and quantitative data from a single organism by automatic high-throughput manipulation. For example, the worms can be oriented at regular positions on a substrate due to hydrodynamic forces in a microfluidic chip for the determination of gene function in a high-throughput manner [34]. In this section, we discuss general manipulation techniques, such as culturing, sorting, and immobilization.

Culturing. To interpret the underlying metabolic changes and specific developmental processes during nematode ageing, longitudinal experiments over the entire lifespan are necessary. Imaging and monitoring of the embryogenesis require specific techniques, which include single embryo isolation and mounting. Worm culturing can be dramatically improved using an automated microfluidic platform for culturing, phenotyping, and long-term live imaging of C. elegans embryo and larvae using microfluidic chambers (Figure 2A) and droplet encapsulation (Figure 2B) [3547]. In combination with image recognition algorithms these “worm-chips” have successfully demonstrated their high potential at enhancing worm handling (e.g., automatic nutrient and waste exchange), accurate imaging, and automated analysis of embryonic morphogenesis during embryonic development [48]. Requiring the loading of only a few adult worms into the chip, the ensuing C. elegans embryo population could be processed at the same time.

Figure 2.

Schematic illustration of the microfluidic platforms for C. elegans culturing (A) [35] and sorting based on droplet encapsulation (B) [4244] or electrotaxis (C) [49].

Sorting. C. elegans exhibits age-dependent specific neuron and behavioral responses. For instance, usage of both the early-stage and adult worms may increases the physiological relevance of drug candidates during the identification process and reveals potential toxic effects. Therefore, sorting (age or size synchronization) of worm populations or individuals is often required for further diagnostics. For example, a passive sorting method is based on self-regulated worm distribution and loading into an array of narrowing channels [36, 50, 51] or microchambers [36, 37, 40, 5255] with an average loading effectiveness rate of approximately 65 %. Once the worm enters the microchannel, the hydrodynamic resistance increases dramatically, thereby locking a single worm inside the chamber.

When considering high-throughput manipulation, automatic classification of worms (e.g., wild-type from mutants) becomes of high relevance. Typically, sorting involves individual C. elegans loading and separation, for instance, according to genetic phenotype for downstream analysis [31, 32, 34, 36, 5660]. Together with real-time rapid image extraction and data processing, media flow in the microfluidic channel is driven by a syringe and is controlled by on-chip functional components, such as PDMS valves. Automatic sample positioning can guarantee rapid classification based on synaptic characteristics with sorting throughput at a rate of 900 worms per hour and an overall sorting accuracy of 96.5 % [32]. Depending on the extracted data, the worm could be flushed to either a waste or a sorting outlet by valve actuation.

Several other techniques have been successfully implemented in high-throughput studies [6163]. In these systems, sorting is accomplished based on size difference in a passive, but extremely high throughput (up to 1200 worms per min) and selective manner (94 % of adults with 0.2 % larva contamination) [62].The device body contains an array of microstructured post (or filters) and a network of microfluidic channels allowing a large population of adult worms and larvae to be oriented in the desired direction.

Immobilization. Because of the high mobility, C. elegans immobilization is required for further developmental studies (e.g., neurosurgery). Manual handling and immobilization (e.g., by gluing or anesthesia) suffer from low throughput and is limited in long-term monitoring [1]. Immobilization in a microfluidic channel provides a simple well-controlled mechanism for automated handling of worms in a high-throughput manner. One of the widely used techniques is based on mechanical force. Several immobilization techniques are used to lock an individual worm against a microchannel wall in a robust and reversible manner. The first method involves microarrays of fixed-geometry clamps for C. elegans immobilization, which is a simple way to restrict the motion employing a single PDMS layer [36, 51, 6467]. Using a constant pressure difference between the inlet and outlet of the device to drive fluid flow, an array of up to 128 wedge-shaped microchannels can be filled by nematodes with up to 90 % efficiency [68]. The second method includes a flexible PDMS membrane for squeezing the worm into the side of microfluidic channel under an external pressure through a control layer above the main chamber [51, 5759, 67, 69, 70]. A third method is to trap of C. elegans by suction flow (Figure 1E), which is based on a vacuum-assisted restraint that aligns the worm along its axis [31, 71]. To highlight the utility of both immobilization techniques several research groups have combined suction posts with either flexible membranes [56, 65, 72, 73] or microchannel narrowing [30, 7477] for extremely stable immobilization during microsurgery (Figure 1G).

Microfluidic devices offer advantages for both spatial and temporal control of the animal’s position and microenvironment at the microscale. Based on acoustic wave in a single-layer microfluidic chip, on-chip manipulation technique permitted trapping and rotational manipulation C. elegans regardless of shape and physical properties in the x- or y-directions for extended periods of time without inducing physiological damage [78, 79]. By implementation of a cooling liquid supply through a control layer to lower a worm's temperature down to 4 °C, C. elegans can be immobilized with a throughput up to 400 worms per hour for short-term cooling (~2 s) [32, 52]. Alternatively, light-induced sublethal heat can be used to increase the worm temperature (up to 31–37 °C) for its immobilization [80].

Gases, including carbon dioxide (CO2) and nitrogen (N2), are sensed by C. elegans and serve as a partial and complete method to eliminate worm mobility [67, 69, 76]. By passing pure gas from a control layer into a flow layer microenvironment, C. elegans can be immobilized in a channel with improved sensitivity and increased resolution.

Advertisement

4. Microfluidic approach for C. elegans sensing function and behavior

In this section, we review the use of microfluidic chips for C. elegans investigations under controlled physical and chemical conditions that have been advantageously used, for example, as integrated biosensors for toxicological experiments and drug screening. The two major methods of assaying are behavioral and molecular (or neuronal) studies. In behavioral studies, discussed in Section 4.1, the movement is generally analyzed by observation the animal’s behavior in response to stimulation (e.g., touch, drugs, odorants, food, temperature, gases, osmolytes, or light). The key issues in neuronal studies, covered by Section 4.2, are the intracellular processes and neuronal signaling.

4.1. Behavioral studies of C. elegans’s responses to different stimuli

C. elegans explores its surrounding environment and moves according to environmental stimuli, including temperature, chemical, electric field, and light, which are detected by 24 sensilla organs and various isolated sensor neurons [1]. Obtaining meaningful data about the mechanism of environmental sensing requires strict control over the experimental conditions. Moreover, when a high number of identical biological samples are needed to be screened, a common need and challenge of the experimental procedure is the precise manipulation of worms with an emphasis on high throughput. Microfluidics offers a straightforward solution for automation and parallelization of screening in a rapid, sensitive, and accurate manner.

The environmental cues can be applied by devices, embedded in a chip-based microfluidic system, to analyze the behavioral response of the microorganism. For example, active and automated local manipulation and chemical stimulation of the individual worms can be achieved by implementation of multilayer PDMS layers. Because of C. elegans’ small size and its ability to grow in liquid, on-chip imaging systems and automated data processing facilitate the observation and characterization of key behavioral parameters in vivo with micrometer and nanometer resolution.

Locomotion. Owing to the precision achievable by microlithographic techniques, researchers have been able to investigate in-depth different locomotion patterns (by varying the size and spacing of posts), applied muscular forces (by measuring the deflection of posts) (Figure 1D) and motility quantification (time-averaged kinetic power over the swimming cycle) [8190]. In contrast to traditional experimental techniques, these systems permitted straightforward dynamic force detection of moving nematodes. Whereas the motion of the animals in the artificial soil device exhibited the same principal characteristics of the motion of crawling on agar, the constraints on motion caused by the posts better mimicked the complexity of C. elegans’ natural environment [91]. To further study crawling behavior, a number of PDMS microfluidic devices were presented that consisted of sinusoidal channels of varying wavelengths [84, 86, 87].

