polymerwaveguide_klee htlee

Upload: handoyoeko20017573

Post on 10-Apr-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    1/7

    Polymer waveguide backplanes for optical sensor interfaces inmicrofluidics{{

    Kevin S. Lee,* Harry L. T. Lee and Rajeev J. Ram

    Received 28th June 2007, Accepted 14th August 2007

    First published as an Advance Article on the web 21st August 2007DOI: 10.1039/b709885p

    A polymer optical backplane capable of generic luminescence detection within microfluidic chips

    is demonstrated using large core polymer waveguides and vertical couplers. The waveguides are

    fabricated through a new process combining mechanical machining and vapor polishing with

    elastomer microtransfer molding. A backplane approach enables general optical integration with

    planar array microfluidics since optical backplanes can be independently designed but still

    integrated with planar fluidic circuits. Fabricated large core waveguides exhibit a loss of 0.1 dB

    cm21 at 626 nm, a measured numerical aperture of 0.50, and a collection efficiency of 2.86% in an

    n = 1.459 medium, comparable to a 0.50 NA microscope objective. In addition to vertical couplers

    for out-of-plane collection and excitation, polymer waveguides are doped with organic dyes to

    provide wavelength selective filtering within waveguides, further improving optical device

    integration. With large core low loss waveguides, luminescence collection is improved and

    measurements can be performed with simple LEDs and photodetectors. Fluorescein detection via

    fluorescence intensity with a limit of detection (3s) of 200 nM in a 1 mL volume is demonstrated.

    Phosphorescence lifetime based oxygen detection in water in an oxygen controllable microbial cell

    culture chip with a limit of detection (3s) of 0.08% or 35 ppb is also demonstrated utilizing the

    waveguide backplane. Single waveguide luminescence collection performance is equivalent to a

    back collection geometry fiber bundle consisting of nine 500 mm diameter collection fibers.

    1. Introduction

    Luminescence based sensors are used for determining environ-

    mental parameters such as dissolved oxygen or pH in

    biological systems as well as for providing markers necessary

    for cell identification and sorting. Such optical sensingtechniques can provide real time monitoring capabilities and

    dynamic control without disturbing a biological systems

    equilibrium or introducing contaminants into the system. As

    these new lab-on-a-chip systems become more integrated, the

    number of sensors required increases, leading to an increase in

    size and complexity of the off-chip detection systems. While

    CCD camera imaging is well suited to applications with slow

    timescale (,100 Hz) intensity measurements, spatially dis-

    tributed systems requiring measurements at both high speed

    and low intensity, such as fluorescence lifetime or flow

    cytometry, need the speed and sensitivity of photodiodes and

    non-imaging optics.1,2 Fiber optics, or more generally,

    waveguides are typically used to decouple the location ofmeasurement points and photodetectors, increasing design

    flexibility. In addition waveguides can reduce the number of

    photodetectors in these systems by routing multiple signals to

    a common photodetector.

    One common approach utilizing fiber optics places fibers

    close to fluorescent sources to provide excitation light as well

    as route collected signals to detectors with minimal loss.

    However, in general, luminescence collected by a single fiber is

    very low. In one instance, a 550 mm fiber performing excitation

    and detection was only 0.0009% efficient in detecting

    fluorescence from an oxygen sensitive ruthenium complex2

    due to direct dye deposition on the small fiber. Fiber bundles

    can overcome the low collection efficiency of single fibers when

    collecting light from a distributed source, but typically require

    many fibers, which can be expensive to assemble and not

    useful when high spatial resolution is required. For example, a

    fiber bundle system designed to collect fluorescence from

    stained tissue samples consisted of thirty 100 mm fibers and still

    required a photomultiplier tube to perform detection.3 In

    addition, large fiber bundles are impractical for devices

    requiring multiple sensors.

    We have developed a platform of waveguide based passive

    optical components utilizing the same polymer materials usedto fabricate microfluidic systems. Integration of excitation and

    collection optics into microfluidic systems improves alignment

    tolerances with external optical elements and also increases

    system design flexibility. Many different designs have been

    proposed for absorption or luminescence detection,4 including

    many waveguide based geometries such as waveguide arrays,1

    right angle collection,5 in-line absorption,6 evanescent cou-

    pling,7 and hollow prisms.8 While each of the waveguide based

    systems demonstrated optofluidic integration, all of the

    designs either required waveguides to be in-plane or in direct

    contact with fluid channels, or required fluidic channels to be

    MIT, EECS, 32 Vassar St. 26-459, Cambridge, MA, 02139, USA.E-mail: [email protected]{ The HTML version of this article has been enhanced with colourimages.{ Electronic supplementary information (ESI) available: Maximumwaveguide width calculation, loss and roughness measurements andsystem measurements. See DOI: 10.1039/b709885p

    PAPER www.rsc.org/loc | Lab on a Chip

    This journal is The Royal Society of Chemistry 2007 Lab Chip, 2007, 7, 15391545 | 1539

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    2/7

    redesigned to accommodate waveguides. Such designs make

    the layout of integrated optical systems dependent on fluidic

    component locations and hinder device scalability. Previous

    designs also did not consider the filters required to implement

    on-chip fluorescence detection and often required lasers and

    expensive detectors to detect a signal. In contrast to the strictly

    planar designs demonstrated previously, waveguide back-

    planes with vertical couplers allow for the integration of opticsinto fluidic systems with minimal impact on the fluidic design.

