Open access peer-reviewed chapter

Microfluidics and Nanofluidics: Science, Fabrication Technology (From Cleanrooms to 3D Printing) and Their Application to Chemical Analysis by Battery-Operated Microplasmas-On-Chips

Written By

Vassili Karanassios

Submitted: 17 October 2017 Reviewed: 25 January 2018 Published: 22 August 2018

DOI: 10.5772/intechopen.74426

From the Edited Volume

Microfluidics and Nanofluidics

Edited by Mohsen Sheikholeslami Kandelousi

Chapter metrics overview

2,242 Chapter Downloads

View Full Metrics

Abstract

The science and phenomena that become important when fluid-flow is confined in microfluidic channels are initially discussed. Then, technologies for channel fabrication (ranging from photolithography and chemical etching, to imprinting, and to 3D-printing) are reviewed. The reference list is extensive and (within each topic) it is arranged chronologically. Examples (with emphasis on those from the authors’ laboratory) are highlighted. Among them, they involve plasma miniaturization via microplasma formation inside micro-fluidic (and in some cases millifluidic) channels fabricated on 2D and 3D-chips. Questions addressed include: How small plasmas can be made? What defines their fundamental size-limit? How small analytical plasmas should be made? And what is their ignition voltage? The discussion then continues with the science, technology and applications of nanofluidics. The conclusions include predictions on potential future development of portable instruments employing either micro or nanofluidic channels. Such portable (or mobile) instruments are expected to be controlled by a smartphone; to have (some) energy autonomy; to employ Artificial Intelligence and Deep Learning, and to have wireless connectivity for their inclusion in the Internet-of-Things (IoT). In essence, those that can be used for chemical analysis in the field for “bringing part of the lab to the sample” types of applications.

Keywords

  • microfluidics
  • nanofluidics
  • wet chemical etching
  • embossing
  • polymeric substrates
  • 3D printing
  • rapid prototyping
  • microplasmas
  • portability
  • postage stamp sized 2D-chips
  • 3D-chips
  • Lab‐on‐a‐chip
  • MEMS
  • NEMS

1. Science, technology and advantages of microfluidics

As is the case in many fields of scientific research, the field of microfluidics has three main components: a science, a technology and an applications component.

For microfluidics, a common thread between all of these components is that they are micro-sized, so size will be briefly discussed first. The dimensions shown in Figure 1 are approximate because size of naturally-occurring objects (and of some manufactured-things) varies, for example the diameter of a human hair is between 50 and 100 μm; the diameter of the tip of a rollerball pen is between fine, medium and bold (e.g., between 0.5 and 0.7 mm); and of a 1 cent coin with its diameter varying slightly depending on the jurisdiction the penny was minted (typically around 20 mm or somewhat more).

Figure 1.

Examples of an approximate scale of things. The boundaries between micro and nanofluidics and between micro and millifluidics are fuzzy. In many cases, the strict definition adopted by the National Science Foundation (NSF) of the US for nano as anything with one critical dimension ≤100 nm is not strictly adhered to, thus there is a gap between 100 nm and 1 μm. Similar arguments apply to the NSF definition for micro (defined as one with a critical dimension between 1 and 100 μm). In many cases, the micro-scale is arbitrarily widened to ~1 mm and sometimes slightly more. The term millifluidics has recently been used for channels (or structures) with one critical dimension of a few mm.

1.1. Microfluidics as a science

Microfluidics has been defined [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17] as the study of the behavior of fluids (or whatever is in them, e.g., colloids, discrete nanoparticles or individual cells), in micro or in sub-millimeter channels or around microstructures. Although microchannels can be relatively long (e.g., several 10’s of mm), they are still called microchannels as long as one critical dimension (e.g., channel-width or channel-depth or tube radius) is in the micro scale. Microfluidic channels can be used for example to confine or to guide or to mix or to manipulate fluids.

  • The science of scaling as applied to microfluidics: a number of physical properties of fluids change as size gets smaller [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47], to quote “smaller brings new capability” [31]. These changes are often non-linear and have been discussed in books [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17] and in journal papers [18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29]. A non-exhaustive list of size-dependent phenomena and effects is outlined below.

  • The length-cube relationship: the geometrical scale of length varies linearly but volume varies as length-to-the-power-of-three. As a consequence, volume changes rapidly as length decreases. Typical volumes of fluids in microfluidic channels range between nano-liter (nL) and femtoliter (fL). At the μm-scale, some properties of fluids change (as compared to a mesoscale, arbitrarily defined as the intermediate scale between the micro-scale and the macro-scale). Example properties that dominate at a micro-scale and that are different than those observed at the meso and macroscales include dominance of laminar-flow, diffusion-dominated mixing and capillary action. To highlight one such effect, a counter-intuitive example (from an every-day scale point of view) involves two parallel-flowing fluid-streams that come into contact in a microchannel. Since there are no eddy currents or turbulence (due to laminar flow), the only mixing that occurs is a result of slow-occurring diffusion at the interface between the two fluid-flows. Since there is no bulk mixing, mixture-separations in microchannels are faster and have shorter separation times.

  • The square-cube law: states that volume increases faster than surface area. In microfluidics, fluid-flows in microchannels are influenced or controlled or are a function of surface area (e.g., surface tension) while others (e.g., weight) are a function of volume. Typically, in microfluidics there are no gravity effects but dominance of surface tension and of interface effects.

  • Examples of other phenomena influenced by size and expressed by dimensionless numbers: these include laminar flow expressed by the Reynolds number; surface tension expressed by the Bond number; transient thermal effects expressed by the Fourier number; viscous heating by the Brinkman number; and fluid compressibility by the Mach number.

  • As a result of channel-size, microfluidics enables one to probe individually whatever it is in a fluid constrained in a microchannel (e.g., a single cell), thus providing additional avenues for scientific inquiry and discovery (important especially in the bio-analytical sciences).

Overall, the relevant literature [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47] describes efforts at exploring and understanding the Physics of flow-related phenomena. Developments enabled by microfluidics will be highlighted in this chapter, with emphasis on ionized gases (e.g., Paschen’s law for electrical gas breakdown; plasma sheaths and the Debye length) as applied to microplasmas formed inside fluidic channels.

1.2. Microfluidics as a technology

Microfluidics refers to a variety of approaches that enable exploitation of the phenomena mentioned above by fabricating microfluidic channels on a variety of substrates. For instance, on crystalline Silicon (of c-Si) wafers, on amorphous glass or on polymeric substrates. Due to the advantages of confining flow in microfluidic channels, several fabrication technologies have been developed and tested and will be briefly reviewed. These technologies are often collectively called micro Total Analysis Systems (μTAS) or Lab-on-a-Chip (LoC) or Micro Electro Mechanical Systems (MEMS). Microfluidics or whatever acronym is used to describe it, has attracted significant attention in books [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17] and in journals [18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29]. While in the topic of publications, older references have been purposely included in this chapter followed by some recent publications. Where possible, the citations in the reference list have been grouped either according to fabrication technology or according to the type of substrate used (e.g., c-Si, amorphous, polymeric) or according to application. Within each technology, the reference list has been sorted out chronologically to help interested readers follow the origin and evolution of ideas and technologies. Despite of the relatively large number of references included, this is not a comprehensive review. The reference list simply offers starting points. Getting back to the main theme, the question still remains: why does microfluidics continue to receive increased attention? What are the advantages of using microfluidics, especially for chemical analysis applications?

1.3. Advantages and selected applications of microfluidics

The science and technology mentioned above are widely exploited and applied to give microfluidics a host of advantages. A brief list includes use of small volumes of sample and reagents (thus reducing cost per analysis and minimizing waste disposal); rapid sample processing; potential for automation (thus reducing cost); reduced risk of contamination; short analysis time (e.g., by increasing speed of separations); small footprint and light-weight thus enabling development of future portable microfluidic-based, portable micro-instruments that can be employed on-site or for personal use or for personal dosimetry; potential for massive parallelism (for high sample throughput); and overall, lower ownership and operating costs (vis-à-vis conventional, lab-sized systems). Application areas (to name but a few), include analytical chemistry, synthetic chemistry (including nanomaterials synthesis), microbiology, biotechnology, point-of-care diagnostics, drug delivery, immunoassays and medicine, health-monitoring and health-diagnostics, agriculture, food safety and environmental monitoring [30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47].

Advertisement

2. Technology for fabrication of microfluidic channels

2.1. Fabrication using either crystalline Si (c-Si) or other substrates

Microchannel fabrication technology has been borrowed from the semiconductor industry. Initially, bulk micromachining [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 48, 49, 50, 51] was employed on crystalline Si (c-Si) substrates and on amorphous glass. To use it, a photolithographically patterned wafer was dipped into a chemical etching solution to etch-away (or subtract) material from the substrate, thus forming microchannels of desired geometry. This method is often referred to as wet chemical etching [48, 49, 50, 51]. Inadequate control of channel depth (resulting unevenly etched channels) due to spatial etch-rate variations and to pyramid formation when crystalline-Si (c-Si) substrates and deep microchannels were etched are two key disadvantages. In contrast, surface micromachining [52, 53, 54] involves repetitive patterning, thin layer deposition and selective etching of sacrificial layers. The challenge here stems from the many photolithography steps involved and from the precautions required so that previously deposited layers are not damaged.

We used (as far back as the 1990’s) cleanroom-based photolithography, bulk micromachining and wet chemical etching [48, 49, 50, 51] to fabricate shallow-depth microchannels (with relatively low width-to-depth aspect ratio). This approach is often referred to as 2D sculpting of Manhattan-like structures and it offers a planar, 2D- rather than a 3D-perspective. Some examples will be briefly discussed later.

For completeness, other methods of microchannel fabrication on inorganic substrates (either crystalline or amorphous) have been described. A short list includes laser machining [55, 56, 57, 58]; lithographie galvanoformung adformung (LIGA) or lithography electroplating molding [59, 60, 61] which is well suited for fabrication of high aspect ratio channels; deep reactive ion etching (DRIE) [62, 63, 64, 65] often used for fabrication of microchannels with a high aspect ratio; and, SU-8 (an epoxy-based negative photoresist) and its variants such as SU-8 series 2000 and SU-8 Series 3000) [66, 67, 68].

Technologies involving polymeric substrates [69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81, 82, 83] include replication via imprinting [69, 70, 71, 72, 73] or embossing [74, 75, 76]. Polymeric substrates are selected due to their bio-compatibility or to reduce cost of ownership. Examples will be shown later. The terms disposable or recyclable microfluidic devices is often used for microfluidic channels on polymeric substrates. Soft lithography [77, 78, 79, 80, 81, 82, 83] (defined as a collection of fabrication techniques for replication of microchannels) is a technology that does not require access to a clean room. It is called soft because it uses soft and flexible (primarily) elastomeric materials such as poly di methyl siloxane (PDMS) and often cyclic olefin copolymer (COC).