These devices allow researchers to manipulate the oscillating body motion of the crawling animals and investigate the biophysical and neuronal mechanisms of locomotion and proprioception. Microfluidics facilitates precise environmental control that was demonstrated by modifying the chemicals’ concentration of the main chamber rapidly or and immediately observing the effect on locomotion [83]. Obviously, tracking animals through such a rapid media exchange would not be possible in a larger environment.

Electrotaxis. C. elegans exhibits responsive behavior to electric fields, mediated by certain amphid sensory neurons. These neurons are sensitive to both the direction and strength of the electrical signal, and forced the animal to move toward a negatively charged pole [1]. Analysis of the nematode’s electrotaxis provides a detailed model of how neurons function together to generate a behavioral response to electric fields. When microfluidic chambers are combined with electrodes to deliver electrical stimuli, both behavioral and neuronal screening can be performed, providing the chance to elucidate potential treatment for human muscular disorders. Many microfluidic systems were proposed for fully automated control of electrotaxis, which overcame many of the inherent problems of manual operation [49, 9298].

Normally, worms are exposed to a uniform electric field generated by two electrodes (e.g., platinum wires) embedded in inlet and outlet reservoirs and connected to external electrical drive circuitry (Figure 2C). Exposure to direct (DC), alternating (AC), and pulsed DC electric fields in a specified range of strengths has been employed as a means of guiding nematodes in a binary manner (e.g., start and stop), for sorting, and for immobilization, aiming to provide a close look at the mechanism of neuronal signaling transduced into behavioral responses [86, 92, 93, 98100]. Such movement-based microfluidic devices permit the differentiation of worms according to locomotive abilities and similar physiological states, for instance, to distinguish adults from larva, or healthy worms from uncoordinated, and to locate individuals defective in electric field sensing. This guiding technique allows high throughput (up to 60 worms per min) and method selectivity of 70–90 %.

The progress achieved in microfabrication technologies has made monolithic integration of electrodes into microfluidic platform possible (Figure 1F). Micropatterned electrodes on the sidewalls of microfluidic channels (i.e., without blocking optical visibility) provides a simple means of creating electrofluidic glass chips to flexibly control the movement of C. elegans in a sensitive and reproducible manner [101]. Placing the microelectrodes inside the microfluidic environment as close to the animal as possible allows one to create transient pores in the cell membrane, which permits the diffusion of extracellular compounds that are present in the vicinity of the pore into the interior of the cell [102]. All of these results demonstrate the potential of using active microfluidic devices as an alternative to Petri dishes for C. elegans assays.

Chemicals. Microfluidics is particularly attractive for many applications where C. elegans are used as integrated biosensors for toxicological experiments and drug screening. Behavioral investigations in response to chemical stimuli include real-time locomotion diagnostics of C. elegans. The effect on worm physiology to a variety of anesthetics, such as tricaine, muscimol, sodium azide, and levamisole [29, 55, 69, 93], odors, such as hermaphrodite-conditioned media and nicotine, and odors produced by pathogenic bacteria [55, 103108], chemicals, such as zinc ion (Zn2+) and glucose [109], different osmolarity levels [66, 71], was successfully examined by precise chemical control in a time- and dose-dependent manner. In most cases, pre- and postexposure locomotion phenotypes are compared by a variety of parameters (e.g., average velocity, individual head swing orientations, etc.).

Other stimuli. The ability of integration and individual worm manipulation makes microfluidic devices attractive platforms for understanding the correlation between C. elegans’ neuronal and behavioral responses. Based on the properties of a microfluidic device, temperature stimuli could be delivered to individual worms accurately by flexible chip design and fluidic manipulation. Behavioral mechanisms in response to temperature change is quantified in terms of an average head angle of a semi-restrained animal [74] or swimming movements of the individuals heated in a microdroplet array [107].

Although the C. elegans has no light-sensing organs, it modulates a response to light known as phototaxis [50, 92, 109]. To analyze C. elegans' sensitivity to light, wild-type and mutant nematodes are illuminated with light and their behavioral response are examined. It was experimentally demonstrated that illumination to green light is preferable for animals, while blue light triggers muscle depolarization and further body contraction.

4.2. Neuronal studies

Behavioral studies, such as physiological responses, in a whole organism population include not only movement-based analyses but also monitoring of the C. elegans’ neuronal activity in a confined space. Coupled with microfluidic-based systems, existing neuronal recording techniques (e.g., by calcium imaging or green fluorescent protein (GFP) expression) examine neuronal responses to sensory inputs of a single animal at a time under precise environmental control. For example, it was found that immobilizing a portion of the worm can directly override rhythmic activity and may cause changes in transport parameters of the touch neuron [69, 110]. In order to explore locomotive behavior and the underlying molecular mechanism, Wang et al. monitored a subcellular distribution of the DAF-16 gene that regulates different stress responses [91]. The experiments showed an increase of DAF-16 nuclear localization, attributed to crowding stress, in a microcolumn array with intervals from 40 to 200 µm between microposts (Figure 1D). As a result, a system-level understanding of the worm's motor circuit can be obtained.

One application where microfluidics and fluorescent-based imaging open up aspects that would remain hidden from traditional laboratory techniques is drug screening. C. elegans can be an effective test-bed for a wide range of water-soluble chemical compounds (e.g., glycerol [30, 66, 74, 75], anticancer drugs [48], heavy metals [54], sodium chloride NaCl [58, 65, 71, 83, 111, 112], copper(II) chloride CuCl2 [66, 74], levamisole [70], manganese [102], antibiotics [104], isoamyl alcohol [113], cyanide [114], etc.). Microfluidic network manipulation allows the automation in a high-throughput manner and under reproducible experimental conditions while analysis of the nematode's chemosensitivity provides a detailed model of how neurons function together to generate behavioral response. For example, neurotransmitters and hormones, such as 1-methyl-4-phenylpyridinium (MPP+), 6-hydroxy dopamine (6-OHDA), and rotenone, have widespread effects as chemical regulators for coordinating physiological activity throughout the body of both nematodes and humans [1]. The microfluidic-based experiments proved that MPP +, 6-OHDA, and rotenone induce mobility defects in the animal (i.e., significant reduction in speed) after treatment and was potentially neurotoxic for dopaminergic neurons [43, 51, 112].

Due to PDMS microfluidic devices, much progress has been made to overcome the limitations of precise chemical control. The effect of ageing on physiological properties of the ASH chemosensory neuron can be characterized and quantified by the direct delivery of a chemical odor to the nose of C. elegans [30]. To emphasize the influence of different anesthetics on subcellular activity, a microfluidic platform was used for studying the contribution of vesicle transport to synaptic growth [70]. As a result, imaging of subcellular processes, such as pre-synaptic vesicle transport, intraflagellar transport (IFT), dendritic transport, and migration of neuroblasts during early developmental stages of the nematode, has become feasible. Monitoring of neuron activity (e.g., ASH neurons) with respect to osmotic gradient, can access the pattern-generating activity (e.g., individual head swing orientations) of the chemosensory circuit [66].