    Using a polymer optical backplane, we will demonstrate a

    potentially disposable high sensitivity lifetime based lumines-

    cence detection system utilizing only light emitting diodes

    (LED) and silicon photodetectors.

    2. Design

    In order to maximize the signal to noise ratio of the detection

    system, stray excitation light must be minimized at the detector

    and luminescence collection must be maximized. A back-

    scattering collection geometry satisfies the first requirement

    because only interface reflections contribute to stray light. For a

    planar waveguide system, a backscattering collection geometry

    can be achieved by forcing the light out of plane utilizing

    waveguide vertical couplers. Due to the low index contrast of

    the waveguides, the vertical couplers are metallized to maximize

    efficiency. The backplane design is shown in Fig. 1a.

    The second requirement, maximizing luminescence collec-

    tion, depends on the location and behavior of the fluorescent

    source. To demonstrate the backplane concept, we assume

    little knowledge of the location of the emitting source and

    optimize for collecting luminescence located 1 mm above the

    backplane, or the thickness of a typical glass slide. Under this

    constraint, the minimum core size of the integrated wave-

    guides, which maximizes efficiency, is determined by the

    distance h from the emitting source and the waveguide

    numerical aperture NA = (ncore22 nclad

    2)1/2 which specifies

    the sine of the maximum input angle to the waveguide. With

    ncore = 1.56 and nclad = 1.459 for the materials used (section 3),the NA is 0.55. Through ray geometry, eqn (1) for the

    minimum width can be derived taking into account coupling

    through next = nclad instead of air. This equation is derived in

    the ESI.{ At h = 1 mm, the minimum core width is

    approximately 800 mm.

    w~2hffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    next=NA 2{1

    q (1)

    Sensor excitation and collection efficiency is also influenced

    by the waveguide vertical coupler bend angle. While optical

    backplanes developed for other interconnect applications9

    focus on coupling as much power out of the waveguide as

    possible, a backplane for luminescence detection focuses on

    maximizing both sensor excitation and collection of sensor

    emission. To meet this condition, the sensor should be spatially

    illuminated where the collection waveguide is most efficient.

    This occurs when the excitation and collection waveguide bend

    angles result in intensity profiles which overlap at the sensor

    plane.

    To determine the spatial distribution of excitation light

    incident on the sensor, ray tracing simulations were performed

    starting with a point source located at the waveguide input

    and ending at the sensor surface. For the collection efficiency

    of the output waveguide, the efficiency needs to be determined

    for each position on the sensor plane. Rays representing

    point source emissions were placed along the sensor plane, asshown in Fig. 1a, and propagated through the collection

    waveguide. The ratio of light exiting the output/filter

    waveguide to light emitted by the point source determines

    the collection efficiency at a particular point on the sensing

    plane. By performing this simulation for points throughout

    the surface plotted in Fig. 1b and 1c, a contour of the

    collection area efficiency versus sensor position can be

    determined. Fig. 1b and Fig. 1c show contours of the

    excitation profiles (dashed lines) and point source collection

    efficiencies (solid lines) at the sensing plane for 45u and 30u

    vertical couplers.

    For oxygen sensors used in this paper, fluorophor emission

    is collected through a polystyrene encapsulating layer, 100 mmglass, and then 1 mm of polydimethylsiloxane (PDMS). For a

    well centered point source above this specific material stack, a

    multimode fiber or microscope objective with a 0.55 NA

    achieves a collection efficiency of 3.13% into the first lens

    calculated from ray optics10 ignoring reflection effects. In ray

    tracing simulations, the 30u coupler, 1 mm2 collection

    waveguide design presented here achieves a maximum point

    source collection of 2.86%, close to the multimode fiber

    maximum, suggesting that this design would be suitable for

    cytometry systems using similar NA single lens collection

    geometries, which achieve collection efficiencies of 4%.10

    Fig. 1 (a) A side and top view of the integrated oxygen sensor chip is

    shown. The waveguides are designed to excite and collect luminescence

    from the sensor located 1 mm above the silvered bends. (b,c) Ray

    tracing simulations of the intensity distribution for the input

    waveguide excitation and point source collection efficiency of the

    output waveguide at the sensing plane for 45u bends and 30u bends,

    respectively. System performance increases by 200% after replacing 45u

    bends with 30u bends.