There are other techniques that are rather difficult to classify either according to fabrication technology or according substrate. Despite of being brief, the list includes droplet microfluidics [84, 85, 86, 87, 88, 89], in which discrete droplets or small volumes of immiscible liquids are guided through microchannels. In the early literature, this approach was often called digital microfluidics. As it is known now, digital microfluidics [90, 91, 92, 93, 94, 95] is an outgrowth of electrowetting [90, 92] and it involves use of discrete droplets on arrays of electrodes, with individual droplets manipulated by electrical means. The list also includes centrifugal microfluidics [96, 97, 98, 99, 100, 101], a technique that enables micro-flow manipulation by using rotational forces (e.g., Coriolis) obtained by spinning a CD on top of which there are microfluidic channels. This technique is often called “lab on a CD”. It also includes paper microfluidics [102, 103, 104, 105, 106, 107, 108], a technique that uses paper for development of microfluidic approaches intended for use in resource limited situations (e.g., remote geographical areas or resource-limited locations).

Rapid prototyping via 3D-printing [109, 110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120, 121, 122] involves both a technology (e.g., a 3D printer) and a materials platform (e.g., a polymer) for formation (primarily) of mill-sized fluidic (and recently) micro-sized channels [115, 117, 120]. An example of 3D printing will be discussed later in this chapter.

2.2. Fabrication technology examples

To highlight substrate-dependence of fabrication, the fabrication steps required for microchannels on c-Si and on amorphous glass or quartz substrates are compared and contrasted in Figure 2. It should be noted that depending on crystallographic orientation of the substrate and of the chemical cocktail used in the etching solution, isotropic or anisotropic etching may be obtained [48, 49, 50, 51].

Example 1: Planar 2D-chips and wet chemical etching for fabrication of microchannels on crystalline and amorphous substrates (Figure 2).

Figure 2.

Simplified steps used for fabrication of microchannels on a) a c-Si wafer as a substrate and on b), a wafer made from an amorphous material (abbreviated as a-wafer-above, such as glass).

For completeness, an example of wet chemically etched microchannels on glass is shown in Figure 3.

Figure 3.

(a) Part of a 14.5 mm by 25.6 mm chip of an etched microfluidic channel on corning 7059 glass with the photoresist removed and (for clarity) without a cover plate. Also omitted are pipette-tips used as sample reservoirs that are attached to the sample well. A coin was included for size. (b) Part of a Mylar mask used for photo-lithography. (c) Part of a washed meandering microchannel shown under 10x magnification and (d) shown under 60-fold magnification. (e) an unwashed microchannel immediately after etching showing etching by-products inside the microchannels, thus requiring their removal. (f) a sample-well and a washed microchannel showing the quality of etching, in particular for the round sample-well. For (d), (e) and (f) the photoresist was not removed to provide contrast for the photographs.

The quality of the etched microchannels depended on the composition of the etching solution and on the geometric-primitives that were used to define the channels. To enclose the microchannel of Figure 3, a cover plate was used (but is not shown in Figure 3). Depending on the required optical transparency, a UV-transparent quartz cover plate was employed for most of the work described here. Furthermore, depending on the substrate (e.g., c-Si or glass), a variety of bonding methods can be employed [2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17].

Despite of the ability to fabricate low aspect ratio microchannels, wet chemical etching has shortcomings arising from costs, from limited access by many to photolithography and to cleanrooms, and from time-delays between mask-design (Figure 3b) and receipt of finished prototype (e.g., Figure 3a). At present, access to cleanrooms is not required because microfluidic chips can now be ordered from specialized foundries. In spite of this, there are still costs and time-delays involved. There is another limitation if microchannels are to be used with biological samples, because many biosamples adhere to substrates. Thus, functionalized surfaces or microfluidic channels on polymeric substrates are preferred.

Example 2. Imprinting microchannels on planar polymeric 2D-chips. 2D-microchannel fabrication on polymeric substrates is one way of overcoming some of the limitations mentioned above. But polymers may contain additives, fillers or plasticizers that may contaminate the samples, and they may display auto-fluorescence. As for fabrication (Figure 4), it may be achieved by using Si-stamp imprinting (Figure 4) or by imprinting (by pressing) a wire on a substrate [69] (Figure 5). In the example shown in Figure 4, a c-Si stamp (or master or hard mold) was developed and was employed for replication by imprinting.

Example 3: 3D-printed, milli-sized fluidic channels on polymeric materials for hybrid 3D chips. 3D printing technology [109, 110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120, 121, 122] using polymeric materials is receiving attention for rapid prototyping [109] including fabrication of mm channels (often called millifluidics) and more recently for sub-mm channels (using specialized printers) [120, 121]. We used 3D-printing due to reduced fabrication and ownership costs and due to quick turn-around times (often from concept to prototype in hours). A simple, hybrid, 3D-printed 3D-chip containing a millifluidic channel is shown in Figure 6. The word hybrid was used because the two needle electrodes and the quartz cover plate were not 3D-printed.

Figure 4.

(a) Mask; (b) mask on c-Si chip, coin has been added for size; (c) chemically etched c-Si chip (serving as a stamp), the meandering pattern is protruding from the surface of the chip; (d) imprint generated by pressing the stamp and the polymeric chip together by placing them in a hydraulic press and by applying pressure at room temperature; (e) imprinted sample-well on a polymer chip shown under magnification; and (f), part of an imprinted meandering channel shown under magnification. For (d) and (e) and (f) different polymeric materials were used.

Figure 5.

(a) Imprinted channel on a polymeric chip (60x magnification), (b) sample-well (60x magnification) and (c) Venturi micropump with no moving parts and no electrical power requirements fabricated by imprinting (coin included for size) [73].

Figure 6.

Sugar cube-sized, 3D-printed hybrid-chip with a millifluidic channel to be fitted with a quartz cover plate (selected for UV transparency). A sample introduction system is also shown and it has been included to provide an overall size for this “critical” component of a potential future micro-instrument. An actual sugar-cube (~1 cm by ~1 cm) has been included for scale comparisons.

In my laboratory, some of the fabrication technologies discussed thus far have been used to constrain plasmas in microfluidic or in millifluidic channels. But why plasmas and why microplasmas?

Advertisement

3. Why plasmas?

There are four states of matter: gases, liquids, solids and plasmas [123, 124, 125, 126, 127, 128, 129, 130, 131]. To generalize, atmospheric pressure plasmas are ionized gases that are either hot or cold (about room temperature or somewhat above it). Plasmas occur in nature, for example those found in inter-stellar space, in the ionosphere, in auroras and in lightening. There are also artificially-generated plasmas that are being used in many every-day-life applications. Neon signs and fluorescent lights in which low-pressure plasmas are formed either in Neon (Ne) gas or in Argon (Ar) gas) are two such examples. Other examples include plasmas employed for device fabrication by the semiconductor industry or for materials synthesis in nanoscience and nanotechnology [129, 130, 131]. It has been estimated that over 50% of whatever goes inside any electronic device (e.g., a tablet, a smartphone, TV) is fabricated using a low-pressure plasma.

Conventional-scale (or lab-scale) atmospheric pressure plasmas are widely used in chemical analysis, primarily in the form of atmospheric-pressure, 6000–10,000 K hot Inductively Coupled Plasmas or ICPs [132]. Due to their size and weight (e.g., in the few 100’s of pounds), their gas consumption (e.g., ~20 L/min), their power usage (e.g., 1–2 kW) and their need for cooling, ICPs are primarily used in a lab.

3.1. Some fundamental aspects of plasma science

A plasma is an ionized gas [123, 124, 125, 126, 127, 128, 129, 130, 131]. The term plasma was coined by Langmuir in the 1920’s and it is derived from the ancient Greek word πλάσμα (plasma), freely translated to something “moldable”. A plasma consists of ions (with ion number density ni) and electrons (with an electron number density ne), and on the average it is quasi-neutral, and for singly ionized gases ne≈ni. Thus, a prerequisite for plasma formation is ionization. Singly-charge ionization (in the form of ion-electron pair formation) is done by detaching an electron from a neutral gaseous atom or molecule. Although there are other ways of detaching an electron (e.g., thermally), one way doing it is by placing a gas between two electrodes and by applying an electric field with a sufficiently field-strength to ionize the gas (Figure 7), thus forming an electrical gas discharge. Because neutral gaseous atoms or molecules (ordinarily insulators) become ion-electron pairs, they also become (partial) conductors. Partial because to an approximation, conductivity depends on the degree of ionization (this is important for weakly ionized plasmas).

Figure 7.

Ideal plasma formed in a gas-tight and pressure-controlled enclosure. The plasma is formed between two conducting plates or electrodes positioned at a distance (or gap) d from each other. For dc operation, pertinent literature should be consulted [124].

To obtain electrical gas breakdown, the dielectric strength of the gas must be exceeded. The dielectric strength is the maximum electric field-strength (in V/m) an insulating gas can endure without breaking down into ions and electrons. If there is a sufficiently large field-strength, breakdown of the dielectric strength will cause formation of (typically) a low-current spark (i.e., a momentary electrical discharge, an example is electrostatic discharge from static electricity), or formation of a continuous electric-arc requiring continuous application of an electric field from an external power supply (Figure 7) capable of providing high-current (often in the Amp range). Arcs find applicability in welding of metals.

Conditions for sustaining continuous plasma operation: Following gas breakdown, there must be continuous application of external power to sustain a plasma. Other criteria include an electrode distance d that must be > > λD and that neλ3D must be > > 1 (this is easy to satisfy for the plasmas of interest to this work), where λD is the Debye length [133, 134, 135, 136, 137]. These will be briefly discussed later in this section.

For microplasmas formed inside fluidic microchannels, in addition to gas breakdown and to continuous application of power, a microplasma must be formed in a constrained microchannel.

3.2. Scaling of lab-size, ambient-pressure plasmas to microplasmas

Arbitrarily defined, microplasmas are those with one critical dimension in the micro-meter (μm) or in the sub-milli-meter regime [138, 139]. The words “critical dimension” (i.e., one dimension such as channel depth or width or radius) are important here: an atmospheric pressure microplasma in a microfluidic channel can range in length from μm to a 10’s of mm, as long as its critical dimension fits the definition above. But as the critical dimension is reduced to sub-mm and depending on operating conditions, atmospheric pressure plasmas transition from thermal and 10,000°C hot (e.g., lab-scale ICP [132]) to non-thermal and cold [133, 134, 135, 136, 137, 138, 139] (e.g., microplasmas). They also transition from equilibrium to non-equilibrium (to an approximation, to those with gas temperature Tg << Te (electron T). There are scientific implications due to these transitions (e.g., for nanomaterials synthesis) and for excitation mechanisms (e.g., for chemical analysis). In terms of technology-implications, cold plasmas enable use of inexpensive polymeric substrates that do not melt because microplasmas are cold and they do not require cooling; and they allow use of inexpensive 3D printing technology for fabrication.

