biosensors and bioelectronics · aurich et al. (2007) employed a similar optomagnetic system and a...

8
Blu-ray optomagnetic measurement based competitive immunoassay for Salmonella detection Bo Tian, Rebecca S. Bejhed, Peter Svedlindh, Mattias Strömberg n Department of Engineering Sciences, Uppsala University, Box 534, SE-751 21 Uppsala, Sweden article info Article history: Received 23 July 2015 Received in revised form 28 August 2015 Accepted 30 August 2015 Available online 8 September 2015 Keywords: Biosensors Immunoassays Magnetic particles Magneto-optics abstract A turn-on competitive immunoassay using a low-cost Blu-ray optomagnetic setup and two differently sized magnetic particles (micron-sized particles acting as capture particles and nano-sized particles acting as detection particles) is here presented. For Salmonella detection, a limit of detection of 8 10 4 CFU/mL is achieved within a total assay time of 3 h. The combination of a competitive strategy and an optomagnetic setup not only enables a turn-on read-out format, but also results in a sensitivity limit about a factor of 20 times lower than of volumetric magnetic stray eld detection device based immunoassays. The improvement of sensitivity is enabled by the formation of immuno-magnetic ag- gregates providing steric hindrance protecting the interior binding sites from interaction with the magnetic nanoparticle labels. The formation of immuno-magnetic aggregates is conrmed by uores- cence microscopy. The system exhibits no visible cross-reaction with other common pathogenic bacteria, even at concentrations as high as 10 7 CFU/mL. Furthermore, we present results when using the setup for a qualitative and homogeneous biplex immunoassay of Escherichia coli and Salmonella typhimurium. & 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). 1. Introduction The development of sensitive assays for identication and de- tection of microorganisms is vital for improving diagnosis and preventing disease outbreaks. These assays also play essential roles in control programs provided by food-safety regulatory agencies and disease control and prevention agencies (Swami- nathan and Feng, 1994; Hood et al., 2004; Virgin and Todd, 2011). Many of the latest generation devices, such as surface plasmon resonance (SPR) sensors (Zeng et al., 2014), nuclear magnetic re- sonance (NMR) devices (Harel et al., 2008), and surface enhanced Raman scattering (SERS) spectrometers (Han et al., 2009), are re- garded as promising biosensor setups that have been commer- cialized to a large extent in order to meet the requirements of the society. Unfortunately, the relatively high cost and complexity associated with many of today's diagnostic assay platforms limit their ability to contribute to sustainable health care systems in developing countries (Urdea et al., 2006). To address this problem, the research community has turned to nanoscale labels, e.g. col- loidal gold nanoparticles (Giljohann et al., 2010), quantum dots (Medintz et al., 2005) and magnetic nanoparticles (Haun et al., 2010; Canfarotta and Piletsky, 2014) in the search for user-friendly and inexpensive diagnostic tools that can be used in remote areas where sophisticated lab facilities are not available (Kelley et al., 2014; Brandao et al., 2015). The last decade has yielded breakthroughs within the devel- opment of nano-/microparticle based biosensors, in which the afnity capture is effectively achieved by coating dispersed nano- or microparticles with capture molecules (Gu et al., 2006; Tekin and Gijs, 2013). Due to their ability to be easily manipulated by magnetic eld gradients and the negligible magnetic background of biological objects, superparamagnetic particles have been in- creasingly used as labels in volumetric assays (Borlido et al., 2013; Issadore et al., 2014). In biosensing platforms based on the de- tection of the presence or absence of magnetic particle labeled targets, the read-out is performed in various magnetic stray eld detection devices, measuring changes in either static or dynamic properties of the magnetic particles upon interaction with the target, such as magnetoresistance sensors (Baselt et al., 1998; Graham et al., 2003; Dalslet et al., 2011), Hall effect devices (Mi- hajlovic et al., 2005; Sandhu et al., 2007; Østerberg et al., 2014), magnetic tunnel junctions (Grancharov et al., 2005), super- conducting quantum interference devices (Chemla et al., 2000; Grossman et al., 2004; Strömberg et al., 2008), or AC suscept- ometers (Astalan et al., 2004; Park et al., 2011; Zardán Gómez de la Torre et al., 2011). On the one hand, by employing enzymatic amplication strategies for improving the sensitivity, those mag- netic biosensors are capable of detecting nucleic acid targets with a limit of detection (LOD) in the low pM range (Dalslet et al., 2011; Østerberg et al., 2014; Strömberg et al., 2008; Zardán Gómez de la Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/bios Biosensors and Bioelectronics http://dx.doi.org/10.1016/j.bios.2015.08.070 0956-5663/& 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). n Corresponding author. E-mail address: [email protected] (M. Strömberg). Biosensors and Bioelectronics 77 (2016) 3239

Upload: others

Post on 24-Sep-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

Biosensors and Bioelectronics 77 (2016) 32–39

Contents lists available at ScienceDirect

Biosensors and Bioelectronics

http://d0956-56

n CorrE-m

journal homepage: www.elsevier.com/locate/bios

Blu-ray optomagnetic measurement based competitive immunoassayfor Salmonella detection

Bo Tian, Rebecca S. Bejhed, Peter Svedlindh, Mattias Strömberg n

Department of Engineering Sciences, Uppsala University, Box 534, SE-751 21 Uppsala, Sweden

a r t i c l e i n f o

Article history:Received 23 July 2015Received in revised form28 August 2015Accepted 30 August 2015Available online 8 September 2015