    1540 | Lab Chip, 2007, 7, 15391545 This journal is The Royal Society of Chemistry 2007

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    3/7

    While the point source collection efficiency is large, the

    overall system performance also includes the excitation

    efficiency. To include excitation efficiency, the system effi-

    ciency is recalculated as the ratio of emission leaving the

    output waveguide to excitation entering the input waveguide.

    Efficiency at the sensor plane is determined by taking the

    overlap of the excitation and collection areas given in Fig. 1c.

    The total system efficiency resulting from a maximum angulardistribution excitation is only 0.40% with 45u couplers and a

    factor of 2.8 times better at 1.14% for optimized 30u couplers.

    To further improve the collection efficiency of the waveguides,

    the collection waveguide can be expanded in width to

    maximize collection given the large area sensor illumination.

    A width expansion to 3 mm brings the system efficiency to

    1.73%. While the collection efficiency of this waveguide system

    is approximately half of the maximum achievable with a

    microscope (3.13%), it provides a compact design, incorporat-

    ing both excitation and collection.

    3. Fabrication

    As mentioned in section 2, the 1 mm standoff between the

    sensors and waveguides requires that waveguides be fabricated

    on a mm scale to perform efficient luminescence collection.

    Standard lithography processes are not amenable to fabricat-

    ing master molds with features of this size since mm

    thicknesses require multiple spin coating and exposure steps.

    In SU-8 processes, layers of 100 mm thickness already suffer

    from substrate delamination due to stress.11 As a result, a

    computer numerical control (CNC) machining approach is

    used to fabricate the mm thick master molds necessary for

    polymer molding.

    The general waveguide fabrication process is shown in Fig. 2.

    Polycarbonate positive master molds were machined using a

    CNC mill (Sherline 2000) and a 1 mm diameter square end mill

    (Microcut USA) to fabricate 800 mm 6 1 mm channels or ribs

    (Fig. 2a). To make the waveguide mold as smooth as possible,

    milling was performed around the waveguide structure,

    defining the sidewalls only, with a first pass at 2000 rpm and

    a 25 mm min21 feed speed, and a second pass removing an

    extra 100 mm at 2000 rpm and a 50 mm min21 feed speed to

    improve finish quality. In this way, we took advantage of the

    pre-polished polycarbonate surface as one of the waveguide

    walls. 30u vertical couplers are then milled with 120u full angle

    drill mills (Microcut USA) at each detection location with the

    same machining parameters.

    Roughness due to milling was then reduced through solvent

    vapor polishing where a vapor of methylene chloride diffuses

    into the plastic surface and causes reflow12

    (Fig. 2b). Thedegree of polishing could be controlled by varying the solvent

    pressure or also by varying the exposure time. Under optimal

    polishing conditions, the roughness average (Ra) from sanded

    polycarbonate samples could be reduced from 1000 nm to

    70 nm.13 By combining vapor polishing with CNC milling, this

    method becomes viable for both optical and microfluidic

    master mold fabrication.

    PDMS (Dow Corning Sylgard 184) was used to replicate the

    polycarbonate master molds, forming the waveguide cladding

    (Fig. 2c). Since the refractive index difference between PDMS

    (n = 1.459 at 600 nm via ellipsometry) and the core material,

    polyurethane (NOA 71 Norland Products) (n = 1.56 given by

    manufacturer), was too low to reflect light out of plane using

    total internal reflection, the vertical couplers were coated with

    300 nm of silver to improve reflectivity. By first applying air

    plasma to modify PDMS into a glass-like layer (Fig. 2d), high

    quality silver films could be deposited with e-beam evapora-

    tion (Fig. 2e) on the molded PDMS cladding.14 A 100 mm sheet

    of PDMS, made by spin coating onto a fluorinated silicon

    wafer, was then air plasma bonded to the top cladding mold to

    seal the channels and provide a flat surface for interfacing with

    fluidic chips (Fig. 2f).

    Once the hollow channels in PDMS were complete, the

    waveguide core material was introduced. Before waveguide

    core injection, all of the PDMS channels were surface treated

    with air plasma to increase their surface energy. This reduces

    the contact angle between liquids and cured PDMS, increasing

    capillary forces as well as binding strength at the core cladding

    interface.15 Without first applying a surface treatment, bubbles

    formed along the sidewalls and delamination of the core from

    the cladding occurred due to shrinkage after curing. In this

    work, no bubbles were observed, and while 6% shrinkage

    occurred, there was no sign of delamination. For all

    waveguides, NOA71 was filled into the channels via direct

    injection and capillary filling (Fig. 2g).