Another field where polymer-based fabrication techniques have already demonstrated themselves, is in investigations of gas sensing in nematodes [69, 76, 115117]. In order to understand how oxygen level variation causes behavioral and physiological changes, freely moving adult animals were subjected to a gas-phase oxygen gradient. Experiments showed that specific soluble guanylate cyclase homologues (GCY-31, GCY-33, GCY-35, and GCY-36), located in URX, AQR, and PQR sensory neurons, activate hypoxia or hyperoxia avoidance [115, 116].

For many applications, such as characterizing stochastic neural responses, it should be beneficial to increase experimental throughput at the expense of image resolution. Microfluidics promotes simultaneous recording of calcium transients in individual neurons from multiple animals (up to 20), and increases experimental throughput [82, 118]. Thus, a systematic characterization of chemosensory neuron responses to multiple odors, odor concentrations, and temporal patterns, as well as responses to pharmacological manipulation can be performed.

The described experiments benefit enormously from the use of microfluidic technologies. The precise handling and chemical mixing of chemicals and neurotoxins in nanoliter volume droplets tremendously decreases reagent consumption and reaction time. The combination of brightfield imaging, fluorescent imaging, and microfluidics allows in vivo observation of biomolecules and automated analysis of protein aggregation phenomena in C. elegans for amyotrophic lateral sclerosis (ALS) at unprecedented resolution [119]. The level of precision that researchers have already achieved demonstrates the potential for the dissection of neuronal function and toxin-induced neurodegeneration in vivo.

4.2.1. Intracellular studying techniques

Several other techniques for studying and characterization of intracellular processes, including dielectrophoresis and electrochemical impedance spectroscopy, have been adopted by researchers for whole-animal drug screening (Figure 1F) [77, 92, 120, 121]. These methods use a noninvasive electrophysiological readout of neuromuscular function and can provide high-quality neurogenetic and neuropharmacological data on nematodes. Automatic real-time monitoring and parallelization (up to 8 worms simultaneously) with throughput of up to 12 worms per hour facilitate the rapid neuroactive drug screening, e.g., effects of drugs on neurons, as well as on muscles [77, 121].

4.2.2. Microsurgery and microinjection

In combination with microfluidics and optical image analysis systems, microsurgery and microinjection are employed for in vivo neuronal regeneration and cell-to-cell communication studies [52, 73, 80, 122127]. Because the nervous system is described in great detail, the role of an individual neuron can be directly studied with laser ablation experiments (Figure 1G). Using a laser nanobeam in the UV wavelength region, fluorescent-labeled whole cell ablation is possible and a single synapse removal can be achieved [124127]. Afterward, the resulting phenotypes (e.g., degeneration and regeneration) can be examined in vivo. Advances in optical imaging and microfluidic methods support this procedure. The capabilities of current chip-based systems are sufficient to perform precise animal manipulations, required for high immobilization stability of the worm, and complex image-based assaying with high throughput (up to 200 animals per hour with a success rate of 89 %) [52, 122]. This provides approximately one order of magnitude improvement over manually performed axotomies (when considering study of a single population) and gives an opportunity to perform genetic screening in a reasonable timeframe to identify the molecular mechanisms involved in nerve regeneration and degeneration.

The in vivo injection of chemical materials that have significant implications in genetics, drug discovery, and other biological applications is another way to study the mechanisms underlying intercellular communication in C. elegans (Figure 1H). Using a single needle tip of the micromanipulator, localized chemical stimulation can be delivered to a single intestinal cell of the immobilized worms [123, 128].

Advertisement

5. Conclusions and perspectives

The advances in microfabrication technologies have demonstrated the potential of using active lab-on-a-chip (LoC) devices as an alternative to microwell plates for worm-based assays. LoC technology offers a straightforward solution to all of the problems during manual manipulation. Complex three-dimensional (3D) microenvironments have been created, where a whole population of worms is cultured and analyzed in a reproducible way. Currently available microfluidic-based systems are capable of recording from sensory neurons in animals in vivo, whose neuronal responses could be correlated with behavior. Microsurgery and microinjection allow the investigation of many processes, including the role of individual neurons in neuronal networks and in cell-to-cell interaction. Obviously, this is pushing forward fundamental studies in biology and biochemistry.

The use of fluorescence-based techniques and microfluidics to image the activity of specific neurons requires that animals be labeled either chemically or genetically. However, for monitoring certain biological processes, fluorescent labeling might be inconvenient or may interfere with normal behavior. Moreover, many dynamic phenomena of motile samples might be missed during impedance spectroscopy, microsurgery, and microinjection because of the long-term immobilization required for subcellular-level stabilization of C. elegans. This makes monitoring of actual metabolic activity impossible.

Several other approaches can be used to study the neuronal and metabolic activity of a biological system. For example, nuclear magnetic resonance imaging (MRI) and nuclear magnetic resonance spectroscopy (NMR) are two of the most information-rich methods that provide a unique opportunity to link morphological, functional, and chemically specific spectroscopic information from small volume (e.g., µl) samples (Figure 1I). MRI and NMR uses strong time-varying radio frequency (RF) fields to generate a weak specific RF response from a certain tissue type [129]. Because the technology is noninvasive and only nonionizing radiation is absorbed and emitted, it might be especially suitable for the study of C. elegans in the identification and quantification of metabolites (intermediate products of metabolism) within the metabolic pathway in vivo [130136].

In many of the reviewed research articles, the easy integration of microfluidic control and detection modules was a key factor in helping to link in vitro and ex vivo experimental investigations. The ability of C. elegans tracking in real time (i.e., with minimal latency) for further diagnostic applications could provide a close look at the cellular, molecular, and genetic levels. Consequently, an understanding of the underlying molecular mechanisms in multicellular model organisms would provide a unique opportunity to unthread analogous and complex biological processes in humans. This certainly will promote more automated and higher throughput applications in the future.

Advertisement

Acknowledgments

We gratefully acknowledge financial support from the European Research Council (ERC) (contract number 290586 from 1.07.2012), which funded this work.