Keywords:BiosensorsImmunoassaysMagnetic particlesMagneto-optics

x.doi.org/10.1016/j.bios.2015.08.07063/& 2015 The Authors. Published by Elsevie

esponding author.ail address: [email protected]

a b s t r a c t

A turn-on competitive immunoassay using a low-cost Blu-ray optomagnetic setup and two differentlysized magnetic particles (micron-sized particles acting as capture particles and nano-sized particlesacting as detection particles) is here presented. For Salmonella detection, a limit of detection of8�104 CFU/mL is achieved within a total assay time of 3 h. The combination of a competitive strategyand an optomagnetic setup not only enables a turn-on read-out format, but also results in a sensitivitylimit about a factor of 20 times lower than of volumetric magnetic stray field detection device basedimmunoassays. The improvement of sensitivity is enabled by the formation of immuno-magnetic ag-gregates providing steric hindrance protecting the interior binding sites from interaction with themagnetic nanoparticle labels. The formation of immuno-magnetic aggregates is confirmed by fluores-cence microscopy. The system exhibits no visible cross-reaction with other common pathogenic bacteria,even at concentrations as high as 107 CFU/mL. Furthermore, we present results when using the setup fora qualitative and homogeneous biplex immunoassay of Escherichia coli and Salmonella typhimurium.& 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license

(http://creativecommons.org/licenses/by/4.0/).

1. Introduction

The development of sensitive assays for identification and de-tection of microorganisms is vital for improving diagnosis andpreventing disease outbreaks. These assays also play essentialroles in control programs provided by food-safety regulatoryagencies and disease control and prevention agencies (Swami-nathan and Feng, 1994; Hood et al., 2004; Virgin and Todd, 2011).Many of the latest generation devices, such as surface plasmonresonance (SPR) sensors (Zeng et al., 2014), nuclear magnetic re-sonance (NMR) devices (Harel et al., 2008), and surface enhancedRaman scattering (SERS) spectrometers (Han et al., 2009), are re-garded as promising biosensor setups that have been commer-cialized to a large extent in order to meet the requirements of thesociety. Unfortunately, the relatively high cost and complexityassociated with many of today's diagnostic assay platforms limittheir ability to contribute to sustainable health care systems indeveloping countries (Urdea et al., 2006). To address this problem,the research community has turned to nanoscale labels, e.g. col-loidal gold nanoparticles (Giljohann et al., 2010), quantum dots(Medintz et al., 2005) and magnetic nanoparticles (Haun et al.,2010; Canfarotta and Piletsky, 2014) in the search for user-friendlyand inexpensive diagnostic tools that can be used in remote areas

r B.V. This is an open access articl

e (M. Strömberg).

where sophisticated lab facilities are not available (Kelley et al.,2014; Brandao et al., 2015).

The last decade has yielded breakthroughs within the devel-opment of nano-/microparticle based biosensors, in which theaffinity capture is effectively achieved by coating dispersed nano-or microparticles with capture molecules (Gu et al., 2006; Tekinand Gijs, 2013). Due to their ability to be easily manipulated bymagnetic field gradients and the negligible magnetic backgroundof biological objects, superparamagnetic particles have been in-creasingly used as labels in volumetric assays (Borlido et al., 2013;Issadore et al., 2014). In biosensing platforms based on the de-tection of the presence or absence of magnetic particle labeledtargets, the read-out is performed in various magnetic stray fielddetection devices, measuring changes in either static or dynamicproperties of the magnetic particles upon interaction with thetarget, such as magnetoresistance sensors (Baselt et al., 1998;Graham et al., 2003; Dalslet et al., 2011), Hall effect devices (Mi-hajlovic et al., 2005; Sandhu et al., 2007; Østerberg et al., 2014),magnetic tunnel junctions (Grancharov et al., 2005), super-conducting quantum interference devices (Chemla et al., 2000;Grossman et al., 2004; Strömberg et al., 2008), or AC suscept-ometers (Astalan et al., 2004; Park et al., 2011; Zardán Gómez de laTorre et al., 2011). On the one hand, by employing enzymaticamplification strategies for improving the sensitivity, those mag-netic biosensors are capable of detecting nucleic acid targets witha limit of detection (LOD) in the low pM range (Dalslet et al., 2011;Østerberg et al., 2014; Strömberg et al., 2008; Zardán Gómez de la

e under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Page 2: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–39 33

Torre et al., 2011). On the other hand, when employing a clinicalaffinity capture strategy, e.g. a bacterial immunoassay, the sensi-tivity of those devices (LOD of higher than 106 CFU/mL) is inferiorcompared to that of a miniaturized nuclear magnetic resonance(NMR) system (LOD of 103 CFU/mL), which benefits from a built-insignal amplification (since millions of water molecules can be af-fected by one individual magnetic particle) (Grossman et al., 2004;Chen et al., 2015). Furthermore, all those aforementioned bio-sensors require expensive instrumentation.

An alternative option for magnetic particle labeled target de-tection is to make use of optomagnetic (magneto-optic) sensorsystems, which measure a modulation of an optical signal relatedto the Brownian relaxation dynamics of magnetic particles (Aurichet al., 2007; Chung et al., 2008; Donolato et al., 2015a). Althoughthe optomagnetic devices have provided significant reductions ininstrumentation cost, their performance regarding magnetic par-ticle labeled target detection is similar to the magnetic stray fielddetection systems discussed above. Chung et al. (2008) estimatedthat the minimum detectable number of MNPs for their opto-magnetic setup was 107. Aurich et al. (2007) employed a similaroptomagnetic system and a homogeneous immunoassay strategyto detection human eotaxin and human IgG, and the LOD wasfound to be 1 nM and 70 pM, respectively. So far, optomagneticsensors have not been reported for the detection of bacteria orcells. In addition, to the best of our knowledge, all reported volu-metric magnetic stray field detection devices and optomagneticdevice based immunoassays have employed direct immunoassaystrategies, which means that they are negative (turn-off) assays,implying that it can be difficult to judge whether the reduction insensor signal is due to poor assay performance or the presence oftarget (van Lierop et al., 2011).

