Fluid Flow in Carbon Nanotubes and Nanopipes

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

REVIEW ARTICLE

Fluid flow in carbon nanotubes


and nanopipes
Nanoscale carbon tubes and pipes can be readily fabricated using self-assembly techniques and
they have useful electrical, optical and mechanical properties. The transport of liquids along their
central pores is now of considerable interest both for testing classical theories of fluid flow at the
nanoscale and for potential nanofluidic device applications. In this review we consider evidence for
novel fluid flow in carbon nanotubes and pipes that approaches frictionless transport. Methods for
controlling such flow and for creating functional device architectures are described and possible
applications are discussed.

M. WHITBY AND N. QUIRKE arrays when assembled as nanoporous membranes, are discussed
in later sections.
Chemistry Department, Imperial College, South Kensington, Our understanding of the static properties of liquids in
London SW7 2AZ, UK. macroscopic capillaries is based on the nineteenth century work
e-mail: n.quirke@imperial.ac.uk of Laplace, Poisson and Young2. Steady-state flow of simple
incompressible fluids in a channel width 2h, driven for example
Understanding and controlling the flow of liquids at the nanoscale by gravity ρg or a pressure gradient dP/dy, can be described by
is currently a subject of great interest. More than 100 papers the Navier Stokes equation3. The solution for the velocity in the
describing work in nanofluidics have been published in the past direction of flow, y, as a function of the distance from the wall,
five years. Recent experimental measurement of fluid transport z, has a parabolic profile (Fig. 2) given by:
through channels with widths of a few nanometres has confirmed
predictions derived from computer simulations that such flow Uy(z) = (ρg/2η)[(δ + h)2 – z2]
occurs at a much greater rate than expected. This discovery has
significant implications, both for our understanding of how fluids where η is the viscosity and δ, the slip length, is the distance into
behave at very small length scales and for the design of nanofluidic the wall at which the velocity extrapolates to zero. Conventionally,
devices. Here we review selected papers with an emphasis on fluid the slip length is assumed to be zero. However, if it is not then
flow through the central pores of carbon tubes. we are said to have slip boundary conditions at the wall, which
The practical investigation of fluid flow in nanoscale channels has for very small channels can significantly enhance the fluid flow.
been facilitated by the availability of tubular carbon structures with Integrating the velocity profile over the cross-sectional area of the
open central pore diameters in the range of one to several hundred pipe gives the Hagen-Poiseuille law for the flux through a pipe.
nanometres. Carbon offers a number of attractive features for the This law states that the flow rate of a fluid passing through a tube is
fabrication of nanofluidic devices including: the existence of several directly proportional to the pressure difference between the tube
well studied processes for self-assembly of tubular structures; useful ends and to the fourth power of the tube’s internal radius. The
electronic properties for sensing and signalling; biocompatibility, flow rate is inversely proportional to the tube length and to the
ready chemical modification to allow functionalization; and low- viscosity of the fluid.
friction transport thanks to graphitic surfaces. The dynamics of capillary filling were elucidated by Washburn4,5
Two types of nanoscale carbon tubes have been used using the Hagen-Poiseuille law with the driving force for flow given
experimentally, each with distinct properties and methods by the Laplace equation for the pressure difference across the
of synthesis. The first are carbon nanotubes, both single and invading liquid meniscus. Thus the penetration length L at time t of
multiwalled, with the fullerene molecular structure reported by a fluid in a capillary of radius R is given by:
Ijima in 1991 (ref. 1). The second are carbon nanopipes produced
by chemical vapour deposition (CVD) of amorphous carbon in L2 = (Rγ/2η)t
alumina templates with a honeycomb pore morphology (Fig. 1).
Throughout this review we reserve the term ‘nanotube’ to refer with γ being the liquid/vapour surface tension and η the shear
to true molecular carbon structures. The term ‘nanopipe’ will viscosity of the invading liquid. In deriving the Washburn equation
be used to describe carbon pipes produced in templates using it is assumed that there is no fluid motion at the wall (that is, stick
CVD. The latter are generally larger and composed mainly of boundary conditions). The Washburn equation is very successful
amorphous rather than well ordered graphitic carbon. The on time scales sufficient to establish viscous flow but breaks
properties and fabrication of both types of carbon tube, and their down for short times. A more general solution can be found for

nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology 87


REVIEW ARTICLE

b a

20 μm

1 μm

2 μm

500 nm

Figure 1 Scanning electron microscope images of carbon nanopipes produced by the authors using standard chemical vapour deposition. a, Nanopipes partially
released from an anodic aluminium oxide template following sonication in NaOH. b, Cross section of intact carbon coated membrane. c, Higher magnification view of
individual aligned carbon pipes. d, Surface of carbon membrane showing open pores (diameter ~160 nm). The growth procedures are described in the section on “Flow
through nanoscale channels including carbon nanopipes”.