References

  1. 1. Riddle DL, Blumenthal T, Meyer BJ, Priess JR. C. elegans II, 1997.
  2. 2. Markaki M, Tavernarakis N. Modeling human diseases in Caenorhabditis elegans. Biotechnol. J. 2010, 5: 1261–1276. DOI: 10.1002/biot.201000183.
  3. 3. Alexander AG, Marfil1 V, Li C. Use of Caenorhabditis elegans as a model to study Alzheimer’s disease and other neurodegenerative diseases. Front Genet. 2014, 5: 1–21. DOI: 10.3389/fgene.2014.00279.
  4. 4. Duffy DC, McDonald JC, Schueller OJA, Whitesides G. Rapid prototyping of microfluidic systems in poly(dimethylsiloxane). Anal. Chem. 1998, 70: 4974–4984. DOI: 10.1021/ac980656z.
  5. 5. Sivagnanam V, Gijs MAM. Exploring living multicellular organisms, organs, and tissues using microfluidic systems. Chem. Rev. 2013, 113: 3214–3247. DOI: 10.1021/cr200432q.
  6. 6. Bakhtina NA, Korvink JG. Microfluidic laboratories for C. elegans enhancefundamental studies in biology. RSC Adv. 2014, 4: 4691–4709. DOI: 10.1039/c3ra43758b.
  7. 7. Ben-Yakar A, Chronis N, Lu H. Microfluidics for the analysis of behavior, nerve regeneration, and neural cell biology in C. elegans. Curr. Opin. Neurobiol. 2009, 19: 561–567. DOI: 10.1016/j.conb.2009.10.010.
  8. 8. Chronis N. Worm chips: microtools for C. elegans biology. Lab Chip 2010, 10: 432–437. DOI: 10.1039/b919983g.
  9. 9. Taylor AM, Jeon NL. Micro-scale and microfluidic devices for neurobiology. Curr. Opin. Neurobiol. 2010, 20: 640–647. DOI: 10.1016/j.conb.2010.07.011.
  10. 10. Crane MM, Chung K, Stirman J, Lu H. Microfluidics-enabled phenotyping, imaging, and screening of multicellular organisms. Lab Chip 2010, 10: 1509–1517. DOI: 10.1039/b927258e.
  11. 11. Jian-Ping J, Fan Y, Xin-Chun L, Yan-Yan Y, Zuan-Guang C. Advances on biomedical research in Caenorhabditis elegans based on microfluidic device. Prog. Biochem. Biophys. 2011, 38: 877–883. DOI: 10.3724/SP.J.1206.2011.00079.
  12. 12. Shi W, Wen H, Lin B, Qin J. Microfluidic platform for the study of Caenorhabditis elegans. Top. Curr. Chem. 2011, 304: 323–338. DOI: 10.1007/128_2011_145.
  13. 13. Yanik MF, Rohde CB, Pardo-Martin C. Technologies for micromanipulating, imaging, and phenotyping small invertebrates and vertebrates. Annu. Rev. Biomed. Eng. 2011, 13: 185–217. DOI: 10.1146/annurev-bioeng-071910-124703.
  14. 14. Xu X, Kim SK. The early bird catches the worm: new technologies for the Caenorhabditis elegans toolkit. Nat. Rev. Genet. 2011, 12: 793–801. DOI: 10.1038/nrg3050.
  15. 15. Wlodkowic D, Khoshmanesh K, Akagi J, Williamsand DE, Cooper JM. Wormometry-on-a-chip: innovative technologies for in situ analysis of small multicellular organisms. Cytometry Part A 2011, 79: 799–813. DOI: 10.1002/cyto.a.21070.
  16. 16. San-Miguel A, Lu H. Microfluidics as a tool for C. elegans research, in WormBook, ed. The C. elegans Research Community, 2013. DOI: 10.1895/wormbook.1.162.1.
  17. 17. Rezai P, Salam S, Selvaganapathy PR, Gupta BP. Microfluidic systems to study the biology of human diseases and identify potential therapeutic targets in C. elegans, in Integrated Microsystems, ed. Iniewski K, CRC Press, 581–608; 2011.
  18. 18. Hui W, JianHua Q. Analysis of Caenorhabditis elegans in microfluidic devices. Sci. China: Chem. 2012, 55: 484–493. DOI: 10.1007/s11426-012-4541-x.
  19. 19. Wlodkowic D, Khoshmanesh K, Akagi J, Williams DE, Cooper JM. Wormometry-on-a-chip: innovative technologies for in-situ analysis of small multicellular organisms. Cytometry, Part A 2011, 79: 799–813. DOI: 10.1002/cyto.a.21070.
  20. 20. Gurkan UA, Moon S, Geckil H, Xu F, Wang S, Lu TJ, Demirci U. Miniaturized lensless imaging systems for cell and microorganism visualization in point-of-care testing. Biotechnol. J. 2011, 6: 138–149. DOI: 10.1002/biot.201000427.
  21. 21. Zhu H, Isikman SO, Mudanyali O, Greenbaum A, Ozcan A. Optical imaging techniques for point-of-care diagnostics. Lab Chip 2013, 13: 51–67. DOI: 10.1039/c2lc40864c.
  22. 22. Wu J, Zheng G, Lee LM. Optical imaging techniques in microfluidics and their applications. Lab Chip 2012, 12: 3566–3575. DOI: 10.1039/c2lc40517b.
  23. 23. Heng X, Erickson D, Baugh LR, Yaqoob Z, Sternberg PW, Psaltisa D, Yang C. Optofluidic microscopy — a method for implementing a high resolution optical microscope on a chip. Lab Chip 2006, 6: 1274–1276. DOI: 10.1039/B604676B.
  24. 24. Cui X, Lee LM, Heng X, Zhong W, Sternberg PW, Psaltis D, Yang C. Lensless high-resolution on-chip optofluidic microscopes for Caenorhabditis elegans and cell imaging. Proc. Natl. Acad. Sci. U. S. A. 2008, 105: 10670–10675. DOI: 10.1073/pnas.0804612105.
  25. 25. Lange D, Storment CW, Conley CA, Kovacs GTA. A microfluidic shadow imaging system for the study of the nematode Caenorhabditis elegans in space. Sens. Actuat. B 2005, 107: 904–914. DOI: 10.1016/j.snb.2004.12.039.
  26. 26. Isikman SO, Sencan I, Mudanyali O, Bishara W, Oztoprak C, Ozcan A. Lens-free optical tomographic microscope with a large imaging volume on a chip. Lab Chip, 2010, 10: 1109–1112. DOI: 10.1073/pnas.1015638108.
  27. 27. Bishara W, Zhu H, Ozcan A. Holographic opto-fluidic microscopy. Opt. Express 2010, 18: 27499–27510. DOI: 10.1364/OE.18.027499.
  28. 28. Coskun AF, Sencan I, Su T, Ozcan A. Lensfree fluorescent on-chip imaging of transgenic Caenorhabditis elegans over an ultra-wide field-of-view. PLoS One 2011, 6: e15955. DOI: 10.1371/journal.pone.0015955.
  29. 29. Liu P, Martin RJ, Dong L. Micro-electro-fluidic grids for nematodes: A lens-less, image-sensor-less approach for on-chip tracking of nematode locomotion. Lab Chip 2013, 13: 650–661. DOI: 10.1039/c2lc41174a.
  30. 30. Chokshi TV, Bazopoulou D, Chronis N. An automated microfluidic platform for calcium imaging of chemosensory neurons in Caenorhabditis elegans. Lab Chip 2010, 10: 2758–2763. DOI: 10.1039/c004658b.
  31. 31. Rohde CB, Zeng F, Gonzalez-Rubio R, Angel M, Yanik MF. Microfluidic system for on-chip high-throughput whole-animal sorting and screening at subcellular resolution. Proc. Natl. Acad. Sci. U. S. A. 2007, 104: 13891–13895. DOI: 10.1073/pnas.0706513104.