Here, we present a novel Salmonella detection methodologyusing a low-cost Blu-ray optomagnetic setup. The utilization of acompetitive turn-on immunoassay strategy associated with twokinds of magnetic particles (5 μm and 100 nm magnetic particles)gives a LOD of 8�104 CFU/mL within a total assay time of 3 h.Similar bioassay methods have been reported by Donolato et al.(2015a, 2015b) and Bejhed et al. (2015) for DNA detection, but thisis the first time that the Blu-ray optomagnetic sensor is reportedfor a turn-on immunoassay strategy. In this work, for the com-petitive immunoassay scheme, bacteria are affinity captured bybiotinylated-antibody coated 5 μm magnetic particles (Ab-MPs) toform immuno-magnetic aggregates. For certain bacterial con-centrations, large immuno-magnetic aggregates are formed, i.e.clusters of capture MPs and bacteria. Thereafter, the aggregates areincubated with streptavidin-coated 100 nm magnetic nano-particles (MNPs). Finally, the unreacted (non-bound) MNPs in thesuspension are detected by the Blu-ray optomagnetic setup. Theexistence of immuno-magnetic aggregates containing clusters ofAb-MPs is confirmed using fluorescence microscopy. Furthermore,the competitive strategy is compared with a direct immunoassayscheme involving direct binding of Ab-conjugated 100 nm MNPsto bacteria, which shows that the competitive strategy is about 20times more sensitive. Finally, we present a homogeneous and bi-plex detection strategy for Salmonella typhimurium and Escher-ichia coli. In the biplex detection, E. coli directly binds to Ab-250 nm detection MNPs, i.e. a direct detection strategy, while si-multaneously S. typhimurium binds to the Ab-MPs to form im-muno-magnetic aggregates (thereby preventing binding of detec-tion streptavidin-100 nm MNPs to Ab-MPs), i.e. a competitivestrategy. In addition, it is worth mentioning that the cost of thisread-out system is very low in comparison to other devices, e.g.only a few percent of the cost for a miniaturized NMR device(about 50,000€).

2. Experimental section

2.1. Reagents and bacterial strains

Ultrapure grade phosphate buffered saline (PBS, 20� ) waspurchased from AMRESCO (Solon, USA) and was diluted to 1�before being used as the incubation/reaction/washing buffer andblank control sample. Biotinylated rabbit anti Salmonella group(biotinylated polyclonal antibody, with 6–8 biotin groups per an-tibody according to the manufacturer) and rabbit anti E. coli(polyclonal antibody) were purchased from AbD Serotec. Avidinmodified 5 μm MPs (product code 08-18-503), streptavidin mod-ified 100 nm MNPs (product code 10-19-102) and protein Amodified 250 nm MNPs (product code 09-20-252) were purchasedfrom Micromod (Rostock, Germany). Biotin (5-fluorescein) con-jugate was purchased from Sigma-Aldrich and dissolved in DMSO(purchased from Sigma). Heat-killed S. typhimurium cells (CatalogNo. 50-74-01, KPL, Gaithersburg, USA) were used as the positivecontrol sample. Enterococcus faecalis (CCUG 9997), E. coli (CCUG17620), Staphylococcus aureus (CCUG 17621), Proteus mirabilis(CCUG 26767) and Klebsiella oxytoca (CCUG 42935) were obtainedfrom the Department of Medical Sciences, Clinical Microbiologyand Infectious Medicine, Uppsala University. All bacteria weresuspended in PBS and stored at 4 °C.

2.2. Conjugation of magnetic particles and magnetic nanoparticleswith antibodies

For the competitive immunoassay, polyclonal antibodies wereconjugated to avidin-coated MPs (to obtain capture particles, heredenoted Ab-MPs) by mixing biotinylated rabbit anti Salmonellagroup antibody (4 mg/mL) and 5 μm MPs (25 mg/mL,3.5�108 particles/mL) in a volumetric ratio of 1:2 followed byincubation at 37 °C for 2 h to perform the avidin–biotin reaction.After washing two times using a magnetic separation stand, theAb-MPs were resuspended at a concentration of 25 mg/mL(3.5�108 particles/mL) in PBS.

For the direct immunoassay, antibody-conjugated 100 nmMNPs were prepared by mixing biotinylated rabbit anti Salmonellagroup Ab (4 mg/mL) and 100 nm streptavidin-coated MNPs(10 mg/mL) in a volumetric ratio of 3:4 followed by incubation at37 °C for 2 h. After washing two times, the Ab-MNPs were re-suspended at a concentration of 10 mg/mL in PBS.

For the biplex assay, detection MNPs for E. coli were preparedby mixing anti E. coli Ab (4 mg/mL) and protein A modified 250 nmMNPs (10 mg/mL) in a volumetric ratio of 3:4 followed by in-cubation at 37 °C for 2 h and resuspended at a concentration of10 mg/mL.

Ab-MPs and Ab-MNPs were stored at 4 °C for further use.