non-steady viscous incompressible flow (the Bosanquet equation6), that the smaller the value of α, the greater the slip length, a direct
which approaches the Washburn equation at long times. relationship between the two only exists at low density10.
For nanoscale capillaries the flow behaviour is dominated not
by bulk properties (γ, η) but by the interaction of the fluid with SIMULATION AND THEORY OF NANOSCALE FLOW
the capillary walls. In this regard it is important to include the
role of surface friction due to molecular corrugation, especially A central question in nanofluidics concerns the extent to which
for graphitic surfaces where the high surface density of atoms the classical equations describing the way fluids interact with
leads to a very low corrugation. A convenient parameter here is nanomaterials hold at the nanoscale. Molecular simulation is
the Maxwell coefficient7 α, which represents the fraction of the ideally suited to shed light on this problem and has shown that
molecular collisions with the capillary walls that undergo diffuse static properties such as surface tensions and contact angles
scattering and/or trapping–desorption, the remainder being of simple fluids obey classical relations down to almost single
specularly reflected (no energy loss). The Maxwell coefficient nanometre dimensions11. Steady-state Poiseuille flows maintain a
has been shown to provide a useful description of molecular parabolic velocity profile down to widths of several nanometres12.
friction8 for a variety of materials9. However, although it is clear Deviations from this behaviour can typically be linked to density
inhomogeneities in the confining walls.
In performing simulations relevant to experimental work on
+h
flow in nanopores, a major challenge has been to choose realistic
potentials between the molecules in the fluid and in the pore walls.
Although it has proved possible to be reasonably confident about,
Uy(z) say, determining alkane/carbon potentials by fitting to heats of
z adsorption, for fluids such as water considerable uncertainty has
arisen. For example, reports of the measured contact angle of water on
y
a graphitic surface differ by as much as a factor of two13. The choice of
intermolecular potential determines whether a model nanotube can
be filled with water14,15 or not. However recent work has shown how it
may be possible to calibrate solvent–nanotube potentials using Raman
shifts in solution16. The uncertainty in water potentials notwithstanding,
some general features of nanoscale flow have been determined from
molecular dynamics using simple fluids relevant to the experimental
work. For example Sokhan et al.8,17 were the first to show that fluids
–h flowing through (slit) carbon nanopores experience very low surface
friction, that is very small Maxwell coefficients (~0.01) or equivalently
large slip boundary conditions, whereas pores made from other
Figure 2 Parabolic velocity profile Uy(z) for Poiseuille flow in a capillary of width 2h. materials show higher values and stick boundary conditions.

88 nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology


REVIEW ARTICLE
900
Time = 60 ps
800
700

600

V m s–1
500
400
300
200
100

0
0 2 4 6 8 10 12
log t (ps)

Figure 3 Imbibition of decane in a (7,7) carbon nanotube (diameter = 0.951 nm). Left: Simulated imbibition, showing a snapshot of the fluid after 60 ps50. Right: plot of the
imbibition velocity obtained from the analytic model using parameters derived from nanoscale simulation, the upper (blue) line shows results for a carbon nanotube, the
bottom (black) line, a non-carbon tube50.