    For collection waveguides with integrated filters, NOA71

    was first directly mixed with an organic dye (pinacyanol

    iodide) which absorbs at 600 nm in order to provide integrated

    filtering of the excitation light. The dye was allowed to saturate

    the prepolymer. Integrating filters into the waveguides canimprove both coupling and collection efficiency by removing

    1 to 5 mm thick external filters and allowing LEDs and

    detectors to be placed closer to the waveguide facets. This can

    eliminate reflection losses from external filters and reduce free

    space divergence losses associated with an increased path

    length. Similar dye doped polymer filters have also been

    produced for bulk PDMS layers.16 After the channels were

    filled with core material, the waveguide array was cured under

    UV illumination. While diffusive mixing between dye doped

    and clear NOA71 occurred before complete cure was reached,

    the dye was still able to fully absorb excitation light. No

    Fig. 2 The fabrication process for making large core polymer channel

    waveguides with a CNC milling process. The core preparation process

    for providing integrated filter materials is not shown.

    This journal is The Royal Society of Chemistry 2007 Lab Chip, 2007, 7, 15391545 | 1541

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    4/7

    photobleaching of the dye absorption was observed after a 24 h

    UV-exposure.

    After core curing, the waveguide ends were cleaved to the

    appropriate size (Fig. 2h). Since cleaving elastomeric wave-

    guides results in poor quality end faces, additional NOA71 was

    deposited on each end and cured to repair the cut ends (Fig. 2i).

    This smoothed each of the end faces and improved input

    coupling dramatically.The fabricated device shown in Fig. 3 consists of a

    waveguide backplane interfaced with a fluidic device capable

    of controlling dissolved oxygen concentration. Design and

    fabrication of this fluidic system has been previously

    described.17 The fluidic device consists of 4 wells in parallel

    with integrated peristaltic mixers separated by a thin gas

    permeable membrane. Pressurized gas actuation of the

    membrane translates into diffusion of the gas into the fluid

    in the well. By placing an oxygen sensor in each well, the

    oxygen concentration of the fluid can be measured and

    controlled with changes in gas concentration. This microfluidic

    system is ideal for testing lifetime based sensors since the

    dissolved oxygen in the wells of the chip can be controlled. The

    backplane contains one waveguide sensor for each well

    coupled to the oxygen sensor inside the well.

    The oxygen sensors used for this experiment are composed

    of a mixture of platinum(II) octaethylporphine ketone (PtOEP-

    K Frontier Scientific) and polystyrene deposited on glass disks.

    Glass disks were etched with HF prior to deposition of the

    polymer/dye solution to enhance binding of the fluorophor

    film to the sensor substrate. This phosphorescent dye has a

    maximum absorption at 592 nm and emits at 759 nm with a

    12% quantum yield.18

    4. Results and discussion

    Before using the waveguide backplane for luminescencemeasurements, the fabricated system must be characterized.

    We first measure straight waveguide test structures to

    determine the loss, numerical aperture, and surface roughness

    effects of the fabricated waveguides. We then characterize the

    waveguide filters and the silvered vertical couplers.

    Total system performance is then characterized, including

    the input and output coupling efficiency between an LED and

    the backplane. After demonstrating that luminescence collec-

    tion can be performed above the noise limit of the detection

    system, the backplane is integrated with a fluidic chip capable

    of performing dissolved oxygen control in water and

    measurements of oxygen concentration are performed.

    A. Waveguide characterization

    Fabricated large core waveguides (800 mm 6 1 mm) exhibited

    a propagation loss of 0.10 dB cm21 at 632 nm from a HeNe

    laser via the cut-back method. The measured waveguide

    numerical aperture (NA) was 0.50 (50% intensity). The NA

    reduction from the expected 0.55 to 0.50 in the large core

    waveguides reduces the collection efficiency and most likely

    results from sidewall roughness still present after the polishing

    process. By assuming uniform scattering within the core and

    ignoring interface effects at the waveguide ends, the scattered

    light from the waveguide is also directly proportional to the

    power inside the waveguide. Measurements of the scattered

    light intensity versus the length of the waveguide can therefore

    serve as an alternative method to measure loss. An exponential

    fit to the scattered light intensity imaged using an 8-bit CCDcamera yielded a similar propagation loss as the cut-back

    method of 0.13 dB cm21.

    To compare with previously reported polymer waveguides,

    90 mm 6 90 mm 6 70 mm waveguides were fabricated.