2.3. Competitive immunoassay for Salmonella

The competitive immunoassay consisted of three independentsteps: capture, labeling and detection (see Scheme 1). First, thesample (blank, positive or negative) was mixed with Ab-MPs andincubated at 37 °C for 1.5 h. The mixture was shaken every 20 minto avoid sedimentation. After incubation and magnetic separation,the immuno-magnetic aggregates were washed once and re-suspended (keeping the volume unchanged) in PBS. Thereafterstreptavidin modified 100 nm magnetic nanoparticles, serving asdetection particles, were added and the final concentration ofMNPs was about 0.1 mg/mL. The MNPs were incubated with theimmuno-magnetic aggregates at 37 °C for 1 h. After this step,65 μL of the suspension was transferred into an UV-transparentdisposable cuvette (REF 67.758.001, SARSTEDT, Nümbrecht, Ger-many). In order to let the immuno-magnetic aggregates sediment

Page 3: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

Scheme 1. Schematic illustration of the Blu-ray optomagnetic measurement basedcompetitive immunoassay for S. typhimurium detection. The assay consists of threeindependent steps (from left to right in the illustration): capture, labelling anddetection. Capture: 5 mm-sized magnetic particles equipped with biotinylatedpolyclonal antibody against Salmonella are added to the sample, positive (top row)or blank (bottom row) and incubated. Labelling: Streptavidin modified 100 nmmagnetic nanoparticles, serving as detection nanoparticles, are added followed byincubation. Detection: The suspension is transferred into a cuvette and the im-muno-magnetic clusters are left to sediment to the bottom of the cuvette prior toperforming target quantification in a Blu-ray optomagnetic system.

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–3934

to the bottom of the cuvette, thus not contributing to the mea-sured optomagnetic response, the cuvette was left on the benchfor 20 min prior to optomagnetic read-out.

2.4. Optomagnetic measurement setup

A detailed description is provided in Supplementary Material,section S1. The output from one measurement (used for targetquantification) is the total intensity of transmitted laser light, V0,the in-phase, V2′ , and the out-of-phase, V2″, components of thecomplex second harmonic signal of the transmitted light. In thecompetitive immunoassay approach, V2′ was normalized with re-spect to the simultaneously measured value of V0 to compensatefor the variations in laser light intensity, particle concentration andcuvette reflection/absorption. In the direct immunoassay andhomogeneous biplex immunoassay, V2′ was directly employed asthe output signal for target quantification since V0 was significantlyinfluenced by the concentration of bacteria in the cuvette. Mea-surements of V2′ and V0 at room temperature were performed inthe frequency range 10–10,000 Hz, using an AC magnetic excita-tion field amplitude of 2.6 mT.

Fig. 1. Panel A: Peak value of V V/2 0′ vs. labelling time normalized with respect to the peathe curves represent pure MNP sample, [MPs (no antibody)þMNPs] sample, [Ab-MPrespectively. Panel B: V V/2 0′ spectra measured for a labelling time of 60 min.

2.5. Fluorescence microscopy study of the immuno-magneticaggregations

A detailed description is provided in Supplementary Material,section S2. Immuno-magnetic aggregates were incubated firstwith MNPs and thereafter with biotin (5-fluorescein) conjugates,and studied in a fluorescence microscope (Axioplan II, Zeiss).

2.6. Homogeneous biplex immunoassay for S. typhimurium and E.coli

A detailed description is provided in Supplementary Material,section S3. Two different kinds of antibody conjugated magneticparticles were added to the sample and incubated. After incuba-tion, the immuno-magnetic aggregates were reacted with MNPsand detected by the optomagnetic read-out.

3. Results and discussion

3.1. Establishment of the competitive immunoassay

Scheme 1 shows the principle behind the competitive im-munoassay. In summary, if S. typhimurium cells and/or fragmentsare suspended in the sample, these entities will react with thebiotinylated antibodies on the surface of the MPs and stericallyprevent these antibodies from interacting with the subsequentlyadded detection nanoparticles (streptavidin MNPs). If no bacteriaare present, a part of the streptavidin MNPs will attach to the Ab-MPs by the streptavidin-biotin reaction, causing the peak value ofV V/2 0′ to be smaller than for a bacteria containing sample. The peakamplitude can therefore be used for qualitative detection of S. ty-phimurium in the sample.

To evaluate the kinetics of the competitive immunoassay, Ab-MPs of 2�106 particles/mL were incubated with a positive controlsample (S. typhimurium of 107 CFU/mL) and a blank control sample(PBS) at 37 °C for 1.5 h (as a capture step). After washing, strep-tavidin modified MNPs were added to the suspensions to a finalconcentration of 0.1 mg/mL (as a labelling step). Thereafter V2′ andV0 were recorded as a function of labelling time (cf. Fig. 1A). Asshown in Fig. 1B, the frequency dependence of V V/2 0′ is closelyrelated to the Brownian relaxation dynamics of the MNPs and

k value of V V/2 0′ measured at 0 min for a pure MNP suspension. From top to bottom,sþpositive controlþMNPs] sample and [Ab-MPsþblank controlþMNPs] sample,

Page 4: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–39 35

shows a peak at approximately 200 Hz. The normalized peak valueof the pure MNP suspension ((V V/2 0′ )max, MNPs, 0 min) was chosen asa reference value for normalizing the peak values of the othersamples. From the results shown in Fig. 1A, it can be seen that(V V/2 0′ )max decreases with increasing labelling time. For the pureMNP sample, the decrease is due to the fact that a small amount ofMNPs adheres to the inner wall of the EP tube, thus not con-tributing to the modulation of the transmitted light. The peakvalue of the [MPs (no antibody)þMNPs] sample is slightly lowerthan that of the pure MNP sample. We attribute such features tointer-assay variations and/or to magnetic interaction fields due toMPs sediment, which attract MNPs to the bottom of the cuvetteand prevent them from contributing to the measured optomag-netic response. For the [Ab-MPsþpositive controlþMNPs] and[Ab-MPsþblank controlþMNPs] samples, the peak decrease ismainly related to the specific (biotin–streptavidin) and nonspecific(bacterial surface-streptavidin) bindings of MNPs to immuno-magnetic aggregates. To maintain a balance between the total la-belling time and signal-to-noise ratio, 60 min was chosen as thereaction time of the labelling step.