Why miniaturize atmospheric-pressure plasmas? Operation at (or near) atmospheric-pressure is preferred because it obviates the need for heavy-weight and power-consuming vacuum pumps. By reducing weight and power consumption, atmospheric-pressure operation enables microplasma portability for chemical analysis on-site (i.e., in the field). By bringing a microplasma-based instrument to the field, microplasmas are expected to cause a paradigm shift in classical chemical analysis in which samples are collected in the field and are brought to a lab for analysis [140, 141, 142, 143, 144, 145, 146, 147, 148].

Due to plasma miniaturization, a number of questions arise. For example, how small can microplasmas be made? And, how small analytical microplasmas should be made? From a technology perspective, what is the minimum voltage required to ignite and sustain a microplasma? Would substrates tolerate the required high voltage? And, what is the preferred fabrication technology?

3.3. How small atmospheric-pressure microplasmas can be made?

A plasma (Figure 7, regardless of its size) consists of two plasma sheaths (located in the vicinity of two electrodes bathed in a gas-of-interest in a gas-tight container) and a bulk plasma [133, 134, 135, 136, 137]. Shielding (or damping or screening) of the electric field arises from the presence of charged species in the plasma and from the unequal mobility of ions and electrons in the vicinity of the electrodes. Inside the plasma sheath, macroscopic electrical neutrality is likely not maintained. But outside of it (labeled bulk plasma in Figure 7), macroscopic neutrality is maintained and (time-averaged) electron and ion fluxes are roughly equal. Thus (on a time-average and per unit-volume), ne≈ni (for singly charged species). The distance (or thickness) a sheath screens electric fields is called the Debye lengthD), given by Eq. 1.

​​λ​ D​​ = ​​[​ ​ϵ​ 0​​ kT ____ ​n​ e​​ ​e​​ 2​ ​]​​​ 1/2​​E1

where k is the Boltzmann constant, T is the electron temperature, ε0 is the permeability in vacuum, ne is the electron number density and e is the charge of an electron.

To generalize, a key assumption is that sheath thickness is about the same magnitude as the Debye length. A few, what-if type thought-experiments will be used to obtain an indication on how λD changes as T and ne vary. For example, for an atmospheric pressure plasma when T = 10,000 K (with 1 eV = 11,600 K) and ne = 1016 m−3, then ​​λ​ D​​ =​ 110 μm. But when T = 5000 K and assuming that there is no thermal ionization (thus the degree of ionization is constant and the same as in the example above) with ne = 1016 m−3, then ​​λ​ D​​ =​ 80 μm. For less than atmospheric pressure operation and assuming that ne = 5 x 1014 m−3 and (for simplicity, assuming that the degree of ionization is unchanged) and that T = 5000 K, then ​​λ​ D​​ =​ 350 μm. Because plasmas cannot be made smaller than their boundary layers (per conditions outlined in Section 3.1), plasma sheaths (and Debye length) set a fundamental limit as to how small the inter-electrode distance d (Figure 7) can become, in other words, how small a microplasma can be made.

Since inter-electrode distance d must be >> λD, and for the example with λD = 110 μm and for a two-electrode operation, then the microplasma must be larger (or much larger) than 2 times λD or (for this example) it must be >> 220 μm). As d becomes ~2 times the length of the sheath, the sheath-bulk plasma structure must disappear and thus the plasma must become devoid of a bulk plasma (Figure 7), that is to become a sheath-only plasma. But in a strict interpretation of the definition of a plasma, can such an ionized gas still be called a “plasma” [134]? There are published reports of microplasmas formed in constrained cavities that are smaller than 10 μm by 10 μm [133, 134, 135, 136, 137]. This has been explained by considering that sheath-thickness scales as inter-electrode distance decreases. Several open-ended questions in this research area still remain unanswered for instance, would microplasmas the size of 10’s of μm be useful for chemical analysis? To obtain insights, perhaps this question must be re-phrased to read “how small analytical, atmospheric pressure microplasmas should be made”?

3.4. How small analytical, ambient-pressure plasmas should be made?

There are two answers to this question. One is that microchannels can be 10’s of mm long ([48, 49, 50, 51] and cited literature). Since there does not seem to be a fundamental reason why microplasmas should be constrained in μm-size cavities, microplasmas can occupy part of mm-long microchannels (Figure 8). Therefore, such microplasmas are not limited by Debye length or by plasma sheaths.

Figure 8.

(a) Simplified diagram of a microplasma and (b) microplasma formed at the end of a needle electrode (OD: 470 μm, ID: 130 μm) inside a microfluidic channel on a microfluidic chip. A Canadian 1 cent coin (about the same diameter as that of a US one-cent coin, or UK’s one-pence, or a one-cent euro) has been included for size.

The other answer involves residence time of an analyte in a microplasma (analyte = the chemical species of interest in a sample to be used for chemical analysis). Residence time (important in elemental chemical analysis) is defined as the time an analyte resides in, or is in contact with or it interacts with a microplasma. In general, as microplasma length (dictated by the inter-electrode distance or gap) decreases, so does residence time. But as residence time decreases, so does signal intensity from an analyte introduced into a microplasma. This is mainly due to a reduced interaction-time between an analyte and a microplasma. Thus, from an elemental analysis viewpoint, decreasing the length of a microplasma (e.g., by fabricating microplasmas in μm cavities) may not necessarily be beneficial in terms of signal intensity. This is significant because as signal intensity worsens, signal-to-noise ratio (SNR) degrades, thus degrading the detection limit (defined as the minimum amount or concentration that can be detected with a stated statistical confidence). The detection limit is a key figure of merit in chemical analysis. From the foregoing it can be concluded that mm-long microplasmas formed inside microfluidic channels (e.g., Figure 8) will likely be beneficial for elemental chemical analysis.

3.5. Igniting and sustaining a microplasma at atmospheric pressure

Conceptually, there are two steps involved in forming and sustaining a continuously-operated atmospheric-pressure microplasma. For instance, a microplasma must be first initiated (or “ignited”) and then it must be sustained. The minimum “ignition” (or sparking) voltage (Vb) for which the entire discharge gap is fully formed (often called “bridged”) when a uniform electric field is applied between two flat electrodes at a distance or gap (d) immersed in a gas of interest under pressure (p) can be determined using Paschen’s law (Eq. 2).

​​V​ b​​ = ​ Bpd _____________ In​[​ Apd ___________ In ​(1 / γ)​ ​]​ ​​E2

A and B are constants that depend on the properties of the gas in which the electrodes are immersed in (not accounting for any ionization due to background radiation). The values of A and B are either determined experimentally or they are calculated from literature values [139]. The coefficient γ (also known as Townsend’s coefficient) incorporates properties of the electrode material (e.g., work function) and it assumes that gas breakdown is predominantly a function of electron emission from the electrodes. In short, the two key variables in this equation are pressure (p) and inter-electrode distance (d). The product of p times d is often called “pd scaling.” An example of a Paschen curve is shown in Figure 9.

Figure 9.

Paschen curve for argon gas and for a 2.8 mm inter-electrode gap (d) as a function of pd.

Paschen’s law applies to electrical discharges formed at low-pressures. In high-vacuum or at high pressures (e.g., atmospheric), Paschen’s law fails ([139] and references herein). There are also deviations from the behavior predicted by Eq. 2 when kHz or MHz ac voltages are used or when μm inter-electrode distances (or gaps d) are employed [139]. Undeniably, there are limits to applicability of Paschen’s law. Despite of these limitations, Paschen’s law (presumably, the only choice) can be used to obtain rough estimates of the magnitude of the voltage required to ignite (or initiate) an atmospheric pressure plasma. Thus it can be used as an aid in the design of appropriate power supplies. For instance, when the electrodes are made from Iron (Fe) and the inter-electrode distance d is 2.8 mm, and the discharge gas is Argon (Ar) at (or near) atmospheric pressure, the minimum voltage (Vb) required for gas breakdown (or for microplasma ignition) is about 6000 V. As the inter-electrode distance d decreases from 2.8 to 1 mm (and by keeping all else constant), Vb drops to about 2400 V, and when d further decreases to 0.5 mm, Vb drops to about 1400 V. It should be emphasized that gas breakdown at the minimum voltage Vb is not always necessary and that (once ignited), to sustain a microplasma lower voltages are typically required. An example is the ballast used in fluorescent lights.

Advertisement

4. Microplasma formation inside fluidic channels

The key idea behind microplasma miniaturization [138] is to obtain analytical performance about equal to that of lab-scale ICP-optical emission spectrometry (ICP-OES) systems [132] but by using self-igniting, low-power, low-cost, small-size, light-weight, continuous-flow and low gas-consumption (e.g., 250 mL/min) atmospheric-pressure microplasmas. The expectation is that such microplasmas can be used for “taking part of the lab to the sample” types of applications [140, 141, 142].

Based on these ideas, we fabricated and tested a variety of battery-operated, atmospheric pressure, self-igniting, mm-length microplasmas in fluidic channels [143, 144, 145, 146, 147, 148, 149, 150, 151, 152, 153, 154, 155, 156]. Due to their mm-length, plasma sheath and Debye length are not of a concern. In addition to being “cold”, their high surface area-to-volume ratio further facilitates heat dissipation, thus facilitating use of polymeric substrates and 3D-printing fabrication methods. Example microplasmas fabricated in a variety of substrates will be discussed next.

4.1. Microplasmas in fluidic channels on amorphous substrates

For microplasmas formed inside a microfluidic channel on a chip, a dual substrate approach was used (Figure 10). Briefly, cleanroom-technologies (Figure 2) were employed to define and to sputter-deposit Au electrodes E1 and E2 (Figure 10a). Holes were drilled for the inlet and the outlet. On the bottom wafer, a chemically etched microchannel was formed. The top and bottom wafers (Figure 10a and b) were aligned so that the central part of the etched channel matched the protruding part of electrodes E1 and E2. Then the wafers were bonded together (Figure 10c) [143] and glass-tubes were affixed to the inlet and outlet holes (Figure 10d). The inlet was connected to a gas-supply (Ar-3%H2) that was used as the microplasma gas and as the sample-introduction carrier-gas. Upon application of electrical power, the microplasma self-ignited, it was formed between electrodes E1 and E2 and was sustained by continuous application of electrical power (~10 W). To avoid electrode breakage, a high-voltage ac [143] was used.

Figure 10.

(a) Top chip showing electrodes E1 and E2, (b) bottom chip showing the etched microchannel, (c) the top and bottom chips bonded together (the microplasma was formed between electrodes E1 and E2, and (d) an “angle” view of the two bonded chips.