    Perpendicular scattering power measurements yielded a

    waveguide loss of 1.26 dB cm21 at 632 nm. If the majority

    of loss results from sidewall scattering, then a 106 increase in

    loss is expected for a 106 decrease in the width of the

    waveguide core. This loss value compares well with the 1 dB

    cm21 of other previously reported CNC machined wave-

    guides19 but was achieved with more tolerant machining

    parameters due to the additional vapor polishing step. The loss

    is also comparable to 75 mm 6 50 mm 6 70 mm waveguidesfabricated by IBM with measured losses of 1.2 dB cm21 for an

    NA of 0.29.20 Details of the loss measurements and numerical

    aperture measurements are provided in the ESI.{

    B. Dye doped filters

    Filter waveguides were also fabricated to determine their

    absorption and fluorescence characteristics. At maximum

    solubility in NOA71, the filters containing pinacyanol iodide

    (Sigma Aldrich) achieved an absorption per unit length of

    71.8 dB cm21 at 590 nm and less than 0.30 dB cm21 at 760 nm

    as measured with an Ocean Optics white light source and

    spectrometer through a 1 mm path length of pinacyanol iodide

    doped NOA71. Waveguide coupled dye autofluorescence wasmeasured by creating a 70 mm long dye doped waveguide and

    injecting a 632 nm HeNe laser at the input of the waveguide.

    Since the input is entirely absorbed, dye fluorescence can be

    measured by measuring the waveguide output power. This

    results in a 1% output power relative to the input laser power

    as measured with the photodiode circuit used for actual

    experiments. Pinacyanol iodide has a short lifetime (330 ps in

    viscous fluids such as glycerol),21 and therefore the fluores-

    cence emission adds constant intensity and phase to the

    detected signal at kHz frequencies. For frequency lock-in

    measurements, a filter dye with a slower lifetime would be

    Fig. 3 The fabricated device is shown consisting of 4 waveguide

    sensors butt coupled to photodetectors. The optical backplane rests

    underneath a miniature bioreactor array containing 4 integrated

    oxygen sensors. Darker sections of the waveguides result from mixing

    the core polymer with organic dye and acts as a filter for the 590 nm

    excitation source.

    1542 | Lab Chip, 2007, 7, 15391545 This journal is The Royal Society of Chemistry 2007

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    5/7

    preferable to provide both absorption based attenuation and

    frequency response attenuation.

    C. Silvered vertical couplers

    To test the effects of air plasma on silver surface deposition

    quality, 300 nm silver films were deposited on both air plasma

    treated and native flat PDMS surfaces. While silver evaporated

    on air plasma treated surfaces achieved reflectivities of 98% as

    expected, silver evaporated on untreated PDMS surfaces

    resulted in microcracks, reducing reflectivity to 90%. In

    addition to silver deposition quality, bend roughness from

    milling resulted in variations in surface flatness. In comparison

    to flat surfaces, measurements on vertical coupler surfaces

    resulted in a reduction in reflectivity from 98% to 96%, in

    agreement with Monte Carlo simulations of a rough bend

    incorporating measured roughness.

    D. System performance

    By characterizing all of the coupling, roughness, and material

    losses in the experimental system, simulated system perfor-mance was found to agree with experimental system perfor-

    mance. Overall efficiency from LED output to collected

    luminescence was calculated to be 0.0109% while measured

    results were 0.01%. A detailed description of the system losses

    and characterization measurements are described in ESI.{

    To compare the waveguide collection efficiency with

    conventional fiber bundles, we also tested a fiber bundle with

    a 1 mm core diameter PMMA excitation fiber surrounded by

    0.5 mm core diameter PMMA collection fibers. Measurements

    performed with the fiber bundle on the same oxygen sensor

    required 9 collection fibers in order to yield a similar efficiency

    of 0.0095%. Out of 9 collection fibers, a maximum efficiency of

    0.0014% was measured for a single fiber.

    E. Intensity detection (fluorescein fluorescence)

    For intensity based fluorescence detection, a 500 mm thick 1 6

    2 cm2 area microfluidic well was placed above the waveguide

    detection area. Sodium fluorescein (Sigma Aldrich) in deio-

    nized water was pumped through the well in increasing

    concentrations starting with concentrations below the detec-

    tion limit of the system. Due to the close excitation and

    emission spectra, external filters were used. The input

    excitation to the waveguide consisted of a filtered (Omega

    Optical HQ480/406) 470 nm fiber coupled LED with an

    output power of 8.6 mW. The fiber from the LED was coupled

    to a second glass fiber with the excitation filter placed betweento eliminate PMMA autofluorescence, and the second fiber

    was butt coupled to the waveguide. A fluorescence filter

    (Omega Optical HQ510LP) was also placed between the

    collection waveguide output and the silicon photodiode. The

    amplifier circuit provided a gain of 3 6 106 V A21.

    Fig. 4 shows the collected fluorescence intensity for different

    concentrations of sodium fluorescein with the background

    signal measured when DI water is in the well subtracted. From

    the plot, we can determine the 3s noise level for an average of

    ten 1 second acquisitions, where s is the standard deviation of

    the measured intensity when the well contains no fluorescein.