Fig. 2. Panel A: Peak value of V V/2 0′ vs. Ab-MP concentration. Panels B, C and D shoconcentration of 1.25�105, 5�105, 2�106 particles/mL, respectively. Note that the peakblank controls. Error bars indicate one standard deviation based on three independent msamples.

3.2. Qualitative detection of S. typhimurium

For the competitive immunoassay strategy (Scheme 1), it canbe concluded that the amount of Ab-MPs used to capture thebacteria is strongly correlated to the sensitivity and specificity ofthe system. On the one hand, the smaller amount of Ab-MPs used,the fewer bacteria are needed to block the biotinylated antibodiesfrom interaction with the subsequently added streptavidin MNPs,and thus the lower the LOD will be. On the other hand, the smalleramounts of Ab-MPs used, the fewer MNPs can bind to MPs, andthus the lower the signal-to-noise ratio will be. To optimize theamount of Ab-MPs, a series of Ab-MP concentrations (from 2�103

to 2�106 particles/mL) was chosen for reacting with 0.1 mg/mLstreptavidin MNPs, and a cut-off value was calculated as theaverage of (V V/2 0′ )max for a triplicate measurement using a blankcontrol sample (PBS) minus three times the standard deviation ofthe triplicate (Fig. 2A). Although the dose-response curve indicatesthat the smallest concentration below the cut-off value (red da-shed line) was 3�104 particles/mL, the signal-to-noise ratio forthis amount of Ab-MPs was too low to be used for further

w peak value of V V/2 0′ vs. S. typhimurium concentration when using an Ab-MPvalues in panels B–D have been normalized with respect to the peak values of theeasurements. Dashed lines indicate the cut-off value calculated from blank control

Page 5: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–3936

experiments. The coefficients of variation for each point of thecurve shown in Fig. 2A were less than 4.8%.

A concentration series of S. typhimurium was measured usingthe competitive immunoassay. The immunoassay was performedusing Ab-MP concentrations of 1.25�105, 5�105 and2�106 particles/mL, and the results of those measurements areshown in Fig. 2B–D, respectively. Normalized peak values of theblank controls (( V V/2 0′ )max, Blank Control) were used not only tocalculate the cut-off values (average signal for a triplicate mea-surement plus three times the standard deviation of the triplicate)for each experiment, but also to normalize the signals to com-pensate for the inter-assay variations. As mentioned earlier, theLOD for S. typhimurium improves upon decreasing the amount ofAb-MPs, and the best LOD 8�104 CFU/mL was achieved for an Ab-MP concentration of 1.25�105 particles/mL (Fig. 2B). Interestingly,the smallest measured S. typhimurium concentration above thecut-off value corresponded well with the concentration of Ab-MPsused, implying that large immuno-magnetic aggregates were ef-ficiently formed when having a MP/bacteria ratio of 1:1. Thecoefficients of variation for each point of the curves shown inFig. 2B–D were less than 3.7%, 5.4% and 6.7% respectively.

The shape of the dose-response curves in Fig. 2B–D is governedmainly by two mechanisms: (1) formation of immuno-magneticaggregates containing clusters of Ab-MPs, providing steric hin-drance preventing streptavidin-MNPs from binding to biotinylatedantibodies inside clusters; and (2) saturation of Ab-MPs by bac-teria, which prevents direct binding of streptavidin-MNPs to Ab-MPs. Immuno-magnetic aggregates containing clusters of Ab-MPsform below a certain bacterial concentration limit, which dependson the concentration of MPs (around 1:1 ratio of MP/bacteria). TheAb-MP cluster size increases with increasing bacterial concentra-tion up to this limit, which results in an increasing dose-responsecurve. Above this limit the aggregates containing several Ab-MPsno longer form, and the immuno-magnetic aggregates mainlycontain single Ab-MP. At even higher bacterial concentrations,more and more bacteria bind to the surfaces of Ab-MPs, yielding a

Fig. 3. Representative fluorescence photomicrographs of MPs after reaction with the indused in these experiments.

dose-response increasing with bacterial concentration and even-tually to a plateau in the curve when the Ab-MPs become satu-rated with bacteria. We thus infer that the sudden decrease ofsignal between the peak and the plateau represents a situationwhere aggregates containing clusters of Ab-MPs cannot form, butinstead the immuno-magnetic aggregates appear as single Ab-MPswith only a few bound bacteria each, thus having free binding sitesfor MNPs.

In another test described in Supplementary Material, sectionS4, the Ab-MPs/bacteria incubation time was shortened to 0.5 h toinvestigate its influence on the sensitivity of the assay. In themeasurement employing 5�105 particles/mL Ab-MPs, the sensi-tivity was not significantly affected by the smaller incubation timecompared with the dose-response curve in Fig. 2C. To balance thesensitivity and stability, subsequent competitive immunoassaytests were performed using an Ab-MPs concentration of5�105 particles/mL.

3.3. Fluorescence microscopy study of Ab-MP/bacteria binding

The fluorescence photomicrographs (Fig. 3) confirm the ex-istence of immuno-magnetic aggregates containing clusters of Ab-MPs and also show that the formation of such immuno-magneticaggregates mainly exists below a certain bacterial concentrationlimit. It should be mentioned that the small biotin (5-fluorescein)molecule can easily diffuse into the interior of the immuno-mag-netic aggregations and bind to both free avidin groups on the MPsand to free streptavidin groups on the MNPs. This enables easy andstraightforward visualization of free Ab-MPs and immuno-mag-netic aggregations in a fluorescence microscope.