4.2. Postage stamp-sized microplasmas on polymeric substrates

To reduce ownership, operation and fabrication costs, we developed and evaluated a variety of microplasmas on polymeric substrates (e.g., Figures 11 and 12) [144, 145, 146]. Although a critical microplasma dimension was in μm-meter regime (Figure 11), these microplasmas were formed inside millifluidic channels (e.g., ~2 mm wide). This was done for rapid prototyping [109] and to avoid accidental contact of the microplasma with the channel-walls (important during testing). Once prototypes were produced, channel width was never revisited. Although polymeric substrates have high dielectric strength, to address poor transmission of polymers in the UV, the channels were fitted with a quartz plate (Figure 11b).

Figure 11.

(a) Postage stamp-sized polymeric 3D-chips and (b) microplasma formed between electrodes E1 and E2. Depending on operating conditions, microplasmas with diameters of (b) ~750 μm, (c) ~400 μm and (d) ~200 μm were formed. A 1 cent coin was included for size, the microplasma fit inside the letter a of the coin.

Figure 12.

3D printed microplasma on a hybrid 3D-chip formed between electrodes E1 and E2 (coin has been included for size, the microplasma fit inside the letter a of the 1 cent coin).

4.3. Millifluidic channels in 3D-printed chips for microplasmas

3D-printing [109, 110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120, 121, 122] was accomplished using a 3D-printer (~$1000) to rapidly prototype 3D-chips in a few hours (or less), thus obviating the need for cleanrooms and lithography. We used 3D-printing to fabricate hybrid chips (fitted with a quartz plate and needle electrodes) for microplasma formation in millifluidic channels [146, 147, 149]. An example is shown in Figure 12.

Advertisement

5. Nanofluidics

The nanoscale [157, 158, 159, 160, 161, 162, 163, 164, 165, 166, 167, 168, 169, 170, 171, 172, 173, 174, 175, 176, 177, 178, 179, 180, 181, 182, 183, 184, 185, 186, 187, 188, 189, 190, 191, 192, 193, 194] is a natural extension of the microscale (Figure 1) and it is defined as the science, technology and application of transport phenomena and of fluid-flow in channels ≤100 nm or around nano-size objects [158, 170]. This is not universally accepted, many consider nano-size as anything with one critical dimension ≤1 μm. The range between 100 nm and 1 μm is sometimes referred to as “extended nanofluidics” [181]. Nanofluidics is not new, although the name is [159, 160].

5.1. The science of nanofluidics

In nanofluidics, size (or scale) is important, likely more so than in microfluidics. For instance, at the nano-scale many dimensions of molecules are of similar size as the nano-fluidic channels that constrain them (Figure 1). A few scientific questions that being addressed include: How do properties of individual atoms, ions or molecules, manifest themselves as they are confined in spaces (roughly) of their own size? Would quantum effects become important [173]? Since pressure is not used to force fluids through nanochannels, should electrokinetic flow be preferred? And, as surface-to-volume ratio increases significantly (over microchannels), what is the effect of surface-charge on ions or molecules confined in nanochannels? What is the effect of surface roughness on fluid-flow? And, how do surfaces interact with ions or molecules so close to them? What are the best surface modification approaches? What is the effect of van der Waals forces and of the electric double layer (EDL) at the nm-scale? Some questions arising from technology include: how would one introduce very small volumes of analytical samples into nanofluidic channels? To facilitate discussion, assume a cylindrical nanochannel with 100 nm diameter and 1 μm length. In this case, the volume is 100 atto Liter (aL). How would one introduce an aL volume sample into a nanochannel without evaporation of some of the analyte or of the solvent? Due to the infinitesimal volumes used, would single atom, ion or molecule measurement techniques be essential? In support of this, it has been estimated that in a liquid the volume of a cube with dimensions 100 nm by 100 nm by 100 nm, there are only ~6 analytes when the concentration of the analyte is 1 μm [160]. Would sample separation, pre-concentration and use of highly-sensitive detection techniques (e.g., laser induced fluorescence or LIF) become essential?

5.2. Technology for fabrication of nanofluidic channels

According to the National Science Foundation (NSF) in the US and its National Nanotechnology Initiative (NNI), nanotechnology involves “the application of scientific knowledge to manipulate and control matter in the nanoscale” [158, 170], more or less arbitrarily defined at ≤100 nm [158, 170]. Nanofluidics often falls under nano electro mechanical systems (NEMS) [164, 177, 178, 179, 184] typically fabricated using complementary metal oxide semiconductor (CMOS) technology [177, 178].

For nanofabrication, many technologies have been described [159, 160, 161, 162, 165, 166, 169, 171, 172, 180]. Some of them are nano-specific [159, 169] for example, scanning probe lithography (SPL) [161], etching using a focused ion beam (FIB) [171] and nanoimprinting [159]. In many cases use of a cross-sectional area of a nanochannel is preferred (e.g., 10 nm by 10 nm) rather than aspect ratio. Nanofluidic channels can be nanofabricated using either top-down or bottom-up approaches.

Top down methods of fabrication of nanochannels: By analogy to micromachining, these fabrication methods include bulk nanomachining; surface nanomachining; and, imprinting (as is typical of soft-lithography) [159, 161, 162, 163, 164, 165, 169]. A top plate is typically used to cover nanochannels but due the nanosize of the channels and unless precautions are taken, channels may plug-up during bonding.

Bottom up methods of nanostructure formation: in some cases molecules can be “convinced” to self-assemble into nanostructures by controlling chemical conditions [160, 162].

Associated nanofabrication technologies include scanning probe lithography (SPL) [161], electron beam lithography (EBL) [159] and dip-pen nanolithography [185]. Such approaches are typically used to bypass the diffraction-limit of photolithography or to provide new capabilities.

5.3. Applications of nanofluidics

In addition to enabling fundamental studies of fluid-flow and of transport phenomena (with many studies aimed at the study of naturally occurring processes in biological nanochannels), many applications are aimed at bio-sciences, bio-nano-technology and bio-analytical chemistry where applications exist in abundance [166, 167, 168, 169]. Applications outside of classical nano-fluidics include nano-pores (e.g., for bio-applications and DNA sequencing) [186, 187, 188, 189, 190, 191, 192] and even for the study of fluid-flow in nano-porous media [193, 194]. For chemical analysis, NEMS have been developed for single protein mass spectrometry [174] and for airborne nanoparticle detection [176]. From this short list it can be concluded that nanofluidics has the potential to become a disruptive technology worthy of further investigation.

Advertisement

6. Conclusions

Microfluidics continues to receive attention in science and technology due to its many applications. And as shown, it has the potential to find applicability in constraining atmospheric-pressure microplasmas in 2D-microfluidic channels (Figures 8 and 10) or in 3D-millifluidic chips (Figures 11 and 12). Future developments include coupling of standard CMOS fabrication technology [179, 183, 184, 195, 196, 197] with microfluidics or millifluidics, thus allowing integration of fluidics with electronics. Microinstruments are those with at least one critical (or essential) component operating in the micro-regime. For nanofluidics as may be applied to chemical analysis, it appears that it will be best if nanofluidic channels was packaged alongside microfluidic channels.

It is envisioned that future fluidics (Figure 1) will be embedded within portable micro- or nano-instruments for measurements on-site (i.e., in the field). Such instruments will have (some) energy autonomy [198, 199, 200], will incorporate some “smarts” [201] (e.g., based on Artificial Intelligence and Deep Learning) and will have wireless capability [202] so that they can become a part of the Internet of Things (IoT) [200, 201, 202, 203]. Clearly, fluidics (e.g., milli-, micro- or nano-) have the potential to become critical components of mobile (or even wearable) instruments that are “cheap, smart and under wireless control” [139].

Advertisement

Acknowledgments

Financial assistance from NSERC (Natural Sciences and Engineering Research Council) of Canada is gratefully acknowledged. A special thank you to Professor (now Emeritus, ETH Zurich, Switzerland) Dr. Henry Baltes for the many enlightening discussions we had on MEMS and on miniaturization.