  32. 32. Chung K, Crane MM, Lu H. Automated on-chip rapid microscopy, phenotyping and sorting of C. elegans. Nat. Methods 2008, 5: 637–643. DOI: 10.1038/nmeth.1227.
  33. 33. Yan Y, Ng LF, Ng LT, Choi KB, Gruber J, Bettiol AA, Thakor NV. A continuous-flow C. elegans sorting system with integrated optical fiber detection and laminar flow switching. Lab Chip 2014, 14: 4000–4006. DOI: 10.1039/c4lc00494a.
  34. 34. Caceres IC, Valmas N, Hilliard MA, Lu H. Laterally orienting C. elegans using geometry at microscale for high-throughput visual screens in neurodegeneration and neuronal development studies. PLoS One 2012, 7: e35037. DOI: 10.1371/journal.pone.0035037.
  35. 35. Li S, Stone HA, Murphy CT. A microfluidic device and automatic counting system for the study of C. elegans reproductive aging. Lab Chip 2015, 15: 524–531. DOI: 10.1039/c4lc01028k.
  36. 36. Hulme SE, Shevkoplyas SS, McGuigan AP, Apfeld J, Fontana W, Whitesides GM. Lifespan-on-a-chip: microfluidic chambers for performing lifelong observation of C. elegans. Lab Chip 2010, 10: 589–597. DOI: 10.1039/b919265d.
  37. 37. Krajniak J, Lu H. Long-term high-resolution imaging and culture of C. elegans in chip-gel hybrid microfluidic device for developmental studies. Lab Chip 2010, 10: 1862–1868. DOI: 10.1039/c001986k.
  38. 38. Kim N, Dempsey CM, Zoval JV, Sze J, Madou MJ. Automated microfluidic compact disc (CD) cultivation system of Caenorhabditis elegans. Sens. Actuat. B 2007, 122: 511–518. DOI: 10.1016/j.snb.2006.06.026.
  39. 39. Bringmann H. Agarose hydrogel microcompartments for imaging sleep- and wake-like behavior and nervous system development in Caenorhabditis elegans larvae. J. Neurosci. Methods 2011, 201: 78–88. DOI: 10.1016/j.jneumeth.2011.07.013.
  40. 40. Xian B, Shen J, Chen W, Sun N, Qiao N, Jiang D, Yu T, Men Y, Pang Z, Kaeberlein M, Huang Y, Han JD. WormFarm: a quantitative control and measurement device toward automated Caenorhabditis elegans aging analysis. Aging Cell 2013, 12: 398–409. DOI: 10.1111/acel.12063.
  41. 41. Jung J, Nakajima M, Tajima H, Huang Q, Fukuda T J. A microfluidic device for the continuous culture and analysis of Caenorhabditis elegans in a toxic aqueous environment. Micromech. Microeng. 2013, 23: 085008. DOI: 10.1088/0960-1317/23/8/085008.
  42. 42. Aubry G, Zhan M, Lu H. Hydrogel-droplet microfluidic platform for high-resolution imaging and sorting of early larval Caenorhabditis elegans. Lab Chip 2015, 15: 1424–1431. DOI: 10.1039/c4lc01384k.
  43. 43. Shi W, Qin J, Ye N, Lin B. Droplet-based microfluidic system for individual Caenorhabditis elegans assay. Lab Chip 2008, 8: 1432–1435. DOI: 10.1039/b808753a.
  44. 44. Clausell-Tormos J, Lieber D, Baret J, El-Harrak A, Miller OJ, Frenz L, Blouwolff J, Humphry KJ, Koester S, Duan H, Holtze C, Weitz DA, Griffiths AD, Merten CA. Droplet-based microfluidic platforms for the encapsulation and screening of mammalian cells and multicellular organisms. Chem. Biol. 2008, 15: 427–437. DOI: 10.1016/j.chembiol.2008.04.004.
  45. 45. Uppaluri S, Brangwynne CP. A size threshold governs Caenorhabditis elegans developmental progression. Proc. R. Soc. B 282: 20151283. DOI:10.1098/rspb.2015.1283.
  46. 46. Turek M, Besseling J, Bringmann H. Agarose microchambers for long-term calcium imaging of Caenorhabditis elegans. J. Vis. Exp. 2015, 100: e52742 (8 pp.), DOI: 10.3791/52742.
  47. 47. Turek M, Besseling J, Spies J, Koenig S, Bringmann H. Sleep-active neuron specification and sleep induction require FLP-11 neuropeptides to systemically induce sleep. eLife 2016, 5: e12499 (18 pp.). DOI: 10.7554/eLife.12499.
  48. 48. Cornaglia M, Mouchiroud L, Marette A, Narasimhan S, Lehnert T, Jovaisaite V, Auwerx J, Gijs MAM. An automated microfluidic platform for C. elegans embryo arraying, phenotyping, and long-term live imaging. Scientific Reports 2015, 5:10192 (13 pp.). DOI: 10.1038/srep10192.
  49. 49. Wang X, Hu R, Ge A, Hu L, Wang S, Feng X, Du W, Liu B. Highly efficient microfluidic sorting device for synchronizing developmental stages of C. elegans based on deflecting electrotaxis. Lab Chip 2015, 15: 2513–2521. DOI: 10.1039/c5lc00354g.
  50. 50. Allen PB, Sgro AE, Chao DL, Doepker BE, Edgar JS, Shen K, Chiu DT. Single-synapse ablation and long-term imaging in live C. elegans. J. Neurosci. Methods 2008, 173: 20–26. DOI: 10.1016/j.jneumeth.2008.05.007.
  51. 51. Ma H, Jiang L, Shi W, Qin J, Lin B. A programmable microvalve-based microfluidic array for characterization of neurotoxin-induced responses of individual C. elegans. Biomicrofluidics 2009, 3: 044114. DOI: 10.1063/1.3274313.
  52. 52. Chung K, Lu H. Automated high-throughput cell microsurgery on-chip. Lab Chip 2009, 9: 2764–2766. DOI: 10.1039/b910703g.
  53. 53. Yang J, Chen Z, Ching P, Shi Q, Li X. An integrated microfluidic platform for evaluating in vivo antimicrobial activity of natural compounds using a whole-animal infection model. Lab Chip 2013, 13: 3373–3382. DOI: 10.1039/c3lc50264c.
  54. 54. Wen H, Shi W, Qin J. Multiparameter evaluation of the longevity in C. elegans under stress using an integrated microfluidic device. Biomed. Microdevices 2012, 14: 721–728. DOI: 10.1007/s10544-012-9652-9.
  55. 55. Chung K, Zhan M, Srinivasan J, Sternberg PW, Gong E, Schroederd FC, Lu H. Microfluidic chamber arrays for whole-organism high-throughput chemical screening. Lab Chip 2011, 11: 3689–3697. DOI: 10.1039/c1lc20400a.
  56. 56. Zeng F, Rohde CB, Yanik MF. Sub-cellular precision on-chip small-animal immobilization, multi-photon imaging and femtosecond-laser manipulation. Lab Chip 2008, 8: 653–656. DOI: 10.1039/b804808h.
  57. 57. Guo SX, Bourgeois F, Chokshi T, Durr NJ, Hilliard MA, Chronis N, Ben-Yakar A. Femtosecond laser nanoaxotomy lab-on-a-chip for in vivo nerve regeneration studies. Nat. Methods 2008, 5: 531–533. DOI: 10.1038/NMETH.1203.