    From these measurements, the estimated limit of detection

    (LOD) for sodium fluorescein is 200 nM with a linear

    detection range up to 100 mM. With a 500 mm channel heightand an illumination area of 2 6 1 mm2 estimated in section 2,

    the detection volume is 1 mL. The LOD achieved by this system

    is considerably higher than other similar fluorescence detec-

    tors, where an LOD of less than 0.04 nM was acheived,22,23

    mainly resulting from autofluorescence of the waveguide core

    material in combination with the filters used. Measurements of

    the background signal show a larger detected signal than the

    noise floor of the photodetector circuit by over 2 orders of

    magnitude. Improvements to the LOD for fluorescein detec-

    tion are possible if other core materials are used which

    generate less autofluorescence.

    F. Lifetime detection (oxygen sensing)

    An example application replaces an array of optical fiber

    bundles with the waveguide backplane for measurements of

    dissolved oxygen in a microbioreactor chip.17 Sampling

    oxygen quickly and accurately is necessary for feedback

    control of the dissolved oxygen concentration. Due to the

    high density of fluidics and control valves in this device, strictly

    planar waveguides can not be implemented along side the

    microfluidics. The experimental setup for measuring oxygen

    concentration with the waveguide backplane consisted of four

    590 nm fiber-coupled LEDs modulated at 5 kHz. To filter IR

    emission from the LEDs, short pass filters (BG-39) were

    placed at the end of the LED fibers and butt coupled to

    secondary PMMA fibers. The second fibers were then buttcoupled to the input waveguides and four silicon photodiodes

    were placed at each output guide with the option of placing an

    off-chip filter (RG-9) between the output guide and the

    photodiodes. Unlike the 470 nm excitation, no autofluores-

    cence was detected at the photodetector with a 590 nm

    excitation and RG-9 filter. The concentration of dissolved

    oxygen in water was controlled by the peristaltic mixer which

    mixes the liquid in the chamber with pressurized nitrogen or

    pressurized air.17

    The phosphorescence of the PtOEP-K dye, which composes

    the oxygen sensors, is quenched by molecular oxygen, resulting

    Fig. 4 Collected fluorescence intensity versus concentration at an

    input power of 8.6 mW. The measurement setup is also shown. From

    the standard deviation of the data for undetectable concentrations, a

    limit of detection of 200 nM can be estimated for sodium fluorescein.

    This journal is The Royal Society of Chemistry 2007 Lab Chip, 2007, 7, 15391545 | 1543

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    6/7

    in a reduction in the phosphorescence lifetime. By measuring

    the phase shift of the collected phosphorescence relative to the

    modulated input, a signal related to the phosphorescence

    lifetime of the sensor can be determined. Each sensor exhibited

    a maximum phase change of 30u with a standard deviation of

    0.5u from nitrogen to air when the off-chip filter was used as

    shown in Fig. 5a. Sensor calibration was performed by mixing

    ratios of nitrogen and air using an electrically controlledswitching valve (Lee Co.) operating at 20 Hz. Oxygen

    concentration could be varied between 0% and 21% by

    changing the duty cycle of the switch between nitrogen and

    air. Fitting the phase response versus frequency to a first order

    model resulted in phosphorescence lifetimes ranging from

    19.8 ms to 58 ms, which is comparable with previously reported

    calibrations.18 The linearity of the inset in Fig. 5a shows that

    the extracted lifetimes are accurately described by Stern

    Volmer quenching kinetics. Using data from the frequency

    response versus oxygen concentration, a calibration curve

    relating phase to oxygen concentration can be determined at

    specific frequencies as shown in Fig. 5b.

    A step response of the oxygen concentration in water,measured by switching the peristaltic actuation gas from

    nitrogen to air, is shown in Fig. 5c. Since the lifetime based

    sensor measures fluorescence quenching due to oxygen, the

    noise increases when the well is air mixed versus nitrogen

    mixed. The noise at maximum concentration (21% oxygen)

    was obtained from the standard deviation (s) of ten 1 second

    acquisitions during air mixing. By converting the deviation in

    phase into a deviation in concentration, the worst case

    resolution (3s) for detecting oxygen in water was estimated.

    For off-chip glass filtered devices, the minimal resolution was

    measured at 0.4% oxygen. At 100% nitrogen, a similar

    measurement results in an LOD at 0.08% oxygen. For a

    dissolved oxygen concentration of 9.2 ppm in air saturated

    water, this translates to an LOD of 35 ppb oxygen. This LOD

    is comparable to the original work characterizing PtOEP-K18

    as well as other fiber based optodes2,24 and is lower than other

    waveguide oxygen sensors where LODs of 600 and 300 ppboxygen were achieved7,25 It should be noted that while

    dynamic range is not affected by intensity variations due to

    the ratiometric nature of the phase measurement, both the

    LOD and resolution improve with increasing intensity since

    the signal to noise ratio is only between 30 and 56 for air and

    nitrogen, respectively. On-chip filtered waveguides measured a

    worst case resolution of 5% oxygen for air saturated water, and

    an LOD of 2% oxygen for nitrogen saturated water. The

    decrease in resolution can be attributed to autofluorescence in

    the dye doped collection waveguide which further reduces the

    signal to noise ratio.