References

  1. 1. Bruus H. Theoretical Microfluidics. Oxford, UK: Oxford University Press; 2008
  2. 2. Kumar CS. Microfluidics in Nanotechnology. New York: Wiley; 2010
  3. 3. Tabeling P. Introduction to Microfluidics. Oxford, UK: Oxford University Press; 2011
  4. 4. Mitra SK, Chakraborty S, editors. Microfluidics and Nanofluidics. Florida: CRC Press; 2011
  5. 5. Madou MJ. Fundamentals of Microfabrication. Florida: CRC Press; 2011
  6. 6. Kirby BJ. Micro- and Nano-Scale Fluid Mechanics. Cambridge, UK: Cambridge University Press; 2013
  7. 7. Li X(J), Zhou Y. Microfluidics for Biomedical Applications. Cambridge, UK: Woodhead; 2013
  8. 8. Conlisk AT. Essentials of Micro- and Nano-Fluidics. Cambridge, UK: Cambridge University Press; 2013
  9. 9. Lagally E editor. Microfluidics and Nanotechnology: Biosensing to the Single Molecule Limit (Devices, Circuits, and Systems). Florida: CRC Press; 2014
  10. 10. Li D. Encyclopedia of Microfluidics and Nanofluidics. Vol. five volume series. New York: Springer; 2015. DOI: 10.1007/981-1-4614-5491-5
  11. 11. Berthier J, Brakke KA, Berther E. Open Microfluidics. Massachusetts: Scrivener publishing; 2016
  12. 12. Dixit CK, Kaushik A. Microfluidics for Biologists. Germany: Springer; 2016
  13. 13. Panigrahi PK. Transport Phenomena in Microfluidic Systems. New York: Wiley; 2016
  14. 14. Giri B. Laboratory Methods in Microfluidics. Netherlands: Elsevier; 2017
  15. 15. Lei K-M, Mak P-I, Law M-K, Martins RP. Handheld Total Chemical and Biological Analysis Systems: Bridging NMR, Digital Microfluidics, and Semiconductors. Germany: Springer; 2017
  16. 16. Chu L-Y, Wang W. Microfluidics for Advanced Functional Polymeric Materials. Germany: Wiley-VCH; 2017
  17. 17. Piraino F, Selimović Š, editors. Diagnostic Devices with Microfluidics. Florida: CRC Press; 2017
  18. 18. Petersen KE. Silicon as a mechanical material. Proceedings of the IEEE. 1982;70(5):420-457. DOI: 10.1109/PROC.1982.12331
  19. 19. Angell JB, Terry SC, Barth PW. Silicon micromechanical devices. Scientific American. 1983;248(4):44-55. DOI: 10.1038/scientificamerican0483-44
  20. 20. Manz A, Harrison DJ, Verpoorte EMJ, Fettinger JC, Lüdi APH, Widmer HM. Planar chips for miniaturization and integration of separation techniques into monitoring systems: CE on a chip. Journal of Chromatography A. 1992;593(1-2):253-258. DOI: 10.1016/0021-9673(92)80293-4
  21. 21. Harrison DJ, Fluri K, Seiler K, Fan Z, Effenhauser CS, Manz A. Micromachining a miniaturized capillary electrophoresis. Science. 1993;1261(5123):895-897. DOI: 10.1126/science.261.5123.895
  22. 22. Khandurina J, McKnight TE, Jacobson SC, Waters LC, Foote RS, Ramsey JM. Integrated system for rapid PCR-based DNA analysis in microfluidic devices. Analytical Chemistry. 2000;72(13):2995-3000. DOI: 10.1021/ac991471a
  23. 23. Oleschuk RD, Shultz-Lockyear LL, Ning Y, Jed Harrison D. Trapping of bead-based reagents within microfluidic systems. Analytical Chemistry. 2000;72(3):585-590. DOI: 10.1021/ac990751n
  24. 24. Verpoorte E, De Rooij NF. Microfluidics meets MEMS. Proceedings of the IEEE. 2003;91(6):930-953. DOI: 10.1109/JPROC.2003.813570
  25. 25. Squires TM, Quake SR. Microfluidics: Fluid physics at the nanoliter scale. Reviews of Modern Physics. 2005;77(3):977-1026. DOI: 0.1103/RevModPhys.77.977
  26. 26. Whitesides GA. The origins and the future of microfluidics. Nature. July 27, 2006;442:368-373. DOI: 10.1038/nature05058
  27. 27. Janesik D, Franzke J, Manz A. Scaling and the design of miniaturized chemical-analysis systems. Nature. 27 July, 2006;442:374-380. DOI: 10.1038/nature05059
  28. 28. Yuan X, Oleschuk RD. Advances in microchip liquid chromatography. Analytical Chemistry. 2018;90(1):283-301. DOI: 0.1021/acs.analchem.7b04329
  29. 29. Qin N, Wen JZ, Ren CL. Highly pressurized partially miscible liquid-liquid flow in a micro-T-junction. II. Physical Review E. 2017;95(4):043111. DOI: 10.1103/PhysRevE.95.043111
  30. 30. Ramsey JM. The burgeoning power of the shrinking laboratory. Nature Biotechnology. 1999;17:1061-1062. DOI: 10.1038/15044
  31. 31. Whitesides GM. The ‘right’ size in nanobiotechnology. Nature Biotechnology. 2003;21:1161-1165. DOI: 10.1038/nbt872
  32. 32. McClain MA, Culbertson CT, Jacobson SC, Allbritton NL, Sims CE, Ramsey JM. Microfluidic devices for the high-throughput cell analysis. Analytical Chemistry. 2003;75(21):5646-5655. DOI: 10.1021/ac0346510
  33. 33. Neethirajan S, Kobayashi I, Nakajima M, Wu D, Nandagopal S, Lin F. Microfluidics for food, agriculture and biosystems industries. Lab on a Chip. 2011;11:1574-1586. DOI: 10.1039/C0LC00230E
  34. 34. Morin SA, Shepherd RF, Kwok SW, Stokes AA, Nemiroski A, Whitesides GM. Camouflage and display for soft machines. Science. 2012;337(6096):828-832. DOI: 10.1126/science.1222149
  35. 35. Yin H, Marshall D. Microfluidics for single cell analysis. Current Opinion in Biotechnology. 2012;23(1):110-119. DOI: 10.1016/j.copbio.2011.11.002
  36. 36. Jokerst JC, Emory JM, Henry CS. Advances in microfluidics for environmental analysis. Analyst. 2012;137:24-34. DOI: 10.1039/C1AN15368D
  37. 37. Sun Y, Jaffray DA, Yeow JTW. The design and fabrication of carbon nanotube based X-ray cathode. IEEE Transactions on Electronic Devices. 2013;60(1):464-470. DOI: 10.1109/TED.2012.2226036
  38. 38. Zalesskiy SS, Danieli E, Blümich B, Ananikov VP. Miniaturization of NMR systems. Chemical Reviews. 2014;114(11):5641-5694. DOI: 10.1021/cr400063g
  39. 39. Sackmann EK, Fulton AI, Beebe DJ. The present and future role of microfluidics in biomedical research. Nature. 2014;507:181-189. DOI: 10.1038/nature13188
  40. 40. Sun J-Y, Keplinger C, Whitesides GM, Suo Z. Ionic skin. Advanced Materials. 2014;26(45):7608-7614. DOI: 10.1002/adma.201403441
  41. 41. Syms RRA, Wright S. MEMS mass spectrometers: The next wave of miniaturization. Journal of Micromechanics and Microengineering. 2016;26:023001. DOI: 10.1088/0960-1317/26/2/023001
  42. 42. Zhang J, Yan S, Yuan D, Alici G, Nguyen N-T, Warkiani ME, Li W. Fundamentals and applications of inertial microfluidics: Review. Lab on a Chip. 2016;16:10-34. DOI: 10.1039/C5LC01159K
  43. 43. Sohi AN, Nieva PM. Frequency response of curved bilayer microcantilevers with applications to surface stress measurement. Journal of Applied Physics. 2016;119(4):044503. DOI: 10.1063/1.4940951
  44. 44. Chiu DT, deMello AJ, Di Carlo D, Doyle PS, Hansen C, Maceiczyk RM, Wootton RCR. Small but perfectly formed? Journal of Chemistry. 2017;2(2):201-223. DOI: 10.1016/j.chempr.2017.01.009
  45. 45. Rane AS, Rutkauskaite J, deMello A, Stavrakis S. High-throughput multi-parametric imaging flow cytometry. Chem. 2017;3:588-602. DOI: 10.1016/j.chempr.2017.08.005
  46. 46. Miled A, Greener J. Recent advancements towards full-system microfluidics. Sensors. 2017;17:1-7. DOI: 10.3390/s17081707
  47. 47. Weiss M, Frohnmayer JP, Benk LT, Haller B, Janiesch J-W, Heitkamp T, Börsch M, Lira RB, Dimova R, Lipowsky R, Bodenschatz E, Baret J-C, Vidakovic-Koch T, Sundmacher K, Platzman I, Spatz JP. Sequential bottom-up assembly of mechanically stabilized synthetic cells by microfluidics. Nature Materials. October 16, 2017 (online). DOI: 10.1038/nmat5005
  48. 48. Karanassios V, Sharples JT. Microchannels and microcells for gaseous microsamples. Sensors & Materials. 1997;9(6):363-378. http://myukk.org/SM2017/sm_pdf/SM298.pdf
  49. 49. Karanassios V, Sharples JT, Nathan A. In situ ultrasound assisted etching of <100> Si wafers by KOH. Sensors & Materials. 1997;9(7):427-436. http://myukk.org/SM2017/sm_pdf/SM303.pdf
  50. 50. Karanassios V, Mew G. Anisotropic wet chemical etching of Si for chemical analysis applications. Sensors & Material. 1997;9(7):395-416. http://myukk.org/SM2017/sm_pdf/SM301.pdf
  51. 51. Swart NR, Stevens M, Nathan A, Karanassios V. A flow-insensitive thermal conductivity microsensor. Sensors & Materials. 1997;9(6):387-394. http://myukk.org/SM2017/sm_pdf/SM300.pdf
  52. 52. French PJ, Sarro PM. Surface versus bulk micromachining: The contest for suitable applications. Journal of Micromechanics and Microengineering. 1998;8:4552. DOI: 10.1088/0960-1317/8/2/002
  53. 53. Bustillo JM, Howe RT, Muller RS. Surface micromachining for microelectromechanical systems. Proceedings of the IEEE. 1998;86(8):1552-1574. DOI: 10.1109/5.704260
  54. 54. Kovacs GTA, Maluf NI, Petersen KE. Bulk micromachining of silicon. Proceedings of the IEEE. 1998;86(8):1536-1551. DOI: 10.1109/5.704259
  55. 55. Pethig R, Burt JPH, Parton A, Rizvi N, Talary MS, Tame JA. Development of biofactory-on-a-chip. Journal of Micromechanics and Microengineer. 1998;8(2):57-63. DOI: 10.1088/0960-1317/8/2/004
  56. 56. Klank H, Kuttera JP, Geschke O. CO2-laser micromachining for rapid production of PMMA-based microfluidic systems. Lab on a Chip. 2002;2:242-246. DOI: 10.1039/B206409J
  57. 57. Cheng J-Y, Wei C-W, Hsu K-H, Young T-H. Direct-write laser micromachining. Sensors Actuators B. 2004;99(10):186-196. DOI: doi.org/10.1016/j.snb.2003.10.022
  58. 58. Osellame R, Cerullo G, Ramponi R. Femtosecond Laser Micromachining: Photonic and Microfluidic Devices in Transparent Materials. Germany: Springer; 2012