  58. 58. Yang J, Chen Z, Yang F, Wang S, Hou F. A microfluidic device for rapid screening of chemotaxis-defective Caenorhabditis elegans mutants. Biomed. Microdevices 2013, 15: 211–220. DOI: 10.1007/s10544-012-9719-7.
  59. 59. Crane MM, Chung K, Lu H. Computer-enhanced high-throughput genetic screens of C. elegans in a microfluidic system. Lab Chip 2009, 9: 38–40. DOI: 10.1039/b813730g.
  60. 60. Crane MM, Stirman JN, Ou CY, Kurshan PT, Rehg JM, Shen K, Lu H. Autonomous screening of C. elegans identifies genes implicated in synaptogenesis. Nat. Methods 2012, 9: 977–980. DOI: 10.1038/nmeth.2141.
  61. 61. Ai X, Zhuo W, Liang Q, McGrath PT, Lu H. A high-throughput device for size based separation of C. elegans developmental stages. Lab Chip 2014, 14: 1746–1752. DOI: 10.1039/c3lc51334c.
  62. 62. Solvas XC, Geier FM, Leori AM, Bundy JG, Edela JB, Mello AJ. High-throughput age synchronisation of Caenorhabditis elegans. Chem. Commun. 2011, 47: 9801–9803. DOI: 10.1039/c1cc14076k.
  63. 63. Dong L, Cornaglia M, Lehnert T, Gijs MAM. Versatile size-dependent sorting of C. elegans nematodes and embryos using a tunable microfluidic filter structure. Lab Chip 2016, 16: 574–585. DOI: 10.1039/C5LC01328C.
  64. 64. Lee H, Kim SA, Coakley S, Mugno P, Hammarlund M, Hilliard MA, Lu H. A multi-channel device for high-density targetselective stimulation and long-term monitoring of cells and subcellular features in C. elegans. Lab Chip 2014, 14: 4513–4522. DOI: 10.1039/c4lc00789a.
  65. 65. Wang Y, Wang J, Du W, Feng XJ, Liu B. Identification of the neuronal effects of ethanol on C. elegans by in vivo fluorescence imaging on a microfluidic chip. Anal. Bioanal. Chem. 2011, 399: 3475–3481. DOI: 10.1007/s00216-010-4148-z.
  66. 66. Chronis N, Zimmer M, Bargmann C. Microfluidics for in vivo imaging of neuronal and behavioral activity in Caenorhabditis elegans. Nat. Methods 2007, 4: 727–731. DOI: 10.1038/nmeth1075.
  67. 67. Chokshi TV, Ben-Yakar A, Chronis N. CO2 and compressive immobilization of C. elegans on-chip. Lab Chip 2009, 9: 151–157. DOI: 10.1039/b807345g.
  68. 68. Hulme SE, Shevkoplyas SS, Apfeld J, Fontana W, Whitesides GM. A microfabricated array of clamps for immobilizing and imaging C. elegans. Lab Chip 2007, 7: 1515–1523. DOI: 10.1039/b707861g.
  69. 69. Mondal S, Ahlawat S, Rau K, Venkataraman V, Koushika SP. Imaging in vivo neuronal transport in genetic model organisms using microfluidic devices. Traffic 2011, 12: 372–385. DOI: 10.1111/j.1600-0854.2010.01157.x.
  70. 70. Mondal S, Ahlawat S, Rau K, Venkataraman V, Koushika SP. Imaging in vivo neuronal transport in genetic model organisms using microfluidic devices. Traffic 2011, 12: 372–385. DOI: 10.1111/j.1600-0854.2010.01157.x.
  71. 71. McCormick KE, Gaertner BE, Sottile M, Phillips PC, Lockery SR. Microfluidic Devices for analysis of spatial orientation behaviors in semi-restrained Caenorhabditis elegans. PLoS One 2011, 6: e25710. DOI: 10.1371/journal.pone.0025710.
  72. 72. Gilleland CL, Rohde CB, Zeng F, Yanik MF. Microfluidic immobilization of physiologically active Caenorhabditis elegans. Nat. Protoc. 2010, 5: 1888–1902. DOI: 10.1038/nprot.2010.143.
  73. 73. Rohde CB, Gilleland C, Samara C, Norton S, Haggarty S, Yanik MF. Microfluidic in vivo screen identifies compounds enhancing neuronal regeneration. Conf. Proc. IEEE Eng. Med. Biol. Soc. 2009, 5950–5952. DOI: 10.1109/IEMBS.2009.5334771.
  74. 74. Wang J, Feng X, Du W, Liu B. Microfluidic worm-chip for in vivo analysis of neuronal activity upon dynamic chemical stimulations. Anal. Chim. Acta. 2011, 701: 23–28. DOI: 10.1016/j.aca.2011.06.007.
  75. 75. Wang J, Li Z, Xu Z, Hu L, Feng X, Chen M, Du W, Wu Z, Luo Q, Xu T, Liu B. Development of an integrated microfluidic device for evaluating of in vivo chemo-sensing of intact Caenorhabditis elegans. Sens. Actuat. B 2013, 178: 343–349. DOI: 10.1016/j.snb.2012.12.102.
  76. 76. Hu L, Wang J, Feng X, Du W, Liu B. Microfluidic device for analysis of gas-evoked neuronal sensing in C. elegans. Sens. Actuat. B: Chem 2015, 209: 109–115. DOI:10.1016/j.snb.2014.11.081.
  77. 77. Lockery SR, Hulme SE, Roberts WM, Robinson KJ, Laromaine A, Lindsay TH, Whitesides GM, Weeks JC. A microfluidic device for whole-animal drug screening using electrophysiological measures in the nematode C. elegans. Lab Chip 2012, 12: 2211–2220. DOI: 10.1039/c2lc00001f.
  78. 78. Ding X, Lin SS, Kiraly B, Yue H, Li S, Chiang I, Shi J, Benkovic SJ, Huang TJ. On-chip manipulation of single microparticles, cells, and organisms using surface acoustic waves. Proc. Natl. Acad. Sci. U.S.A. 2012, 1–5. DOI: 10.1073/pnas.1209288109.
  79. 79. Ahmed D, Ozcelik A, Bojanala N, Nama N, Upadhyay A, Chen Y, Hanna-Rose W, Huang TJ. Rotational manipulation of single cells and organisms using acoustic waves. Nat. Commun. 2016, 7: 11085 (11 pp.). DOI: 10.1038/ncomms11085.
  80. 80. Chuang H, Chen H, Chen C, Chiu W. Immobilization of the nematode Caenorhabditis elegans with addressable light-Induced heat knockdown (ALINK). Lab Chip 2013, 13: 2980–2989. DOI: 10.1039/c3lc50454a.
  81. 81. Qiu Z, Tu L, Huang L, Zhu T, Nock V, Yu E, Liu X, Wang W. An integrated platform enabling optogenetic illumination of Caenorhabditis elegans neurons and muscular force measurement in microstructured environments. Biomicrofluidics 2015, 9: 014123. DOI: 10.1063/1.4908595.
  82. 82. Larsch J, Ventimiglia D, Bargmann CI, Albrecht DR. High-throughput imaging of neuronal activity in Caenorhabditis elegans. PNAS 2013, 110: e4266 (7 pp.). DOI: 10.1073/pnas.1318325110.
  83. 83. Lockery SR, Lawton KJ, Doll JC, Faumont S, Coulthard SM, Thiele TR, Chronis N, McCormick KE, Goodman MB, Pruitt BL. Artificial dirt: microfluidic substrates for nematode neurobiology and behavior. J. Neurophysiol. 2008, 99: 3136–3143. DOI: 10.1152/jn.91327.2007.
  84. 84. Park S, Hwang H, Nam S, Martinez F, Austin RH, Ryu WS. Enhanced Caenorhabditis elegans locomotion in a structured microfluidic environment. PLoS One 2008, 3: e2550. DOI: 10.1371/journal.pone.0002550.