In this study, Ab-MPs with a concentration of5�105 particles/mL were reacted with samples containing differ-ent concentrations of bacteria. Because of steric hindrance, bioti-nylated antibodies cannot block all the biotin binding sites on thesurface of avidin-MPs, implying that the bacteria–Ab-MPs con-jugates can still be labeled by the biotin (5-fluorescein) conjugate.

icated S. typhimurium samples. An Ab-MP concentration of 5�105 particles/mL was

Page 6: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

Fig. 4. Specificity study of the competitive immunoassay. Common pathogenicbacteria at a concentration of 107 CFU/mL in PBS were investigated. The con-centration of Ab-MPs was 5�105 particles/mL. Error bars indicate one standarddeviation based on three independent measurements.

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–39 37

When exposed to a high concentration of S. typhimurium (i.e.107 CFU/mL), Ab-MPs were covered by excessive bacterial frag-ments and could, thus, not form aggregates containing clusters ofAb-MPs (Fig. 3A). This in turn resulted in that only a few MNPs(with bound fluorescent labels) were bound to the Ab-MPs, inaccordance with the observed plateau at the position of107 CFU/mL in Fig. 2C. As the concentration ratio of bacteria/Ab-MPs was reduced, non-saturated immuno-magnetic aggregatescontaining single Ab-MPs coexisting with a few aggregates con-taining small clusters of Ab-MPs were observed (Fig. 3B), but theseimmuno-magnetic aggregates cannot sterically block the bindingsites for MNPs, in accordance with the valley at the position of2�106 CFU/mL in Fig. 2C. Immuno-magnetic aggregates contain-ing large clusters of Ab-MPs (Fig. 3C) create steric hindrancearound the interior binding sites which reduces the amount ofbound streptavidin MNPs, thus giving rise to a peak around106 CFU/mL S. typhimurium in Fig. 2C. In Fig. 3D, Ab-MPs were stillfree after being incubated with the blank control sample (PBS),and the fluorescence intensity was slightly higher than for the MPsin the other images. This observation can be explained by the factthat in this case a large number of MNPs (with fluorescent labelson surface) were attached to the Ab-MPs.

3.4. Comparison of competitive and direct immunoassay strategies

By using the same reagents and read-out device, a direct im-munoassay strategy for S. typhimurium detection was evaluatedand compared with the competitive assay approach (details aregiven in Supplementary Material, section S5). The LOD for thedirect immunoassay was found to be approximately2�106 CFU/mL (cf. Fig. S4), which corroborates previous results(5.6�106 CFU/mL) reported by Grossman et al. (2004). A dynamicrange of at least two orders of magnitude is obtained for the directimmunoassay strategy. A competitive immunoassay is generallyless sensitive than the corresponding direct assay, but in this studyit was proven to be more sensitive. This could be explained by thefact that the amount of biotin binding sites on the surface of theAb-MPs is much larger and easier to control than the amount ofantigen sites on the surface of typical bacteria. In addition, theformation of immuno-magnetic aggregates containing clusters ofAb-MPs strongly reduces the number of available binding sites,thus further enhancing the sensitivity. In the direct immunoassaystrategy, immuno-magnetic aggregates could also be formed byAb-MNPs and bacteria. (Wang et al., 2015) However, cross-linkingbetween Ab-MNPs and bacteria means that the bacteria are notsaturated with Ab-MNPs (bacteria are sharing the labels), andhence, the sensitivity is limited. If we regard the 5 μm Ab-MPs as“ideal bacteria” with target molecules covering their entire sur-face, the LOD of a direct immunoassay for this “bacteria” would be3�104 particles/mL, which can be appreciated by inspection ofFig. 2A. Since it is nearly impossible to achieve such an ideal si-tuation in real clinical tests, the competitive assay strategy per-forms better than the direct immunoassay strategy. This can beexplained by the fact that the strength of antibody–antigen in-teraction (employed in the direct immunoassay) is variable andnot as strong as the biotin–streptavidin pair used in this compe-titive immunoassay.

3.5. Specificity evaluation of competitive and direct immunoassays

The specificity (selectivity) of the competitive and direct im-munoassay approaches was investigated by considering five typesof other common pathogenic bacteria (see Section 2.1) as negativecontrols (at 107 CFU/mL and 108 CFU/mL, respectively). Resultspresented in Fig. 4 show that the normalized peak values,(V V/2 0′ )max/(V V/2 0′ )max, Blank Control, of the negative control samples

are close to 1, meaning that the response from the negative con-trols are similar to the blank control. Similar results were obtainedfor the direct immunoassay strategy (see Fig. S5), indicating thatno significant amounts of bacteria other than S. typhimurium werecaptured by the Ab-MPs (in competitive immunoassay) or Ab-MNPs (in direct immunoassay). Therefore, the current approacheswere able to differentiate S. typhimurium from other clinicalcommon pathogenic bacteria and are therefore suitable for thespecific detection of S. typhimurium.