Subsequent theoretical work shows that in a cylindrical geometry, lithographically. The height of the channels varied from 2.7 μm down
decreasing the radius can reduce the Maxwell coefficients by orders to 40 nm. The results show a marked and progressive departure from
of magnitude compared with a slit (~0.000002 for decane in a Poiseuille flow as the channel height falls below 200 nm. This effect
(7,7) nanotube, diameter 0.75 nm) and that carbon nanotubes are was observed most strongly for hexadecane and is also evident for
predicted to be essentially frictionless pipes18,19. Skoulidas et al.20 decane and for silicone oil. Slip lengths were estimated at 25 to 30 nm,
compared transport diffusivities for carbon nanotubes with zeolites ~9 nm and ~14 nm respectively. The effect was not seen for water at
using molecular dynamics and found exceptionally high transport the length scales investigated.
rates for the nanotubes due to the smooth carbon surface. We Several groups have recently reported results from experiments
should expect therefore that crystalline carbon channels (as opposed involving flow of liquids and gases through membranes with
perhaps to amorphous surfaces) show significantly increased flow. nanoscale pores composed of carbon. Two principal methods have
Another consequence is the extremely rapid imbibition of wetting been used to construct suitable membranes. The first involves CVD
fluids (decane is predicted to fill (7,7) nanotubes, at 800 m s–1; ref. 18) of carbon in an anodic aluminium oxide (AAO) template. The second
observed in non-equilibrium molecular dynamics. Note that the is based on production of polymer composites containing an aligned
same model predicts external wetting to be much slower by factors array of carbon nanotubes spanning the thickness of the membrane.
of roughly 30 for a (13,13) tube with diameter 1.765 nm. This has Both techniques allow liquid flow through the central pores of carbon
implications for nanotube arrays, where flow may occur along either nanotubes and nanopipes to be studied without the complication of
the external or internal surfaces of the tubes. external wetting.
From a Bosanquet-like equation (where we approximate the The first method for creating carbon nanomembranes takes
surface friction with an expression involving the Maxwell coefficient advantage of the self-assembly of highly-ordered nanopores arranged
and the surface tension and the driving force is represented by the in a dense hexagonal pattern in alumina when aluminium foil is
wall-wetting liquid tension), it is possible to obtain a solution for the etched in acid23,24. Using this approach, monodisperse-sized pores with
imbibition velocity and penetration length L at all times18. There is an densities as high as 1011 cm–2 are possible. To line the nanopores, the
L ∝ t dependence for short times tending to t1/2 at long times (as in the AAO template is placed in a furnace and a hydrocarbon gas is flowed
Washburn equation), the timescale for which depends on the value for several hours, whereupon carbon is deposited on the hot surface of
of the Maxwell coefficient. Figure 3 shows how the imbibition speed the alumina25. Carbon pipes with an outside diameter as small as 10 nm
for decane varies as a function of time with ultrafast uptake for t<μs have been formed in this way although larger 100 to 200 nm nanopipes
falling to cm s–1 at longer times. are more usual. The length of these pipes is typically 50 to 100 μm.
The situation is more complex for hydrogen bonding fluids such as It is important to appreciate that, as synthesized, the typically
water. In very narrow tubes the water molecules form a one-dimensional 20-nm-thick walls of carbon pipes produced by the AAO templating
chain with linking hydrogen bonds. At equilibrium, conduction method will be at least partly amorphous. Che et al.26 reported electron
through such tubes occurs in bursts14 and this behaviour is attributed diffraction studies in a transmission electron microscope (TEM),
to density fluctuations at the openings of the tube. which produced a pattern characteristic of non-aligned graphite
fragments. There are reports that high-temperature annealing can
FLOW THROUGH NANOSCALE CHANNELS INCLUDING CARBON NANOPIPES improve the order of the tube walls through graphitization27,28,
which may enhance transport properties. Miller et al.25 confirmed
Initial indications of rapid fluid flow through very small channels electro-osmotic flow through these carbon nanomembranes and
in an experimental system were published by Pfahler et al. in reported a linear relationship between electro-osmotic flow velocity
1990 (ref. 21). Working with rectangular channels etched in a and applied current density.
silicon substrate, they found an n-propanol flux approximately
three times greater than expected in the smallest channels (height FLUID FLOW THROUGH CARBON NANOTUBES
0.8 μm and width 100 μm). Following up on this observation in 2002,
Cheng and Gordanio22 reported measurements for Newtonian Hinds et al.29 published the first account of experiments involving fluid
fluids flowing under pressure through nanoscale channels produced flow through a membrane composed of true molecular nanotubes. This

nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology 89


REVIEW ARTICLE
a closed by end caps or plugged by residual iron catalyst. To overcome
these problems, a polystyrene solution was spin coated onto the array
to fill the gaps. Then plasma etching was used to remove the top layer
of the composite and to open the nanotube central pores. The pore
density was estimated to be 6 (±3) × 1010 cm–2.
Using gas flow apparatus, the MWNT composite array showed
a linear relationship between flow rate and pressure drop across
the membrane. The authors also studied diffusion of aqueous ionic
species (Ru(NH3)63+ diameter = 0.55 nm) and noted that the enhanced
diffusion coefficient was near the bulk aqueous-solution diffusion for
the Ru cation. They infer only limited interaction between the ion and
the nanotube tip and walls.
Majumder et al.30 published the results of further experiments by
the same group on fluid transport through the MWNT composite
50 μm membranes. A simple pressure-driven flow apparatus was used, with
the accumulated mass of transported fluids including water, ethanol
b and alkanes being measured after a fixed time period. Their results
demonstrate dramatically enhanced flow through ~7-nm-diameter
nanotube cores compared with conventional fluid flow theory.
Observed flow rates were four to five orders of magnitude greater
than predicted by hydrodynamics based on macroscale behaviour
(Table 1). The authors conclude that the implied slip lengths
(3 to 70 μm), which are much greater than the tube diameter, are
consistent with a nearly frictionless interface.
In the studies reviewed so far, increasingly interesting fluid flow
results were found at progressively smaller length scales. In 2006,
Holt et al.31 achieved a milestone in measuring water flow through
the central pores of double walled carbon nanotubes (DWNT). These
tubes had inner diameters less than 2 nm with defect-free graphitic
50 μm
walls that are expected from simulation (see previous section) to
present a low friction surface to transported fluids.
The practical challenges involved in studying transport through
very small carbon nanotubes are formidable. The approach used
c by Holt et al. was to use fabrication techniques adapted from the
semiconductor industry. A dense array of carbon nanotubes were first
grown on a silicon substrate onto which metal catalyst particles had
been deposited. The spaces between the tubes were then completely
filled with silicon nitride from low-pressure chemical vapour. Finally
the closed ends of the tubes and catalyst particles were removed by
a series of etching and ion milling steps. The resulting membranes
had a thickness in the range of 2.0 to 3.0 μm and pore densities
≤0.25 × 1012 cm–2. Pore diameters, calibrated by passage of gold
nanoparticles, were in the range of 1.3 to 2.0 nm.
Gas flow and water flow measurements were taken in flow cells
sealed with o-rings. Five hydrocarbon and eight non-hydrocarbon
Figure 4 Carbon nanotube arrays and membranes. a, An as-grown, dense, multiwalled gases were tested to determine flow rates and to demonstrate
carbon nanotube array produced with an Fe-catalysed chemical vapour deposition molecular weight selectivity compared with helium. Water flow was
process. b, The cleaved edge of the nanotube-polystyrene membrane after exposure pressure-driven at 0.82 atm and measured by following the level of
to H2O plasma oxidation. The polystyrene matrix is slightly removed to contrast the the meniscus in a feed tube. The results for both gas and liquid show
alignment of the nanotubes across the membrane. c, Schematic of the target membrane dramatic enhancements over flux rates predicted with continuum flow
structure. With a polymer embedded between the nanotubes, a viable membrane models. Gas flow rates were between 16 and 120 times that expected
structure can be readily produced, with the pore being the rigid inner-tube diameter of according to the Knudsen diffusion model in which fluid molecule–
the nanotube. Reproduced with permission from ref. 29. Copyright (2004) AAAS. wall collisions dominate the flow. Water flow rates were 560 to 8,400
times greater than those calculated according to the Hagen–Poiseuille
equation. Minimum slip lengths are estimated in the range of 140 to
1,400 nm. To express these findings in a practical context, the nanotube
group used dense, well aligned arrays of multiwalled carbon nanotubes membranes showed flow rates several orders of magnitude greater
(MWNT) grown using CVD on a flat quartz substrate (Fig. 4). One than those of conventional membranes, despite having pore sizes an
advantage of this approach is that the pore diameters are much smaller order of magnitude smaller. The authors of these studies point out the
(4.3 ± 2.3 nm) than with the AAO template method. Another is that significance of this finding for separation applications.
the crystallinity of the inner carbon nanotube walls is greater, with
consequent benefits for low friction fluid flow. MICROSCOPE STUDIES OF FLUID FLOW IN NANOTUBES
As synthesized MWNT arrays are permeable not only through the
central pores of the nanotubes, but also extensively through the gaps Permeation experiments allow quantitative measurement of
between the individual tubes. Furthermore, many nanotubes will be fluid transport through nanoscale carbon pores. A more qualitative