  85. 85. Johari S, Nock V, Alkaisi MM, Wang W. On-chip analysis of C. elegans muscular forces and locomotion patterns in microstructured environments. Lab Chip 2013, 13: 1699–1707. DOI: 10.1039/c3lc41403e.
  86. 86. Han B, Kim D, Ko UH, Shin JH. A sorting strategy for C. elegans based on size-dependent motility and electrotaxis in a micro-structured channel. Lab Chip 2012, 12: 4128–4134. DOI: 10.1039/C2LC40209B.
  87. 87. Parashar A, Lycke R, Carr JA, Pandey S. Amplitude-modulated sinusoidal microchannels for observing adaptability in C. elegans locomotion. Biomicrofluidics 2011, 5: 024112–024119. DOI: 10.1063/1.3604391.
  88. 88. Doll JC, Harjee N, Klejwa N, Kwon R, Coulthard SM, Petzold B, Goodman MB, Pruitt BL. SU-8 force sensing pillar arrays for biological measurements. Lab Chip 2009, 9: 1449–1454. DOI: 10.1039/b818622g.
  89. 89. Oliver CR, Gourgou E, Bazopoulou D, Chronis N, Hart AJ. On-demand isolation and manipulation of C. elegans by In Vitro maskless photopatterning. PLoS One 2016, 11: e0145935 (16 pp.). DOI:10.1371/journal.pone.0145935.
  90. 90. Kuo WJ, Sie YS, Chuang HS. Characterizations of kinetic power and propulsion of the nematode Caenorhabditis elegans based on a micro-particle image velocimetry system. Biomicrofluidics 2014, 8: 024116. DOI: 10.1063/1.4872061.
  91. 91. Wang X, Tang L, Xia Y, Hu L, Feng X, Du W, Liu B. Stress response of Caenorhabditis elegans induced by space crowding in a micro-column array chip. Integr. Biol. 2013, 5: 728–737. DOI: 10.1039/C3IB20289E.
  92. 92. Chuang H, Raizen D, Lamb A, Dabbish N, Bau H. Dielectrophoresis of Caenorhabditis elegans. Lab Chip 2011, 11: 599–604. DOI: 10.1039/C0LC00532K.
  93. 93. Carr JA, Parashar A, Gibson R, Robertson AP, Martin RJ, Pandey S. A microfluidic platform for high-sensitivity, real-time drug screening on C. elegans and parasitic nematodes. Lab Chip 2011, 11: 2385–2396. DOI: 10.1039/C1LC20170K.
  94. 94. Rezai P, Salam S, Selvaganapathy PR, Gupta BP. Effect of pulse direct current signals on electrotactic movement of nematodes Caenorhabditis elegans and Caenorhabditis briggsae. Biomicrofluidics 2011, 5: 044116. DOI: 10.1063/1.3665224.
  95. 95. Rezai P, Siddiqui A, Salam S, Selvaganapathy PR, Gupta BP. Behavior of Caenorhabditis elegans in alternating electric field and its application to their localization and control. Appl. Phys. Lett. 2010, 96: 153702 (3 pp.). DOI: 10.1063/1.3383223.
  96. 96. Rezai P, Siddiqui A, Selvaganapathy PR, Gupta BP. Electrical sorting of Caenorhabditis elegans. Lab Chip 2012, 12: 1831–1840. DOI: 10.1039/c2lc20967e.
  97. 97. Rezai P, Siddiqui A, Selvaganapathy PR, Gupta BP. Electrotaxis of Caenorhabditis elegans in a microfluidic environment. Lab Chip 2010, 10: 220–226. DOI: 10.1039/B917486A.
  98. 98. Tong J, Rezai P, Salam S, Selvaganapathy PR, Gupta BP. Microfluidic-based electrotaxis for on-demand quantitative analysis of Caenorhabditis elegans' locomotion. J. Vis. Exp. 2013, 75: e50226. DOI: 10.3791/50226.
  99. 99. Maniere X, Lebois F, Matic I, Ladoux B, Meglio JD, Hersen P. Running worms: C. elegans self-sorting by electrotaxis. PLoS One 2011, 6: e16637. DOI: 10.1371/journal.pone.0016637.
  100. 100. Jung J, Nakajima M, Takeuchi M, Najdovski Z, Huang Q, Fukuda T. Microfluidic device to measure the speed of C. elegans using the resistance change of the flexible electrode. Micromachines 2016, 7: 50 (12 pp.). DOI: 10.3390/mi7030050.
  101. 101. Xu J, Wu D, Ip JY, Midorikawaa K, Sugioka K. Vertical sidewall electrodes monolithically integrated into 3D glass microfluidic chips using water-assisted femtosecond-laser fabrication for in situ control of electrotaxis. RSC Adv. 2015, 5: 24072–24080. DOI: 10.1039/c5ra00256g.
  102. 102. Zhang B, Li Y, He Q, Qin J, Yu Y, Li X, Zhang L, Yao M, Liu J, Chen Z. Microfluidic platform integrated with worm-counting setup for assessing manganese toxicity. Biomicrofluidics 2014, 8: 054110. DOI: 10.1063/1.4896663.
  103. 103. Zhang Y, Lu H, Bargmann CI. Pathogenic bacteria induce aversive olfactory learning in Caenorhabditis elegans. Nature 2005, 438: 179–184. DOI: 10.1038/nature04216.
  104. 104. Chalasani SH, Chronis N, Tsunozaki M, Gray JM, Ramot D, Goodman MB, Bargmann CI. Dissecting a circuit for olfactory behaviour in Caenorhabditis elegans. Nature 2007, 450: 63–70. DOI: 10.1038/nature06292.
  105. 105. Zhang Y, Lu H, Bargmann CI. Pathogenic bacteria induce aversive olfactory learning in Caenorhabditis elegans. Nature 2005, 438: 179–184. DOI: 10.1038/nature04216.
  106. 106. Albrecht DR, Bargmann CI. High-content behavioral analysis of Caenorhabditis elegans in precise spatiotemporal chemical environments. Nat. Methods 2011, 8: 599–605. DOI: 10.1038/nmeth.1630.
  107. 107. Luo L, Gabel CV, Ha H, Zhang Y, Samuel ADT. Olfactory behavior of swimming C. elegans analyzed by measuring motile responses to temporal variations of odorants. J. Neurophysiol. 2008, 99: 2617–2625. DOI: 10.1152/jn.00053.2008.
  108. 108. Stirman JN, Brauner M, Gottschalk A, Lu H. High-throughput study of synaptic transmission at the neuromuscular junction enabled by optogenetics and microfluidics. J. Neurosci. Methods 2010, 191: 90–93. DOI: 10.1016/j.jneumeth.2010.05.019.
  109. 109. Song P1, Zhang W, Sobolevski A, Bernard K, Hekimi S, Liu X. A microfluidic device for efficient chemical testing using Caenorhabditis elegans. Biomed. Microdevices. 2015, 17: 38–48. DOI: 10.1007/s10544-015-9939-8.
  110. 110. Wen Q, Po MD, Hulme E, Chen S, Liu X, Kwok SW, Gershow M, Leifer AM, Butler V, C Fang-Yen, Kawano T, Schafer WR, Whitesides G, Wyart M, Chklovskii DB, Zhen M, Samuel ADT. Proprioceptive coupling within motor neurons drives C. elegans forward locomotion. Neuron 2012, 76: 750–761. DOI: 10.1016/j.neuron.2012.08.039.
  111. 111. Hwang H, Kim E, Kim SH, Park S. A sensitive C. elegans chemotaxis assay using microfluidic device generating a linear gradient of chemoeffectors. Bull. Korean Chem. Soc. 2015, 36: 1096–1099. DOI: 10.1002/bkcs.10201.