  59. 59. Becker EW, Ehrfeld W, Hagmann P, Maner A, Münchmeyer D. Fabrication of microstructures by LIGA. Microelectronic Engineering. 1986;4(1):35-56. DOI: 10.1016/0167-9317(86)90004-3
  60. 60. Lorenz H, Despont M, Fahrni N, Brugger J, Vettiger P, Renaud P. High-aspect-ratio, negative-tone near-UV photoresist and its applications for MEMS. Sensors Actuators A. 1998;64(1):33-39. DOI: 10.1016/S0924-4247(98)80055-1
  61. 61. Malek CK, Saile V. Applications of LIGA to precision manufacturing of high-aspect-ratio micro-components: A review. Microelectronics Journal. 2004;35(2):131-143. DOI: 10.1016/j.mejo.2003.10.003
  62. 62. Li X, Abe T, Esashi M. Deep reactive ion etching of pyrex glass using SF6 plasma. Sensors Actuators A. 2001;87(3):139-145. DOI: 10.1016/S0924-4247(00)00482-9
  63. 63. Laermer F, Urban A. Challenges, developments and applications of silicon deep reactive ion etching. Microelectronic Engineering. 2003;67-68:349-355. DOI: 10.1016/S0167-9317(03)00089-3
  64. 64. Marty F, Rousseau L, Saadany B, Mercier B, Français O, Mita Y, Bourouina T. Advanced etching of silicon based on DRIE etching of Si high aspect ratio microstructures and 3D micro- and nanostructures. Microelectronics Journal. 2005;36(7):637-677. DOI: 10.1016/j.mejo.2005.04.039
  65. 65. Laermer F, Franssila S, Sainiemi L, Kolari K. Chapter 21: Deep Reactive Ion Etching. In: Lindroos V, Motooka T, Franssila S, Paulasto-Krockel M, Tilli M, Airaksinen V-M, editors. Handbook of Silicon Based MEMS Materials and Technologies. 2nd ed. 2015. pp. 444-469. DOI: 10.1016/B978-0-323-29965-7.00021-X
  66. 66. Lorenz H, Despont M, Fahrani N, LaBianca N, Renaud P, Vettiger P. SU-8: A low-cost negative resist for MEMS. Journal of Micromechanics and Microengineering. 1991;7:121-124. DOI: 10.1088/0960-1317/7/3/010
  67. 67. Lee KY, LaBianca N, Rishton SA. Micromachining applications of a high resolution ultra-thick photoresist. Journal of Vacuum Science and Technology B. 1995;13(6):3000-3006. DOI: 10.1116/1.588295
  68. 68. Tay FHE, van Kan JA, Watt F, Choong WO. A novel micro-machining method for the fabrication of thick-film SU-8 micro-channels. Journal of Micromechanics and Microengineering. 2001;11:27-32. DOI: 10.1088/0960-1317/11/1/305
  69. 69. Xu J, Locascio L, Gaitan M, Lee CS. Room-temperature imprinting method for plastic microchannel fabrication. Analytical Chemistry. 2000;72(8):1930-1933. DOI: 10.1021/ac991216q
  70. 70. Becker H, Locascio LE. Polymeric microfluidic devices. Talanta. 2002;86(2):267-287. DOI: 10.1016/S0039-9140(01)00594-X
  71. 71. Mills CA, Martinez E, Bessueille F, Villanueva G, Bausells J, Samitier J, Errachid A. Production of structures for microfluidics using polymer imprint techniques. Microelectronic Engineering. 2005;78-79:695-700. DOI: 10.1016/j.mee.2004.12.087
  72. 72. Bridget XS, Weichun AP, Hector Y, Becerril A, Woolley AT. Rapid prototyping of PMMA microfluidic systems using solvent imprinting and bonding. Journal of Chromatography A. 2007;1162(2):162-166. DOI: 10.1016/j.chroma.2007.04.002
  73. 73. Curtis C, Eshaque B, Badali K, Karanassios V. Rapid prototyping of a microfluidic Venturi pump imprinted on polymeric chips. Proceedings of SPIE. 2012;8366:83660. DOI: 10.1117/12.91956
  74. 74. Becker H, Heim U. Hot embossing as a method for the fabrication of polymer high aspect ratio structures. Sensors Actuators A. 2000;83(1-3):130-135. DOI: 10.1016/S0924-4247(00)00296-X
  75. 75. Lee G-B, Chen S-H, Huang G-R, Sung W-C, Lin Y-H. Microfabricated plastic chips by hot embossing methods. Sensors Actuators B. 2001;75(1-2):142-148. DOI: 10.1016/S0925-4005(00)00745-0
  76. 76. Qi S, Liu X, Ford S, Barrows J, Thomas G, Kelly K, McCandless A, Lian K, Goettert J, Soper SA. Microfluidic devices fabricated in PMMA using hot-embossing with integrated sampling capillary. Lab on a Chip. 2002;2:88-95. DOI: 10.1039/B200370H
  77. 77. Xia Y, Whitesides GM. Soft lithography. Annual Review of Material Science. 1998;28:153-184. DOI: 10.1146/annurev.matsci.28.1.153
  78. 78. Unger MA, Chou H-P, Thorsen T, Scherer A, Quake SR. Monolithic microfabricated valves and pumps. Science. 2000;288(5463):113-116. DOI: 10.1126/science.288.5463.113
  79. 79. Yang P, Wirnsberger G, Huang HC, Cordero SR, McGehee MD, Scot B. Mirrorless lasing from mesostructured waveguides in COC patterned by soft lithography. Science. 2000;287(5452):465-467. DOI: 10.1126/science.287.5452.465
  80. 80. Whitesides GM, Ostuni E, Takayama S, Jiang X, Ingber DE. Soft lithography in biology and biochemistry. Annual Review of Biomedical Engineering. 2001;3:335-373. DOI: 10.1146/annurev.bioeng.3.1.335
  81. 81. Odom TW, Love JC, Wolfe DB, Paul KE, Whitesides GM. Improved pattern transfer in soft lithography. Langmuir. 2002;18(13):5314-5320. DOI: 10.1021/la020169l
  82. 82. Qin D, Xia Y, Whitesides GM. Soft-lithography for micro- and nanoscale patterning. Nature Protocols. 2003;5(3):491-502. DOI: 10.1038/nprot.2009.234
  83. 83. Li C, Yang Y, Craighead HG, Lee KH. Isoelectric focusing in COC microfluidic channels. Electrophoresis. 2005;26(9):1800-1806. DOI: 10.1002/elps.200410309
  84. 84. Teh S-Y, Lin R, Hung L-H, Lee AP. Droplet microfluidics. Lab on a Chip. 2008;8:198-220. DOI: 10.1039/B715524G
  85. 85. Theberge AB, Courtois F, Schaerli Y, Fischlechner M, Abell C, Hollfelder F, Huck WTS. Microdroplets in microfluidics. Angewandte Chemie International. 2010;49:5846-5868. DOI: 10.1002/anie.200906653
  86. 86. Seemann R, Brinkmann M, Pfohl T, Herminghaus S. Droplet based microfluidics. Reports on Progress in Physics. 2011;75:016601. DOI: 10.1088/0034-4885/75/1/016601
  87. 87. Zhua Y, Fang Q. Analytical detection techniques for droplet microfluidics: A review. Analytica Chimica Acta. 2013;787:24-35. DOI: 10.1016/j.aca.2013.04.064
  88. 88. Courtney M, Chen X, Chan S, Mohamed T, Rao PN, Ren CL. Droplet microfluidic system for drug screening applications. Analytical Chemistry. 2017;89:910-915. DOI: 10.1021/acs.analchem.6b04039
  89. 89. Ding Y, Choo J, deMello AJ. From single-molecule detection to next-generation sequencing. Microfluidics Nanofluidics. 2017;21(58). DOI: 10.1007/s10404-017-1889-4
  90. 90. Pollack MG, Fair RB. Electrowetting-based actuation of liquid droplets for microfluidic applications. Applied Physics Letters. 2000;77:1725. DOI: 10.1063/1.1308534
  91. 91. Fair RB. Digital microfluidics: Is a true lab-on-a-chip possible? Microfluidics and Nanofluidics. 2007;3:245-281. DOI: 10.1007/s10404-007-0161-8
  92. 92. Wheeler AR. Putting electrowetting to work. Science. October 24, 2008;322:539-540. DOI: 10.1126/science.1165719
  93. 93. Yang H, Luk VN, Abelgawad M, Barbulovic-Nad I, Wheeler AR. A world-to-chip interface for digital microfluidics. Analytical Chemistry. 2009;81(3):1061-1106. DOI: 10.1021/ac802154h
  94. 94. Choi K, Ng AHC, Fobel R, Wheeler AR. Digital microfluidics. Annual Review of Analytical Chemistry. 2012;5:413-440. DOI: 10.1146/annurev-anchem-062011-143028
  95. 95. Berthier J. Micro-drops and Digital Microfluidics. 2nd ed. Netherlands: Elsevier; 2013
  96. 96. Madou M, Zoval J, Jia G, Kido H, Kim J, Kim N. Lab on a CD. Annual Review of Biomedical Engineering. 2006;8:601-628. DOI: 10.1146/annurev.bioeng.8.061505.095758
  97. 97. Gorkin R, Park J, Siegrist J, Amasia M, Lee BS, Park J-M, Kim JKH, Madou M, Cho Y-K. Centrifugal microfluidics. Lab on a Chip. 2010;10:1758-1773. DOI: 10.1039/B924109D
  98. 98. Gorkin R, Clime L, Madou M, Kido H. Pneumatic pumping in centrifugal microfluidic platforms. Microfluidics and Nanofluidics. 2010;9(2-3):541-549. DOI: 0.1007/s10404-010-0571-x
  99. 99. Kong MCR, Salin ED. Pneumatically pumping fluids radially inward on centrifugal microfluidic platforms in motion. Analytical Chemistry. 2010;82(19):8039-8041. DOI: 10.1021/ac102071b
  100. 100. Wang L, Li PCH. Microfluidic DNA microarray analysis: A review. Analytica Chimica Acta. 2011;687(1):12-27. DOI: doi.org/10.1016/j.aca.2010.11.056
  101. 101. Schwemmer F, Zehnle S, Mark D, von Stetten F, Zengerle R, Paust N. A microfluidic timer for centrifugal microfluidics. Lab on a Chip. 2015;15:1545-1553. DOI: 10.1039/C4LC01269K
  102. 102. Martinez AW, Phillips ST, Whitesides GM. Three-dimensional microfluidic devices fabricated in layered paper and tape. PNAS. 2008;105(50):19606-19611. DOI: 10.1073/pnas.0810903105
  103. 103. Carrilho E, Martinez AW, Whitesides GM. Understanding wax printing. Analytical Chemistry. 2009;81(16):7091-7095. DOI: 0.1021/ac901071p
  104. 104. Martinez AW, Phillips ST, Whitesides GM, Carrilho E. Diagnostics for the developing world: Microfluidic paper-based analytical devices. Analytical Chemistry. 2010;82(1):3-10. DOI: 10.1021/ac9013989
  105. 105. Nie Z, Nijhuis CA, Gong J, Chen X, Kumachev A, Martinez AW, Narovlyansky M, Whitesides GM. Electrochemical sensing in paper-based microfluidic devices. Lab on a Chip. 2010;10:477-483. DOI: 10.1039/B917150A
  106. 106. Dungchai W, Chailapakul O, Henry CS. A low-cost, simple, and rapid fabrication method for paper-based microfluidics. Analyst. 2011;136:77-82. DOI: 10.1039/C0AN00406E
  107. 107. Mao X, Huang TJ. Microfluidic diagnostics for the developing world. Lab on a Chip. 2012;12:1412-1416. DOI: 10.1039/C2LC90022J
  108. 108. Hamedi MM, Ainla A, Güder F, Christodouleas DC, Fernández-Abedul MT, Whitesides GM. Integrating electronics and microfluidics on paper. Advanced Materials. 2016;28(25):5054-5063. DOI: 10.1002/adma.201505823