    To test the waveguide sensor repeatability, each waveguide

    sensor was exposed to the same set of nitrogen and oxygenmixtures and the responses of different sensors to the same

    concentrations were compared. Gas mixtures were created

    with the same switching valve used for calibrations. Sensors

    were exposed to six different oxygen concentrations and the

    measured concentration was taken as the average of 20 acquisi-

    tions. Fig. 5d shows the standard deviation of the measured

    concentration between the four sensors. From the data, we can

    see that the off-chip filtered waveguides measure concentration

    very consistently, with intersensor deviations of only 0.025%

    and 0.14% oxygen concentration for nitrogen purged and air

    saturated conditions. On-chip filtered waveguides performed

    worse, due mainly to signal contributions from stray excitation

    and dye autofluorescence. This reduced the full scale phase

    range to only 6u, inflating phase deviations. Even with a

    reduced phase range, on-chip filtered waveguides, using the

    same calibration curve measured in Fig. 5b, measure

    intersensor deviations of 0.22% and 0.53% oxygen under

    nitrogen purged and air saturated conditions, respectively, for

    averaged measurements.

    5. Conclusions

    The polymer based optical backplane fabricated in PDMS

    consisted of large area waveguides, silvered vertical couplers,

    and filters. A fabrication method combining CNC milling and

    vapor polishing was shown to generate sufficiently smooth

    surfaces for waveguiding. Measured propagation loss for800 mm 6 1 mm waveguides was 0.10 dB cm21 at 632 nm

    and 30u silvered vertical coupler reflectivity was 96%. By

    fabricating waveguides in a polymer backplane platform, low

    cost optical detection solutions can be designed easily for

    applications where CCDs are either too expensive or incapable

    of detection, such as at IR wavelengths or for high frequency

    measurements. Organic dye doping of waveguides provided

    integrated filters, further reducing the complexity and cost of

    off-chip optics. The fabricated backplane demonstrated

    accurate oxygen concentration measurements via the phos-

    phorescence lifetime of sensors located within a complex

    Fig. 5 (a) A measured phase vs. frequency plot (dotted) is shown for

    the oxygen sensors as measured with the waveguides and off-chip

    filters. Curve fits (solid) using phosphorescence lifetimes ranging from

    19.8 ms to 58 ms for air and nitrogen, respectively, are also shown. (b)

    The inset shows the calibration curve of normalized phase vs. %

    saturated oxygen concentration at 5 kHz. (c) Measured dissolved

    oxygen concentration step responses from the chip shown in Fig. 3

    with 4 different waveguide sensors utilizing both off-chip filters

    (dashed) and on-chip filters (solid). (d) Repeatability measurements

    showing deviations in measured oxygen concentration for the four

    waveguides with on-chip (square) and off-chip filters (circle).

    1544 | Lab Chip, 2007, 7, 15391545 This journal is The Royal Society of Chemistry 2007

  • 8/8/2019 PolymerWaveguide_KLee HTLee

    7/7

    microfluidic device with required input powers as low as 17 mW

    for an SNR of 56. Simulations of the system indicate a point

    source collection efficiency of 2.86%, suggesting that the

    backplane could be used for fast particle fluorescence detection

    systems such as flow cytometers in addition to the demon-

    strated distributed area luminescence detection system.

    References1 C. H. Chen, High-Sensitivity Scattering-based Detection under

    Symmetrical Arrayed-Waveguide Platform, presented at the IEEESummer Topicals Optofluidics, Quebec City, QC, Canada, July1719, 2006, paper TuA4.2.

    2 P. A. S. Jorge, P. Caldas, C. C. Rosa, A. G. Oliva and J. L. Santos,Optical fiber probes for fluorescence based oxygen sensing, Sens.Actuators, B, 2004, 103, 290299.

    3 B. W. Pogue and G. Burke, Fiber-optic bundle design forquantitative fluorescence measurement from tissue, Appl. Opt.,1998, 37(31), 74297436.

    4 M. Dandin, P. Abshire and E. Smela, Optical filtering technologiesfor integrated fluorescence sensors, Lab Chip, 2007, 7, 955977.

    5 D. A. Chang-Yen and B. K. Gale, An Integrated Optical GlucoseSensor Fabricated Using PDMS Waveguides on a PDMSSubstrate, Proc. SPIEInt. Soc. Opt. Eng., 2004, 5345, 98107.

    6 K. Windhorn, L. Meixner, S. Drost, H. Schroder, E. Kilgus,W. Scheel, H. Reichl, F. Ebling, M. Kuchler, J. Nester, T. Otto andT. Gebner, Polymer BioMEMs with integrated optical and fluidicfunctionality, Proc. Micro Syst. Technol. 2003, 2003, 416423.