  112. 112. Salam S, Ansari A, Amon S, Rezai P, Selvaganapathy PR, Mishra RK, Gupta BP. A microfluidic phenotype analysis system reveals function of sensory and dopaminergic neuron signaling in C. elegans electrotactic swimming behavior. Worm 2013, 2: e24558. DOI: 10.4161/worm.24558.
  113. 113. Chalasani SH, Chronis N, Tsunozaki M, Gray JM, Ramot D, Goodman MB, Bargmann CI. Dissecting a circuit for olfactory behaviour in Caenorhabditis elegans. Nature 2007, 450: 63–70. DOI: 10.1038/nature06292.
  114. 114. Saldanha JN, Parashar A, Pandey S, Powell-Coffman JA. Multiparameter behavioral analyses provide insights to mechanisms of cyanide resistance in Caenorhabditis elegans. Toxicol. Sci. 2013, 135: 156–168. DOI: 10.1093/toxsci/kft138.
  115. 115. Gray JM, Karow DS, Lu H, Chang AJ, Chang JS, Ellis RE, Marletta MA, Bargmann CI. Oxygen sensation and social feeding mediated by a C. elegans guanylate cyclase homologue. Nature 2004, 430: 317–322. DOI: 10.1038/nature02714.
  116. 116. Zimmer M, Gray JM, Pokala N, Chang AJ, Karow DS, Marletta MA, Hudson ML, Morton DB, Chronis N, Bargmann CI. Neurons detect increases and decreases in oxygen levels using distinct guanylate cyclases. Neuron 2009, 61: 865–879. DOI: 10.1016/j.neuron.2009.02.013.
  117. 117. Santos SI, Mathew M, Loza-Alvarez P. Real time imaging of femtosecond laser induced nano-neurosurgery dynamics in C. elegans. Opt. Express 2010, 18: 364–377. DOI: 10.1364/OE.18.000364.
  118. 118. Buonanno M, Garty G, Grad M, Gendrel M, Hobert O, Brenner DJ. Microbeam irradiation of C. elegans nematode in microfluidic channels. Radiat. Environ. Biophys. 2013, 52: 531–537. DOI 10.1007/s00411-013-0485-6.
  119. 119. Cornaglia M, Krishnamani G, Mouchiroud L, Sorrentino V, Lehnert T, Auwerx J, Gijs MAM. Automated longitudinal monitoring of in vivo protein aggregation in neurodegenerative disease C. elegans models. Mol. Neurodegener 2016, 11: 1–13. DOI: 10.1186/s13024-016-0083-6.
  120. 120. Hu C, O’Connor V, Holden-Dye L, Morgan H. Conf. Proc. MicroTAS 2013, pp. 1441–1443. DOI: 10.13140/2.1.1565.0244.
  121. 121. Hu C, Dillon J, Kearn J, Murray C, O'Connor V, Holden-Dye L, Morgan H. NeuroChip: a microfluidic electrophysiological device for genetic and chemical biology screening of Caenorhabditis elegans adult and larvae. PLoS One 2013, 8: e64297. DOI: 10.1371/journal.pone.0064297.
  122. 122. Gokce SK, Guo SX, Ghorashian N, Everett WN, Jarrell T, et al. (2014) A fully automated microfluidic femtosecond laser axotomy platform for nerve regeneration studies in C. elegans. PLoS One 2014, 9: e113917 (28 pp.). DOI: 10.1371/journal.pone.0113917.
  123. 123. Zhao X, Xu F, Tang L, Du W, Feng X, Liu BF. Microfluidic chip-based C. elegans microinjection system for investigating cell-cell communication in vivo. Biosens. Bioelectron. 2013, 50: 28–34. DOI: 10.1016/j.bios.2013.06.024.
  124. 124. Gilleland CL, Rohde CB, Zeng F, Yanik MF. Microfluidic immobilization of physiologically active Caenorhabditis elegans. Nat. Protoc. 2010, 5: 1888–1902. DOI: 10.1038/nprot.2010.143.
  125. 125. Ben-Yakar A, Bourgeois F. Ultrafast laser nanosurgery in microfluidics for genome-wide screenings. Curr. Opin. Biotechnol. 2009, 20: 100–105. DOI: 10.1016/j.copbio.2009.01.008.
  126. 126. Samara C, Rohde CB, Gilleland CL, Norton S, Haggarty SJ, Yanik MF. Large-scale in vivo femtosecond laser neurosurgery screen reveals small-molecule enhancer of regeneration. Proc. Natl. Acad. Sci. U. S. A. 2010, 107: 18342–18347. DOI: 10.1073/pnas.1005372107.
  127. 127. Fang-Yen C, Gabel CV, Samuel ADT, Bargmann CI, Avery LS. Laser microsurgery in Ceanorhabditis elegans. Methods Cell Biol. 2012, 107: 177–206. DOI: 10.1016/B978-0-12-394620-1.00006-0.
  128. 128. Song P, Dong X, Liu X. A microfluidic device for automated, high-speed microinjection of Caenorhabditis elegans. Biomicrofluidics 2016, 10: 011912 (12 pp.). DOI: 10.1063/1.4941984.
  129. 129. Callaghan P. Principles of nuclear magnetic resonance microscopy, UK: Oxford Science Publications, Corby; 1991.
  130. 130. Trautwein C, Spengler N, MacKinnon N, Wallrabe U, Eimer S, Korvink JG. Quantitative metabolomics of the nematode C. elegans in a microfluidic chip device with NMR spectroscopy. Conf. Proc. Euromar 2014, 29 June– 3 July, Zurich, Switzerland.
  131. 131. Spengler N, Moazenzadeh A, Meier RC, Badilita V, Korvink JG, Wallrabe U. Micro-fabricated Helmholtz coil featuring disposable microfluidic sample inserts for applications in nuclear magnetic resonance. J. Micromech. Microeng. 2014, 24: 034004 (10 pp.). DOI: 10.1088/0960-1317/24/3/034004.
  132. 132. Gruschke OG, Baxan N, Clad L, Kratt K, von Elverfeldt D, Peter A, Hennig J, Badilita V, Wallrabe U, Korvink JG. Lab on a chip phased-array MR multi-platform analysis system. Lab Chip 2012, 12: 495–502. DOI: 10.1039/C2LC20585H.
  133. 133. Aguayo JB, Blackband SJ, Schoeniger J, Mattingly M, Hintermann M. Nuclear magnetic resonance imaging of a single cell. Nature 1986, 322: 190–191. DOI: 10.1038/322190a0.
  134. 134. Kalfe A, Telfah A, Lambert J, Hergenröder R. Looking into living cell systems: planar waveguide microfluidic NMR detector for in vitro metabolomics of tumor spheroids. Anal. Chem. 2015, 87: 7402–7410. DOI: 10.1021/acs.analchem.5b01603.
  135. 135. Meier RC, Höfflin J, Badilita V, Wallrabe U, Korvink JG. Microfluidic integration of wirebonded microcoils for on-chip applications in nuclear magnetic resonance. J. Micromech. Microeng. 2014, 24: 045021 (12 pp.). DOI: 10.1088/0960-1317/24/4/045021.
  136. 136. Wong A, Li X, Molin L, Solari F, Elena-Herrmann B, Sakellariou D. μhigh resolution-magic-angle spinning NMR spectroscopy for metabolic phenotyping of Caenorhabditis elegans. Anal. Chem. 2014, 86: 6064–6070. DOI: 10.1021/ac501208z.

Written By

Natalia A. Bakhtina, Neil MacKinnon and Jan G. Korvink

Submitted: 22 May 2015 Reviewed: 18 May 2016 Published: 23 November 2016