  109. 109. Weagant S, Li L, Karanassios V. Rapid Prototyping of Hybrid, Plastic-Quartz 3D-Chips for Battery-Operated Microplasmas. London, UK: InTech Publishing; 2011. Chapter 10. pp. 1-18. DOI: 10.5772/24994
  110. 110. Kitson PJ, Rosnes MH, Cronin L. Configurable 3D-printed millifluidic and microfluidic ‘lab on a chip’ reactionware devices. Lab on a Chip. 2012;12:3267-3271. DOI: 10.1039/c2lc40761b
  111. 111. Au AK, Lee W, Folch A. Mail-order microfluidics: Evaluation of stereolithography for the production of microfluidic devices. Lab on a Chip. 2014;14:1294-1301. DOI: 10.1039/C3LC51360B
  112. 112. Bhargava KC, Thompson B, Malmstadt N. Discrete elements for 3D microfluidics. PNAS. 2014;111(42):15013-15018. DOI: 10173/pnas/14114764111
  113. 113. Meng C, Ho B, Ng SH, Lia KHH, Yoon Y-J. 3D printed microfluidics for biological applications. Lab on a Chip. 2015;15:3627-3637. DOI: 10.1039/C5LC00685F
  114. 114. Au AK, Huynh W, Horowitz LF, Folch A. 3D-printed microfluidics. Angewandte Chemie International. 2016;55(12):3862-3881. DOI: 10.1002/anie.201504382
  115. 115. Ahmadian A, Adam Y, William P, Wong W, Nguen T, Pan Y, Xu J. 3D printing. Microfluidics Nanofluidics. 2016;20(50). DOI: 10.1007/s10404-016-1715-4
  116. 116. Shatford R, Karanassios V. 3D printing in chemistry: Past, present and future. Proceedings of SPIE. 2016;9855:98550B-98560. DOI: 10.1117/12.2224404
  117. 117. Singh M, Tong Y, Webster K, Cesewski E, Haring AP, Laheri S, Carswell B, O'Brien TJ, Aardema CH, Senger RS, Robertson JL, Johnson BN. 3D printed conformal microfluidics for isolation and profiling of biomarkers. Lab on a Chip. 2017;17:2561-2571. DOI: 10.1039/C7LC00468K
  118. 118. Mattio E, Robert-Peillard F, Puzio CBK, Margaillan A, Brach-Papa C, Knoery J, Boudenne J-L, Coulom B. 3D-printed flow system for determination of lead in natural waters. Talanta. 2017;168:298-302. DOI: 10.1016/j.talanta.2017.03.059
  119. 119. Macdonald NP, Cabot JM, Smejkal P, Guijt RM, Paull B, Breadmore MC. Comparing microfluidic performance of three-dimensional (3D) printing platforms. Analytical Chemistry. 2017;89(7):3858-3866. DOI: 10.1021/acs.analchem.7b00136
  120. 120. Beauchamp MJ, Nordin GP, Woolley AT. Moving from millifluidic to truly microfluidic sub-100-μm cross-section 3D printed devices. Analytical and Bioanalytical Chemistry. 2017;409:4311-4319. DOI: 10.1007/s00216-017-0398-3
  121. 121. Gong H, Bickham BP, Woolley AT, Nordin GP. Custom 3D printer and resin for 18 μm × 20 μm microfluidic flow channels. Lab on a Chip. 2017;17:2899-2909. DOI: 10.1039/C7LC00644F
  122. 122. Salentijn GI-J, Oleschuk RD, Verpoorte E. 3D-printed paper spray ionization cartridge with integrated desolvation feature and ion optics. Analytical Chemistry. 2017;89(21):11419-11426. DOI: 10.1021/acs.analchem.7b02490
  123. 123. Howston AM. Introduction to Gas Discharges. Oxford, UK: Pergamon Press; 1965
  124. 124. Marcus RK. Glow Discharge Spectroscopies. New York: Plenum Press; 1993
  125. 125. Goldston RJ, Rutherfrd PH. Plasma Physics. Pennsylvania: Institute of Physics Publishing; 1995
  126. 126. Smirnov BM. Physics of Ionized Gases. New York: Wiley; 2001
  127. 127. Bittecourt JA. Fundamentals of Plasma Physics. 3rd ed. New York: Springer-Verlag; 2004
  128. 128. Lieberman MA, Lichtenberg AJ. Principles of Plasma Discharges and Materials Processing. New Jersey: Wiley-Interscience; 2005
  129. 129. Ostrikov K. Plasma Nanoscience. New York: Wiley; 2008
  130. 130. Ostrikov K, Cvelbar U, Murphy AB. Plasma nanoscience: Setting directions, tackling grand challenges. Journal of Physics D: Applied Physics. 2011;44(17):174001. DOI: 10.1088/0022-3727/44/17/174001
  131. 131. Ostikov K, Neyts EC, Meyyappan M. Plasma nanoscience: From nano-solids in plasmas to nano-plasmas in solids. Advances in Physics. 2013;02(02):113-0224. DOI: 10.1080/00018732.2013.808047
  132. 132. Montaser A, Golightly DW, editors. Inductively Coupled Plasmas in Analytical Atomic Spectrometry. New York: VCH; 1992
  133. 133. Schulze A, Jeong JY, Babayan SE, Park J, Selwyn GS, Hicks RF. The atmospheric-pressure plasma jet. IEEE Transactions on Plasma Science. 1998;16(6):1685-1694. DOI: 10.1109/27.747887
  134. 134. Eden JG, Park S-J. New opportunities for plasma science in non-equilibrium, low-temperature plasmas confined to microcavities: There’s plenty of room at the bottom. Physics of Plasmas. 2006;13:057101. DOI: 10.1063/1.2179413
  135. 135. Shin JJ, Kong MG. Evolution of discharge structure in capacitive radio-frequency atmospheric microplasmas. Physical Review Letters. 2006;96:105009. DOI: 10.1103/PhysRevLett.96.105009
  136. 136. Wagner AJ, Mariotti D, Yurchenko KJ, Das KT. Experimental study of a planar atmospheric-pressure plasma operating in the microplasma regime. Physical Review E. 2009;80:065401R. DOI: 10.1103/PysRevE80.065401
  137. 137. Lu X, Wu S, Gou J, Pan Y. An atmospheric-pressure, high-aspect ratio, cold micro-plasma. Scientific Reports. 2014;4(7488). DOI: 10.1038/srep07488
  138. 138. Karanassios V. Microplasmas for chemical analysis: Analytical tools or research toys? Spectrochimica Acta Part B. 2004;59:909-928. DOI: 10.1016/j.sab.2004.04.005
  139. 139. Weagant S, Smith AT, Karanassios V. Mobile micro- and nano-instruments: Fast, cheap and under wireless control. ECS Transactions. 2010;28(14):1-6. DOI: 10.1149/1.3490180
  140. 140. Badiei HR, Lai B, Karanassios V. Micro- and nano-volume samples by electrothermal, near-torch vaporization (NTV) using removable, interchangeable and portable rhenium coiled-filament assemblies and ICP-AES. Spectrochimica Acta Part B. 2012;77:19-30. DOI: 10.1016/j.sab.2012.07.025
  141. 141. Badiei HR, McEnaney J, Karanassios V. Bringing part of the lab to the field: On-site Cr speciation by electrodeposition of Cr(III)/Cr(VI) on portable coiled-filaments and measurement in a lab by NTV-ICP-AES. Spectrochimica Acta Part B. 2012;78:42-49. DOI: 10.1016/j.sab.2012.10.002
  142. 142. Badiei HR, Liu C, Karanassios V. Taking part of the lab to the sample: On-site electrodeposition of Pb followed by measurement in a lab using electrothermal, near torch vaporization (NTV) sample introduction and ICP-AES. Microchemical Journal. 2013;108:131-136. DOI: 10.1016/j.microc.2012.10.013
  143. 143. Karanassios V, Johnson K, Smith AT. Micromachined, planar-geometry, atmospheric-pressure, battery-operated microplasma devices (MPDs) on chips for liquids, gases or solids by optical emission spectrometry. Analytical and Bioanalytical Chemistry. 2007;388:1595-1604. DOI: 10.1007/s00216-007-1273-4
  144. 144. Weagent S, Karanassios V. Helium-hydrogen microplasma device (MPD) on postage-stamp-size plastic-quartz chips. Analytical and Bioanalytical Chemistry. 2009;395:577-589. DOI: 10.1007/s00216-009-2942-2
  145. 145. Weagant S, Chen V, Karanassios V. Battery-operated, Ar-H2 microplasma on hybrid, postage stamp-size plastic-quartz chips for elemental analysis using a portable optical emission spectrometer. Analytical and Bioanalytical Chemistry. 2011;401:2865-2880. DOI: 10.1007/s00216-011-5372-x
  146. 146. Weagant S, Karanassios V. Battery-operated, planar-geometry microplasma on a postage-stamp size chips: Some fundamentals. Proceedings of SPIE. 2011;8024:80240L. DOI: 10.1117/12.884329
  147. 147. Zhang X, Karanassios V. Rapid prototyping of solar-powered, battery-operated, atmospheric-pressure, sugar-cube size microplasma on hybrid, 3D chips using a portable optical emission spectrometer. Proceedings of SPIE. 2012;8366:83660D. DOI: 10.1117/12.919550
  148. 148. Abbaszadeh S, Karim KS, Karanassios V. Measurement of UV from a microplasma by a microfabricated amorphous selenium detector. IEEE Transactions on Electronic Devices. 2013;60(2):880-883. DOI: 10.1109/TED.2012.2231682
  149. 149. Devathasan D, Trebych K, Karanassios V. 3D-printed, sugar cube-size microplasma on a hybrid chip used as a spectral lamp to characterize UV-vis transmission of polycarbonate chips for microfluidic applications. Proceedings of SPIE. 2013;8718:87180B. DOI: 10.1117/12.2016222
  150. 150. Nguon O, Gauthier M, Karanassios V. Determination of the loading and stability of Pd in an arborescent copolymer in ethanol by microplasma-optical emission spectrometry. RSC Advances. 2014;4:8978-8984. DOI: 10.1039/C3RA46232C
  151. 151. Shatford R, Karanassios V. Microplasma fabrication: From semiconductor technology for 2D-chips and microfluidic channels to rapid prototyping to 3D-printing of microplasma devices. Proceedings of SPIE. 2014;9106:9106H1-9106H7. DOI: 10.11171/12.2050538
  152. 152. Nguon O, Huang S, Gauthier M, Karanassios V. Microplasmas: From applications to fundamentals. Proceedings of SPIE. 2014;9105:9101061-9101067. DOI: 10.1117/12.2050348
  153. 153. Weagent S, Dulai G, Li L, Karanassios V. Characterization of rapidly-prototyped, battery-operated, Ar-H2 microplasma on a chip for elemental analysis of microsamples by portable optical emission spectrometry. Spectrochimica Acta Part B. 2015;106:75-80. DOI: 10.1016/j.sab.2015.01.009
  154. 154. Shatford R, Kim D, Karanassios V. Microfluidics for spectrochemical applications. Proceedings of SPIE. 2015;9486:94860N1-94960N6. DOI: 10.1117/12.2177513