90 nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology


REVIEW ARTICLE
approach is to observe fluid interaction using optical and electron Miller and Martin35 describe a more sophisticated redox
microscopes. It should be noted that the structures studied are modulation of ionic transport, which they used to switch both the
generally large (tube diameter > 200 nm). The results are therefore rate and direction of electro-osmotic flow. Working again with
unlikely to show the more unusual surface-dominated flow carbon nanopipes, they coated the carbon surfaces with the redox
behaviour expected in true nanoscale channels. polymer poly(vinylferrocene) (PVFc). A variable potential was used
In 2004, Rossi et al.32 reported a series of observations of liquid/ to totally oxidize, totally reduce or to set the redox state of the PVFc
carbon-pipe interactions made in an environmental scanning to an intermediate value. Once set, the membrane would remember
electron microscope (ESEM). The advantage of using an ESEM its state until altered. The authors compare their approach to the field-
is that in situ experiments involving fluids can be undertaken. effect concept in semiconductors.
Furthermore, by carefully varying the pressure in the sample Similar findings are reported by Majumder et al.37 using their
chamber and by controlling the temperature using a Peltier cooling finer MWNT composite membranes (described earlier). Instead
stage, it is possible to cause condensation and evaporation to occur of functionalizing the entire length of the nanopores, the authors
at will in the sample. Using this method, the team captured the selectively modified just the tip region of the nanotubes. The tips were
behaviour of liquid menisci inside 200 to 300 nm diameter CVD exposed by plasma etching and functionalized using a carbodiimide
grown carbon pipes (Fig. 5). mediated coupling. They showed that different chemical end groups
A key result was that water condensed on the internal walls of attached near the entrance to the channels could selectively modulate
the carbon tubes in preference to the external steel surface of the ionic flux.
stage. The authors concluded that the disordered walls of the AAO Babu et al.38 report a technique for guiding the entry of water
template-grown tubes are hydrophilic and they measured contact into carbon nanopipes which were selectively modified by the
angles with water between 5 and 20°. They also reported a slight electrodeposition of polypyrrole on their tips. The authors conclude
deformation of tubes containing plugs of water, indicating negative that this may be useful for constructing devices capable of controlling
pressure inside the channel due to tensile capillary forces. Carbon liquid flows at the nanoscale.
tubes of this type may be useful for guiding aqueous fluids to and Molecular dynamics simulations of water in the pores of single-
from specific locations at the nanoscale and for collecting attolitre walled carbon nanotubes39 show that an electric field can also modulate
and picolitre samples. filling through dipole interactions. This effect might also be exploited
Also in 2004, Kim et al.33 reported work on liquid interactions to achieve selective control of liquid flow at the nanoscale and may
with similar carbon pipes and plotted filling length against time. play a part in natural biological pore transport mechanisms.
Working with ≥200 nm nanopipes, they found a good fit with
x(t) = At1/2 (the Washburn equation) as expected for the long (ms) CHALLENGES IN BUILDING NANOFLUIDIC DEVICES
observation times.
A year later, the same group published a follow-on paper34 Many of the papers reviewed above describe novel phenomena and
reporting on imbibition experiments that involved introducing methods that will be directly applicable to the design and fabrication
fluorescent nanoparticles (diameter ~50 nm) into a variety of of nanofluidic devices with practical application in medical
liquids. The suspended particles were taken up by carbon pipes diagnosis, sensing and materials processing. Before these ambitions
with 500-nm diameters through a combination of capillary action can be achieved, a considerable number of practical hurdles will
and evaporation forces. The introduced particles could be observed need to be overcome. In this section some approaches are considered
moving along the tubes as their fluorescence was visible through that may have useful architectural or procedural application in the
the tube walls. The authors speculate that this technique may be implementation of a viable nanofluidics technology.
useful for observing the interactions of biological macromolecules Microfluidics is a rapidly advancing field with emerging
in the vacuum environment of an electron microscope. analytical and synthetic applications40. Cao et al.41 report
one of the few experimental investigations of the problem of
CONTROLLING LIQUID FLOW THROUGH NANOTUBES microfluidic to nanofluidic interfacing. Liquid samples for
characterization or processing using nanoscale components are
In the previously discussed study by Miller et al.25, electrochemical likely to start out as macroscale droplets. These must be guided
derivitization was used to change the surface functionalization into progressively smaller channels for ultimate delivery to the
of carbon nanopipes. Using this technique, both the magnitude nanoscale device components. A number of potential problems
and the direction of electro-osmotic flow could be modified. arise. One is channel blocking due to large macromolecules
Simple application of a potential across the membrane can also or insoluble debris. This can already be a limiting hazard with
act as an immediate transport switch as reported by several microscale lab-on-a-chip systems. Work is currently taking
researchers29,35,36. place to understand the dynamics of particulate transport at the

Table 1 Comparison of observed flow velocity with predictions for pressure-driven flow through aligned MWNT membranes30. These results indicate liquid flow
rates four to five orders of magnitude faster than would be predicted from conventional fluid-flow theory. The slip lengths are calculated from observed data. *
Units: cm3 per cm2 at minimum pressure. †Flow velocities in cm s–1 at 1 bar.
Liquid Initial permeability* Observed flow velocity† Expected flow velocity† Slip length (µm)
Water 0.58 25 0.00057 54
1.01 43.9 0.00057 68
0.72 9.5 0.00015 39
Ethanol 0.35 4.5 0.00014 28
Iso-Propanol 0.088 1.12 0.00077 13
Hexane 0.44 5.6 0.00052 9.5
Decane 0.053 0.67 0.00017 3.4

nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology 91


REVIEW ARTICLE
a b

~15º

~5º Water

Vapour 200 nm
200 nm

c d

200 nm 200 nm

e f
Vapour

Water

200 nm 50 nm

Figure 5 ESEM images of the dynamic behaviour of a water plug in a carbon nanopipe. a–e, The meniscus shape changes when the stage temperature is constant but the
vapour pressure of the water in the chamber is changed: a, 5.5 Torr, b, 5.8 Torr, c, 6.0 Torr, d, 5.8 Torr and e, 5.7 Torr (where the meniscus returns to the shape seen in a). The
asymmetrical shape of the meniscus, especially the complex shape of the meniscus on the right side in a and e is a result of the difference in the vapour pressure caused by
the open left end and closed right end of the tube. Estimated contact angles between the meniscus and the tube wall are indicated at two locations. f, TEM image showing a
similar plug shape in a closed carbon nanotube under pressure. Reproduced with permission from ref. 32. Copyright (2004) ACS.

nanoscale and to find low friction coatings or channel materials and will form a random coil with a radius of some 700 nm in aqueous
that help to reduce blocking42. solution at 20 °C. This is several times greater than the pore diameter
Another problem is more fundamental: the large size of polymers, of even large carbon pipes and two orders of magnitude greater than
including biologically relevant molecules such as DNA, which are the diameter of a single walled carbon nanotube such as might be
often tangled and tightly folded in vivo. As Cao et al.41 point out, a functionalized to detect the presence of specific base sequences43. DNA
typical DNA molecule from a virus has a length of 100 to 200 kilobases molecules will fit into the central channel of even small nanotubes,