    7 D. A. Chang-Yen and B. K. Gale, An integrated optical oxygensensor fabricated using rapid-prototyping techniques, Lab Chip,2003, 3, 297301.

    8 A. Llobera, R. Wilke and S. Buttgenbach, Poly(dimethylsiloxane)hollow Abbe prism with microlenses for detection based onabsorption and refractive index shift, Lab Chip, 2004, 4, 2427.

    9 J. Moisel, J. Guttmann, H. P. Huber, O. Krumpholz, M. Rode,R. Bogenberger and K. P. Kuhn, Optical backplanes withintegrated polymer waveguides, Opt. Eng., 2000, 39, 673679.

    10 P. M. Goodwin, W. P. Ambrose, J. C. Martin and R. A. Keller,Spatial Dependence of the Optical Collection Efficiency in FlowCytometry, Cytometry, 1995, 21, 133144.

    1 1 S . J . H on g, S . C ho i, Y . C ho i, M . A ll en a nd G . M ay ,

    Characterization of Low-Temperature SU-8 PhotoresistProcessing for MEMS Applications, IEEE Conf. WorkshopASMC 2004, 2004, 404408.

    12 V. D. McGinniss, Vaporous Solvent Treatment of ThermoplasticSubstrates, U. S. Pat. 4529563, July 16, 1985.

    13 K. Lee, H. L. T. Lee and R. J. Ram, Large Core PolymerWaveguides for Optical Backplanes in Microfluidic Systems,LEOS Summer Topical Meetings, 2006, 2006, 3839.

    14 H. Schmid, H. Wolf, R. Allenspach, H. Riel, S. Karg, B. Micheland E. Delamarche, Preparation of metallic films on elastomericstamps and their application for contact processing and contactprinting, Adv. Funct. Mater., 2003, 13, 145153.

    15 S. Bhattacharya, A. Datta, J. M. Berg and S. Gangopadhyay,Studies on Surface Wettability of Poly(Dimethyl) Siloxane(PDMS) and Glass Under Oxygen-Plasma Treatment andCorrelation With Bond Strength, J. Microelectromech. Syst.,2005, 14(3), 590597.

    16 O. Hofmann, X. Wang, A. Cornwell, S. Beecher, A. Raja,D . D . C . B ra dl ey , A . J . d eM el lo a nd J . C . d eM el lo ,Monolithically integrated dye-doped PDMS long-pass filters fordisposable on-chip fluorescence detection, Lab Chip, 2006, 6,981987.

    17 H . L. T. Lee, P. Boccazzi, R. J. Ram and A. Sinskey,Microbioreactor arrays with integrated mixers and fluid injectorsfor high-throughput experimentation with pH and dissolvedoxygen control, Lab Chip, 2006, 6, 12291235.

    18 D. B. Papkovsky, G. V. Ponomarev, W. Trettnak and P. OLeary,Phosphorescent Complexes of Porphyrin Ketones: OpticalProperties and Application to Oxygen Sensing, Anal. Chem.,1995, 67, 41124117.

    19 D. Snakenborg, G. Perozziello, H. Klank, O. Geschke and

    J. P. Kutter, Direct milling and casting of polymer-based opticalwaveguides for improved transparency in the visible range,J. Micromech. Microeng., 2006, 16, 375381.

    20 C. M. Olsen, J. M. Trewheila, B. Fan and M. M. Oprysko,Propagation Properties in Short Lengths of Rectangular EpoxyWaveguides, IEEE Photon. Technol. Lett., 1992, 4(2), 145148.

    21 S. Schneider, Flashlamp-pumped mode-locked dye lasers, Philos.Trans. R. Soc. London, Ser. A, 1980, 298, 233245.

    22 R. Irawan, C. M. Tay, S. C. Tjin and C. Y. Fu, Compactfluorescence detection using in-fiber microchannels its potentialfor lab-on-a-chip applications, Lab Chip, 2006, 6, 10951098.

    23 R. Irawan, S. C. Tjin, X. Fang and C. Y. Fu, Integration ofoptical fiber light guide, fluorescence detection system, andmultichannel disposable microfluidic chip, Biomed. Microdev.,2007, 9, 413419.

    24 E. J. Park, K. R. Reid, W. Tang, R. T. Kennedy and R. Kopelman,Ratiometric fiber optic sensors for the detection of inter- and intra-

    cellular dissolved oxygen, J. Mater. Chem., 2005, 15, 29132919.25 C. S. Burke, O. McGaughey, J. M. Sabattie, H. Barry, A.K. McEvoy, C. McDonagh and B. D. MacCraith, Development ofan integrated optic oxygen sensor using a novel, generic platform,Analyst, 2005, 130, 4145.

    This journal is The Royal Society of Chemistry 2007 Lab Chip, 2007, 7, 15391545 | 1545