  155. 155. Yang R, Sazonov A, Karanassios V. Flexible, self-powered, visible-light detector characterized using a battery-operated, 3D-printed microplasma operated as a light source. 2016 IEEE Sensors. 2016:997-999. ID 978-1-4799-8287-5/16. DOI: 10.1109/ICSENS.2016.7808738
  156. 156. So H, Cebula DA, Karanassios V. Towards chromium speciation in lake-waters by microplasma optical emission spectrometry. Proceedings of SPIE. 2017;10215:102150I. DOI: 10.1117/12.2262955
  157. 157. Feynman RP. There is plenty of room at the bottom. Engineering and Science at Caltech. 1960;13(5):22-36. http://calteches.library.caltech.edu/47/3/ES.23.5.1960.0.pdf
  158. 158. NSF “Partnership in Nanotechnology”, Program announcement, societal implications of nanoscience and nanotechnology. Arlington, VA (1997), http://www.nsf.gov/nano (Accessed Dec. 2017)
  159. 159. Gates BD, Xu Q, Stewart M, Ryan D, Willson CG, Whitesides GM. New approaches to nanofabrication. Chemical Reviews. 2005;105(4):1171-1196. DOI: 10.1021/cr030076o
  160. 160. Eijkel JC, van der Berg A. Nanofluidics: What is it and what can we expect from it?. Microfluid Nanofluidics. 2005;1:249-267. DOI: 10.1007/s10404-004-0012-9
  161. 161. Tsenga AA, Notargiacomo A. Nanofabrication by scanning probe microscope lithography: A review. Journal of Vacuum Science and Technology B. 2005;23:877. DOI: 10.1116/1.1926293
  162. 162. Mijatovic D, Eijek JCT, Van der Tert VA. Technologies for nanofluidic systems: Top-down vs. bottom-up: A review. Lab on a Chip. 2005;5:402-500. DOI: 10.1039/b416951d
  163. 163. Huh D, Mills KL, Zhu X, Burns MA, Thouless MD, Takayama S. Tuneable elastomeric nanochannels. Nature Materials. 2007;6:424-428. DOI: 10.1038/nmat1907
  164. 164. Li M, Tang HX, Roukes ML. Ultra-sensitive NEMS-based cantilevers for sensing, scanned probe. Nature Nanotechnology. 2007;2:114-120. DOI: 10.1038/nnano.2006.208
  165. 165. He Q, Chen S, Su Y, Fang Q, Chen H. Fabrication of 1D nanofluidic channels on glass substrate by wet etching. Analytica Chimica Acta. 2008;628:1-8. DOI: 10.1016/j.aca.2008.08.040
  166. 166. Prakash S, Piruska A, Gatimu EN, Bohn PW, Sweedler JV, Shannon MA. Nanofluidics: Systems and applications. IEEE Sensors Journal. 2008;8(5):441-450. DOI: 10.1109/JSEN.2008.918758
  167. 167. Abgrall P, Nguyen NT. Nanofluidic devices and their applications. Analytical Chemistry. 2008;80(7):2326-2341. DOI: 10.1021/ac702296u
  168. 168. Sparreboom W. van den Berg A, Eijkel JCT. Principles and applications of nanofluidic transport. Nature Nanotechnology. 2009;4:713-720. DOI: 10.1038/nnano.2009.332
  169. 169. Piruska A, Gong M, Sweedler JV, Boha PW. Nanofluidics in chemical analysis. Chemical Society Reviews. 2010;39:1060-1072. DOI: 10.1039/b900409m
  170. 170. Roco MC. The long view of nanotechnology development: The National Nanotechnology Initiative at 10 years. Journal of Nanoparticle Research. 2011;13:427-445. DOI: 10.1007/s11051-010-0192-z
  171. 171. Menard LD, Ramsey JM. Fabrication of sub-5 nm nanochannels in insulating substrates using focused ion beam milling. Nano Letters. 2011;11(2):512-517. DOI: 10.1021/nl103369g
  172. 172. Chantiwas R, Park S, Soper SA, Kim BC, Takayama S, Sunkara V, Hwangc H, Cho Y-K. Flexible fabrication and applications of polymer nanochannels and nanoslits. Chemical Society Reviews. 2011;40:3677-3702. DOI: 10.1039/C0CS00138D
  173. 173. Poot M, van der Zant HSJ. Mechanical systems in the quantum regime. Physics Reports. 2012;511(5):273-335. DOI: 10.1016/j.physrep.2011.12.004
  174. 174. Hanay MS, Kelber S, Naik AK, Chi D, Hentz S, Bullard EC, Colinet E, Duraffourg L, Roukes ML. Single-protein nanomechanical mass spectrometer in real time. Nature Nanotechnology. 2012;7:602-608. DOI: 10.1038/nnano.2012.119
  175. 175. Duan C, Wang W, Xin Q. Fabrication of nanofluidic devices. Biomicrofluidics. 2013;7. DOI: 026501, 10.1063/1.4794973
  176. 176. Schmid S, Kurek M, Adolphsen JQ, Boisen A. Real-time single airborne nanoparticle detection with nanomechanical resonant filter-fiber. Scientific Reports. 2013;3(1288). DOI: 10.1038/srep01288
  177. 177. Uranga A, Verd J, Marigó E, Giner J, Muñóz-Gamarra JL, Barniol N. Exploitation of non-linearities in CMOS-NEMS. Sensors Actuators A. 2013;197:88-95. DOI: 10.1016/j.sna.2013.03.032
  178. 178. Muñoz-Gamarra JL, Alcaine P, Marigó E, Giner J, Uranga A, Esteve J, Barniol N. Integration of NEMS resonators in CMOS. Microelectronics Engineering. 2013;110:246-249. DOI: 10.1016/j.mee.2013.01.038
  179. 179. Berman D, Krim J. Surface science, MEMS and NEMS. Progress in Surface Science. 2013;88(2):171-211. DOI: 10.1016/j.progsurf.2013.03.001
  180. 180. Bocquet L, Tabeling P. Physics and technological aspects of nanofluidics. Lab on a Chip. 2014;14:3143-3158. DOI: 10.1039/c4lc00325j
  181. 181. Mawatari K, Kazoe Y, Shimizu H, Pihosh Y, Kitamori T. Extended-nanofluidics. Analytical Chemistry. 2014;86(9):4068-4077. DOI: 10.1021/ac4026303
  182. 182. Sage E, Brenac A, Alava T, Morel R, Dupré C, Hanay MS, Roukes ML, Duraffourg L, Masselon C, Hentz S. Neutral particle mass spectrometry with nanomechanical systems. Nature Communications. 2015;6(6482). DOI: 10.1038/ncomms7482
  183. 183. Hui Y, Nan T, Sun NX, Rinaldi M. High resolution magnetometer based on a MEMS-CMOS oscillator. Journal of Microelectromechanical Systems. 2015;24(1):134-143. DOI: 10.1109/JMEMS.2014.2322012
  184. 184. Schmid S, Villanueva LG, Roukes ML. Fundamentals of Nanomechanical Resonators. New York: Springer; 2016
  185. 185. Kandemir AC, Erdem D, Ma H, Reiser A, Spolenak R. Polymer nanocomposite patterning by dip-pen nanolithography. Nanotechnology. 2016;27(13):135303. DOI: 10.1088/0957-4484/27/13/135303
  186. 186. Hou X, Guoa W, Jiang L. Biomimetic smart nanopores and nanochannels. Chemical Society Reviews. 2011;40:2385-2401. DOI: 10.1039/C0CS00053A
  187. 187. Schneider GF, Dekker C. DNA sequencing with nanopores. Nature Biotechnology. 2012;30:326-328. DOI: 10.1038/nbt.2181
  188. 188. Ayub M, Hardwick SW, Luisi BF, Bayley H. Nanopore-based identification of individual nucleotides for direct RNA sequencing. Nano Letters. 2013;13(12):6144-6150. DOI: 10.1021/nl403469r
  189. 189. Miles BN, Ivanov AP, Wilson KA, Doğan F, Japrung D, Edel JB. Single molecule sensing with solid-state nanopores. Chemical Society Reviews. 2013;42:15-28. DOI: 10.1039/C2CS35286A
  190. 190. Taniguchi M. Selective multidetection using nanopores. Analytical Chemistry. 2015;87(1):188-199. DOI: 10.1021/ac504186m
  191. 191. Deng T, Wang Y, Chen Q, Chen H, Liu Z. Massive fabrication of silicon nanopore arrays with tunable shapes. Applied Surface Science. 2016;390:681-688. DOI: doi.org/10.1016/j.apsusc.2016.07.171
  192. 192. Taniguchi M. Single-molecule sequencing. In: Kiguchi M, editor. Single-Molecule Electronics. Singapore: Springer; 2016. DOI: 10.1007/978-981-10-0724-8_9
  193. 193. Sheikholeslami M. Influence of coulomb forces on Fe3O4-H2O nanofluid thermal improvement. International Journal of Hydrogen Energy. 2017;42:821-829. DOI: 10.1016/j.ijhydene.2016.09.185
  194. 194. Sheikholeslami M, Seyednezhad M. Simulation of nanofluid flow in a porous media. International Journal of Heat and Mass Transfer. 2018;120:772-781. DOI: 10.1016/j.ijheatmasstransfer.2017.12.087
  195. 195. Hagleitner C, Hierlemann A, Lange D, Kummer A, Kerness N, Brand O, Baltes H. Smart single-chip gas sensor microsystem. Nature. November 15, 2001;414:293-296. DOI: 10.1038/35104535
  196. 196. Hierlemann A, Baltes H. CMOS-based chemical microsensors. Analyst. 2003;128:15-28. DOI: 10.1039/B208563C
  197. 197. Graf M, Barrettino D, Baltes H, Hierlemann A. Microtechnology and MEMS series. In: CMOS Hotplate Chemical Microsensors. Germany: Springer; 2007. DOI: 10.1007/978-3-540-69562-2
  198. 198. Lee D, Dulai G, Karanassios V. Survey of energy harvesting and energy scavenging for powering wireless networks. Proceedings of SPIE. 2013;8028:8720S1. DOI: 10.1117/12.2016238
  199. 199. Yang R, Sazonov A, Karanassios V. Flexible, self-powered, visible-light detector characterized using a microplasma. Proceedings of the IEEE Sensors. 2016;978:997-999. DOI: 10.1109/ICSENS.2016.7808738
  200. 200. Trizcinski P, Nathan A, Karanassios V. Approaches to energy harvesting and energy scavenging for energy autonomy. Proceedings of SPIE. 2017;10194:10194A1-10194A9. DOI: 10.1117/12.2262957
  201. 201. Li Z, Karanassios V. Artificial Neural Networks (ANNs) for Spectral Interference Correction Using a Large-Size Spectrometer and ANN-Based Deep Learning for a Miniature one. London, UK: InTech Publishing; 2018. Chapter 12. pp. 227-249. DOI: 10.5772/intechopen.71039
  202. 202. Trzcinski P, Weagent S, Karanassios V. Wireless data acquisition of transient signals for mobile spectrometry applications. Applied Spectroscopy. 2016;70:905-915. DOI: 10.1177/0003702816638304
  203. 203. Trzcinski P, Karanassios V. How can wireless, mobile data acquisition be used for taking part of the lab to the sample, and how can it join the internet of things?. Proceedings of SPIE. 2016;9855:985503. DOI: 10.1117/12.2224400

Written By

Vassili Karanassios

Submitted: 17 October 2017 Reviewed: 25 January 2018 Published: 22 August 2018