92 nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology


REVIEW ARTICLE
but only when unravelled and fed into the pore opening lengthwise. a
The entropic barrier to achieve this from the disordered state is very
high and therefore such long molecules are normally excluded.
To overcome this problem of interfacing the microscale to
the nanoscale, Cao et al. used optical lithography to fabricate an
array of microchannels that form a gradient from wide to narrow
fluid passages (Fig. 6a)41. They used a novel modified form of
diffraction gradient lithography involving a photosensitive
blocking mask resist on a silicon wafer substrate. The technique
10 μm
is inherently parallel and much faster and more efficient than
using electron-beam lithography. The end result is a massive
array of microposts, with a continuous reduction in the gaps,
which form fluidic channels as the chip is traversed from one
side to the other. b
To test the device, long DNA strands stained with a fluorescent
dye were introduced on the microscale side (right-hand side of
Fig. 6b). Diffusion of the molecules was then observed under
ultraviolet light and captured on video. Still frames from the
recording show individual DNA molecules straightening out and
moving through the interface in an extended configuration towards
the nanoscale region (left-hand side of Fig. 6b). The authors report
transport of the stretched DNA molecules with significantly greater 10 μm
efficiency compared with random diffusion through comparable
nanoscale pores in the absence of any gradient interface.
Melechko et al.36 working at Oak Ridge National laboratory in
Tennessee report a general method for creating patterned arrays of
silica nanopipes precisely positioned over pores in a silicon nitride
membrane on a silicon substrate. The method requires costly Figure 6 Interfacing microfluidics with nanofluidics. a, Optical image of fluidic
equipment and top-down fabrication techniques, but allows a high channels defined by microposts on a chip (after photoresist development). A
degree of control over device architecture. Critically, precise control of continuous reduction of the gaps (light areas) between microposts occurs in the
nanopipe location is achieved by deterministic positioning of catalyst gradient zone. b, Integrated CCD video images of DNA flow through the channels.
particles for CVD growth. The right-hand side of the image shows partially stretched long DNA molecules
Bau et al.44 describe another method for constructing a nanofluidic passing through the micropost array and the gradient zone and continuously entering
device comprising a single carbon nanopipe (diameter 250 nm) the nanochannels on the left-hand side, becoming fully stretched in the process.
connecting two fluid reservoirs. Dielectrophoresis was used to orient Reproduced with permission from ref. 41. Copyright (2002) AIP.
and manipulate the nanopipe into position prior to depositing a
dividing wall on top. These general approaches enable prototyping
of specific functional devices for research and testing. Once effective
architectures have been developed, less expensive methods can be across the tubes and also by functionalizing tube entrances34,35. The
investigated for large-scale manufacture. problem of the transition from micro to nanoflows is being tackled,
Within the past two years, a number of papers have started to for example, by the creation of patterned surfaces, which guide
appear reporting experimentally realized nanofluidic devices. For biomolecules through progressively smaller apertures36. The more
example Flashbart and co-workers45 describe fabrication using contact control that can be engineered into nanopore arrays, the greater the
printing of a microfluidic array with nanochannel interconnects. This prospect of using them in applications requiring high selectivity and
allows their device to take advantage of nanocapillary effects such fast throughput.
as in situ size-based separation, enhanced mixing and molecular It is significant that the size range of nanopores discussed
concentration. Han et al.46 have used surface micromachining in this review is essentially the size range of many important
on silicon nitride, silica and polysilicon in combination with biological entities (antibodies, viruses). Larger molecules such as
micromoulded polydimethylsiloxane to fabricate flexible nanofluidic DNA can be uncoiled36 to fit. Thus carbon nanopipes are potential
device architectures. They have demonstrated attolitre electrokinetic conduits, collimators, sensors, encapsulators and probes for medical
injection using this technique. Finally, Wang et al.47 have created a applications. Clearly there are still many challenges ahead before
nanoscale preconcentration device using standard photolithography such devices become viable including: controlling the mechanical
and etching methods, which has achieved factors in the range of strength of nanoelements in contact with cells and tissue; methods for
106 to 108. They exploit the electrokinetic trapping effect found in the assembly of huge numbers of very small components; fouling of
nanofluidic filters. the nanopipes and surfaces; management of defects in components;
fluid/biomolecule/disease marker input and output; and managing
CONCLUSIONS AND FUTURE DIRECTIONS the information flow from large arrays of nanoscale sensors to the
outside world.
Molecular dynamics simulations16,17,48 have suggested that carbon In a biological context, it is worth noting the vital importance of
nanotubes have very low surface friction with respect to fluid flow. nanoscale pores in natural systems, particularly the channels in cell
This is now being confirmed by direct experiment on timescales membranes involved in ion and water transport. Aquaporin proteins,
which show fast flow corresponding to large slip lengths18,28,25. Water for the elucidation of which Peter Agre earned the Nobel Chemistry
appears to wet carbon nanopipes with flow governed by the Washburn Prize in 2003 (ref. 49), are one example. As is often the case, science
equation, but not for subnanosecond timescales. Control of liquid is only now starting to investigate and appreciate long-established
uptake has been achieved by applying a potential difference27,32,33 natural mechanisms.

nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology 93


REVIEW ARTICLE
doi: 10.1038/nnano.2006.175 27. Kyotani, T., Tsai, L. F. & Tomita, A. Preparation of ultrafine carbon tubes in nanochannels of an
anodic aluminum oxide film. Chem. Mater. 8, 2109–2113 (2006).
28. Mattia, D., et al. Effect of graphitization on the wettability and electrical conductivity of CVD-
References carbon nanotubes and films. J. Phys. Chem. B 110, 9850–9855 (2006).
1. Iijima, S., Helical microtubules of graphitic carbon. Nature 354, 56–58 (1991). 29. Hinds, B. J. et al. Aligned multiwalled carbon nanotube membranes. Science 303, 62–65 (2004).
2. Rowlinson, J. S. & Widom, B. Molecular Theory of Capillarity. Vol. Section 1.3. (Oxford Univ. Press, 30. Majumder, M., Chopra, N., Andrews, R. & Hinds, B. J. Nanoscale hydrodynamics: enhanced flow in
Oxford, 1982). carbon nanotubes. Nature 438, 44 (2005).
3. Palmer, S. B. & Rogalski, M. S. Advanced University Physics (Gordon and Breach, New York, 1996). 31. Holt, J. K. et al. Fast mass transport through sub-2-nanometer carbon nanotubes. Science 312,
4. Washburn, E. W. The dynamics of capillary flow. Phys. Rev. 17, 273–278 (1921). 1034–1037 (2006).
5. de Gennes, P.-G., Brochard-Wyart, F. & Quere, D. Capillarity and Wetting Phenomena. 32. Rossi, M. P. et al. Environmental scanning electron microscopy study of water in carbon nanopipes.
(Springer, Berlin, 2004). Nano Lett. 4, 989–993 (2004).
6. Bosanquet, C. H., The flow of liquids into capillary tubes. Phil. Mag. 6, 525–531 (1923). 33. Kim, B. M., Sinha, S. & Bau, H. H. Optical microscope study of liquid transport in carbon
7. Maxwell, J. C. On stresses in rarified gases arising from inequalities of temperature. Philos. Trans. R. nanotubes. Nano Lett. 4, 2203–2208 (2004).
Soc. 170, 231–256 (1879). 34. Kim, B. M., Murray, T. & Bau, H. H. The fabrication of integrated carbon pipes with sub-micron
8. Sokhan, V. P., Nicholson, D. & Quirke, N. Fluid flow in nanopores: an examination of hydrodynamic diameters. Nanotechnology 16, 1317–1320 (2005).
boundary conditions. J. Chem. Phys. 115, 3878–3887 (2001). 35. Miller, S. A. & Martin, C. R. Redox modulation of electroosmotic flow in a carbon nanotube
9. Quirke, N. (ed.) Adsorption and Transport at the Nanoscale (CRC Press, Boca Raton, membrane. J. Am. Chem. Soc. 126, 6226–6227 (2004).
Florida, 2005). 36. Melechko, A. V., McKnight, T. E., Guillorn, M. A. & Merkulov, V. I. Vertically aligned carbon
10. Lauga, E., Brenner, M. & Stone, H. The no-slip boundary condition: a review in The Handbook Of nanofibers as sacrificial templates for nanofluidic structures. Appl. Phys. Lett. 82, 976–978 (2003).
Experimental Fluid Dynamics (Springer, 2005). 37. Majumder, M., Chopra, N. & Hinds, B. J. Effect of tip functionalization on transport through
11. Powell, C., Fenwick, N., Bresme, F. & Quirke, N. Wetting of nanoparticles and nanoparticle arrays. vertically oriented carbon nanotube membranes. J. Am. Chem. Soc. 127, 9062–9070 (2005).
Colloid. Surface. A 206, 241–251 (2002). 38. Babu, S., Ndungu, P., Bradley, J.-C., Rossi, M. P. & Gogotsi, Y. Guiding water into carbon nanopipes
12. Travis, K. P., Todd, B. D. & Evans, D. J. Departure from Navier–Stokes hydrodynamics in confined with the aid of bipolar electrochemistry. Microfluid. Nanofluid. 1, 284–288 (2005).
liquids. Phys. Rev. E 55, 4288–4295 (1997). 39. Vaitheeswaran, S., Rasaiah, J. C. & Hummer, G. Electric field and temperature effects on water in the
13. Werder, T., Walther, J. H., Jaffe, R. L., Halicioglu, T. & Koumoutsakos, P. On the water-carbon narrow nonpolar pores of carbon nanotubes. J. Chem. Phys. 121, 7955–7965 (2004).
interaction for use in molecular dynamics simulations of graphite and carbon nanotubes. J. Phys. 40. de Mello, A. J. Control and detection of chemical reactions in microfluidic systems. Nature 442,