3.6. Homogeneous biplex immunoassay for S. typhimurium andE. coli

This optomagnetic setup can be used for multiplex detection,since MNPs with different hydrodynamic sizes have V 2′ peak/val-ley values at different frequencies. The V 2′ valley (negative peak) of250 nm MNPs is located at 18 Hz, while the V 2′ peak of 100 nmMNPs is located at 184 Hz. Direct and competitive immunoassayswere employed for E. coli and S. typhimurium using 250 nm and100 nm MNPs as detection labels, respectively. As shown inFig. 5A, each spectrum has its own signature in terms of valley/peak values and frequency positions. For the spectra of the E. colicontaining samples, i.e., the yellow triangles and blue diamonds,the absolute values of the valley at 14–18 Hz are smaller than forthe blank sample (red squares), since a fraction of the Ab-250 nmMNPs are bound to the E. coli. When measuring the spectra ofS. typhimurium containing samples, the black circles and bluediamonds, the Salmonella fragments covered the Ab-MPs and leftmore streptavidin-100 nm MNPs free in the suspension, resultingin higher peak values at 184 Hz. In view of the homogeneous bi-plex immunoassay results (Fig. 5B), we conclude that the setup canbe applied for qualitative multiplex immunoassays.

4. Conclusion

In summary, we have presented a turn-on competitive im-munoassay using a low-cost Blu-ray optomagnetic setup for read-out. This qualitative detection strategy utilizes two differently

Page 7: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

Fig. 5. Panel A: Average V 2′ spectra of three independent measurements measured for a homogeneous biplex immunoassay. Direct and competitive immunoassays wereemployed for E. coli and S. typhimurium using 250 nm and 100 nm MNPs as detection labels, respectively. The optomagnetic response of free Ab-250 nm MNPs andstreptavidin-100 nm MNPs was characterized by a valley/peak at 18 Hz and 184 Hz, respectively. E: 107 CFU/mL E. coli; S: 107 CFU/mL S. typhimurium. 0: absence of bacteria.Panel B: Bar graph showing the value of V 2′ at the frequency of 18 Hz and 184 Hz. Error bars indicate one standard deviation based on three independent measurements.

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–3938

sized magnetic particles and gives a LOD of 8�104 CFU/mL S. ty-phimurium within a total assay time of 3 h. The improvement ofthe sensitivity was achieved by the formation of immuno-mag-netic aggregates containing clusters of Ab-MPs providing sterichindrance protecting the interior binding sites from interactionwith the magnetic nanoparticle labels. The existence of immuno-magnetic aggregates containing clusters of Ab-MPs was confirmedby fluorescence microscopy. This presented method is fully com-patible with a large variety of pathogens and can be used for thedetection of urinary tract samples, which have a high positive CFUthreshold (usually 105 CFU/mL) (Mezger et al., 2015). Moreover,our method can be easily combined with the use of a MP pre-treatment procedure to enrich the bacteria before the test, therebyfurther improving the LOD. The immunoassay exhibits no visiblecross-reaction with other clinically common pathogenic bacteria,even at a concentration of 107 CFU/mL. If a positive answer isobtained from the qualitative competitive assay, we can furtherquantify the bacterial concentration by employing a direct im-munoassay approach. Furthermore, we have demonstrated aqualitative and homogeneous biplex immunoassay of E. coli andS. typhimurium. In consideration of the advantages shown in thisstudy, as well as the low-cost of the setup, the system has a greatpotential to be used for cost-efficient and user-friendly identifi-cation and detection of microorganisms relating to e.g. food safety.

Acknowledgments

This work was financially supported by Swedish Research CouncilFormas (Project no. 221-2012-444). Dr. Eva Tano at the Department ofMedical Sciences, Clinical Microbiology and Infectious Medicine, Up-psala University, and Ma Jing at the Department of Immunology, Ge-netics and Pathology (IGP), Uppsala University are gratefully ac-knowledged for providing bacterial samples. Dr. Camilla Russell at IGP,Uppsala University, is gratefully acknowledged for helpful assistancewhen running the fluorescence microscope.

Appendix A. Supplementary information

Supplementary data associated with this article can be found inthe online version at http://dx.doi.org/10.1016/j.bios.2015.08.070.

References

Astalan, A.P., Ahrentorp, F., Johansson, C., Larsson, K., Krozer, A., 2004. Biosens.Bioelectron. 19 (8), 945–951.

Aurich, K., Nagel, S., Glockl, G., Weitschies, W., 2007. Anal. Chem. 79 (2), 580–586.Baselt, D.R., Lee, G.U., Natesan, M., Metzger, S.W., Sheehan, P.E., Colton, R.J., 1998.

Biosens. Bioelectron. 13 (7–8), 731–739.Bejhed, R.S., Zardán Gómez de la Torre, T., Svedlindh, P., Strömberg, M., 2015. Bio-

technol. J. 10 (3), 469–472.Borlido, L., Azevedo, A.M., Roque, A.C., Aires-Barros, M.R., 2013. Biotechnol. Adv. 31

(8), 1374–1385.Brandao, D., Liebana, S., Pividori, M.I., 2015. N. Biotechnol. 32 (5), 511–520.Canfarotta, F., Piletsky, S.A., 2014. Adv. Healthc. Mater. 3 (2), 160–175.Chemla, Y.R., Crossman, H.L., Poon, Y., McDermott, R., Stevens, R., Alper, M.D.,

Clarke, J., 2000. Proc. Natl. Acad. Sci. USA 97 (26), 14268–14272.Chen, Y.P., Xianyu, Y.L., Wang, Y., Zhang, X.Q., Cha, R.T., Sun, J.S., Jiang, X.Y., 2015. ACS

Nano 9; , pp. 3184–3191.Chung, S.H., Grimsditch, M., Hoffmann, A., Bader, S.D., Xie, J., Peng, S., Sun, S., 2008.

J. Magn. Magn. Mater. 320 (3–4), 91–95.Dalslet, B.T., Damsgaard, C.D., Donolato, M., Strømme, M., Strömberg, M., Svedlindh,

P., Hansen, M.F., 2011. Lab Chip 11 (2), 296–302.Donolato, M., Antunes, P., Bejhed, R.S., Zardán, Gómez, de la Torre, T., Østerberg, F.

W., Strömberg, M., Nilsson, M., Strømme, M., Svedlindh, P., Hansen, M.F., Va-vassori, P., 2015a. Anal. Chem. 87 (3), 1622–1629.

Donolato, M., Antunes, P., Zardán Gómez de la Torre, T., Hwu, E.T., Chen, C.H.,Burger, R., Rizzi, G., Bosco, F.G., Strømme, M., Boisen, A., Hansen, M.F., 2015b.Biosens. Bioelectron. 67, 649–655.

Giljohann, D.A., Seferos, D.S., Daniel, W.L., Massich, M.D., Patel, P.C., Mirkin, C.A.,2010. Angew. Chem. Int. Ed. 49 (19), 3280–3294.

Graham, D.L., Ferreira, H.A., Freitas, P.P., Cabral, J.M., 2003. Biosens. Bioelectron. 18(4), 483–488.

Grancharov, S.G., Zeng, H., Sun, S.H., Wang, S.X., O’Brien, S., Murray, C.B., Kirtley, J.R.,Held, G.A., 2005. J. Phys. Chem. B 109 (26), 13030–13035.

Grossman, H.L., Myers, W.R., Vreeland, V.J., Bruehl, R., Alper, M.D., Bertozzi, C.R.,Clarke, J., 2004. Proc. Natl. Acad. Sci. USA 101 (1), 129–134.

Gu, H., Xu, K., Xu, C., Xu, B., 2006. Chem. Commun. (Camb) 9, 941–949.Han, X.X., Zhao, B., Ozaki, Y., 2009. Anal. Bioanal. Chem. 394 (7), 1719–1727.Harel, E., Schroder, L., Xu, S.J., 2008. Annu. Rev. Anal. Chem. 1, 133–163.Haun, J.B., Yoon, T.J., Lee, H., Weissleder, R., 2010. WIREs Nanomed. Nanobiotechnol.

2 (3), 291–304.Hood, L., Heath, J.R., Phelps, M.E., Lin, B., 2004. Science 306 (5696), 640–643.Issadore, D., Park, Y.I., Shao, H., Min, C., Lee, K., Liong, M., Weissleder, R., Lee, H.,

2014. Lab Chip 14 (14), 2385–2397.Kelley, S.O., Mirkin, C.A., Walt, D.R., Ismagilov, R.F., Toner, M., Sargent, E.H., 2014.

Nat. Nanotechnol. 9 (12), 969–980.Medintz, I.L., Uyeda, H.T., Goldman, E.R., Mattoussi, H., 2005. Nat. Mater. 4 (6),

435–446.Mezger, A., Fock, J., Antunes, P., Østerberg, F.W., Boisen, A., Nilsson, M., Hansen, M.F.,

Ahlford, A., Donolato, M., 2015. ACS Nano 9 (7), 7374–7382.Mihajlovic, G., Xiong, P., von Molnar, S., Ohtani, K., Ohno, H., Field, M., Sullivan, G.J.,

2005. Appl. Phys. Lett. 87, 11.Østerberg, F.W., Rizzi, G., Donolato, M., Bejhed, R.S., Mezger, A., Strömberg, M.,

Page 8: Biosensors and Bioelectronics · Aurich et al. (2007) employed a similar optomagnetic system and a homogeneous immunoassay strategy to detection human eotaxin and human IgG, ... (2015a,

B. Tian et al. / Biosensors and Bioelectronics 77 (2016) 32–39 39

Nilsson, K.T., Strømme, M., Svedlindh, P., Hansen, M.F., 2014. Small 10 (14),2877–2882.

Park, K., Harrah, T., Goldberg, E.B., Guertin, R.P., Sonkusale, S., 2011. Nanotechnology22, 8.

Sandhu, A., Kumagai, Y., Lapicki, A., Sakamoto, S., Abe, M., Handa, H., 2007. Biosens.Bioelectron. 22 (9–10), 2115–2120.

Strömberg, M., Göransson, J., Gunnarsson, K., Nilsson, M., Svedlindh, P., Strømme,M., 2008. Nano Lett. 8 (3), 816–821.

Swaminathan, B., Feng, P., 1994. Annu. Rev. Microbiol. 48, 401–426.Tekin, H.C., Gijs, M.A., 2013. Lab Chip 13 (24), 4711–4739.Urdea, M., Penny, L.A., Olmsted, S.S., Giovanni, M.Y., Kaspar, P., Shepherd, A., Wilson,

P., Dahl, C.A., Buchsbaum, S., Moeller, G., Burgess, D.C.H., 2006. Nature 444,73–79.

van Lierop, D., Faulds, K., Graham, D., 2011. Anal. Chem. 83 (15), 5817–5821.Wang, D.B., Tian, B., Zhang, Z.P., Wang, X.Y., Fleming, J., Bi, L.J., Yang, R.F., Zhang, X.E.,

2015. Biosens. Bioelectron. 67, 608–614.Virgin, H.W., Todd, J.A., 2011. Cell 147 (1), 44–56.Zardán Gómez de la Torre, T., Mezger, A., Herthnek, D., Johansson, C., Svedlindh, P.,

Nilsson, M., Strømme, M., 2011. Biosens. Bioelectron. 29 (1), 195–199.Zeng, S.W., Baillargeat, D., Ho, H.P., Yong, K.T., 2014. Chem. Soc. Rev. 43 (10),

3426–3452.