Chem. B 107, 1345–1352 (2003). 394–402 (2006).
14. Hummer, G., J. C. Rasaiah, and J. P. Noworyta, Water conduction through the hydrophobic channel 41. Cao, H., Tegenfeldt, J. O., Austin, R. H. & Chou, S. Y. Gradient nanostructures for interfacing
of a carbon nanotube. Nature 414, 188–190 (2001). microfluidics and nanofluidics. Appl. Phys. Lett. 81, 3058–3060 (2002).
15. Waghe, A., Rasaiah, J. C. & Hummer, G. Filling and emptying kinetics of carbon nanotubes in water. 42. Kenis, P. J. A. & Stroock, A. D. Materials for micro- and nanofluidics. MRS Bull. 31,
J. Chem. Phys. 117, 10789–10795 (2002). 87–94 (2006).
16. Longhurst, M. & Quirke, N. Environmental effects on the radial breathing modes of carbon 43. Heller, D. A. et al. Optical detection of DNA conformational polymorphism on single-walled carbon
nanotubes in water. J. Chem. Phys. 124, 234708 (2006). nanotubes. Science 311, 508–511 (2006).
17. Sokhan, V. P., Nicholson, D. & Quirke N. Fluid flow in nanopores: accurate boundary conditions for 44. Bau, H. H., Sinha, S., Kim, B. & Riegelman, M. Fabrication of nanofluidic devices and the study of
carbon nanotubes. J. Chem. Phys. 117, 8531–8539 (2002). fluid transport through them in Proc. SPIE. Vol. 5592 (eds Lai, W. Y., Pau, S., Lopez, O. D.) 201–213
18. Supple, S. & Quirke, N. Molecular dynamics of transient oil flows in nanopores I: Imbibition speeds (SPIE, Bellington, Washington, 2005).
for single wall carbon nanotubes. J. Chem. Phys. 121, 8571–8579 (2004). 45. Flachsbart, B. R., et al. Design and fabrication of a multilayered polymer microfluidic chip with
19. Supple, S. & Quirke, N. Rapid imbibition of fluids in carbon nanotubes. Phys. Rev. Lett. 90, nanofluidic interconnects via adhesive contact printing. Lab Chip 6, 667–674 (2006).
214501 (2003). 46. Han, A. P., de Rooij, N. F. & Staufer, U. Design and fabrication of nanofluidic devices by surface
20. Skoulidas, A. I. et al. Rapid transport of gases in carbon nanotubes. Phys. Rev. Lett. 89, micromachining. Nanotechnology 17, 2498–2503 (2006).
185901 (2002). 47. Wang, Y. C., Stevens, A. L. & Han, J. Y. Million-fold preconcentration of proteins and peptides by
21. Pfahler, J., Harley, J, Bau, H. & Zemel, J. Liquid transport in micron and submicron channels. Sensor. nanofluidic filter. Anal. Chem. 77, 4293–4299 (2005).
Actuat. A 22, 431–434 (1990). 48. Skoulidas, A. I., Ackerman, D. M., Johnson, J. K. & Sholl, D. S. Rapid transport of gases in carbon
22. Cheng, J. T. & Giordano N. Fluid flow through nanometer-scale channels. Phys. Rev. E 65, nanotubes. Phys. Rev. Lett. 89, 185901 (2002).
214501 (2002). 49. Agre, P. Aquaporin water channels (Nobel lecture). Angew. Chem. Int. Edn 43, 4278–4290 (2004).
23. Masuda, H., Hasegwa, F. & Ono, S. Self-ordering of cell arrangement of anodic porous alumina 50. Supple, S. Molecular dynamics simulation of capillary flow at the nanoscale. PhD Thesis, Imperial
formed in sulfuric acid solution. J. Electrochem. Soc. 144, L127–L130 (1997). College London (2005).
24. Routkevitch, D., Bigioni, T., Moskovits, M. & Xu, J. M. Electrochemical fabrication of CdS nanowire
arrays in porous anodic aluminum oxide templates. J. Phys. Chem. 100, 14037–14047 (1996).
25. Miller, S. A., Young, V. Y. & Martin, C. R. Electroosmotic flow in template-prepared carbon nanotube Acknowledgements
membranes. J. Am. Chem. Soc. 123, 12335–12342 (2001). M. Whitby acknowledges support from the EPSRC through a Doctoral Training Award and thanks
26. Che, G., Lakshmi, B. B., Fisher, E. R. & Martin, C. R. Carbon nanotubule membranes for F. Barclay and T. Cotter at RGB Research for helpful discussions. The authors are members of the EU
electrochemical energy storage and production. Nature 393, 346–349 (1998). network Inside POReS.

94 nature nanotechnology | VOL 2 | FEBRUARY 2007 | www.nature.com/naturenanotechnology

You might also like