Open Access
1 July 2010 Direct detection of aptamer-thrombin binding via surface-enhanced Raman spectroscopy
Cynthia V. Pagba, Stephen M. Lane, Hansang Cho, Sebastian Wachsmann-Hogiu
Author Affiliations +
Abstract
In this study, we exploit the sensitivity offered by surface-enhanced Raman scattering (SERS) for the direct detection of thrombin using the thrombin-binding aptamer (TBA) as molecular receptor. The technique utilizes immobilized silver nanoparticles that are functionalized with thiolated thrombin-specific binding aptamer, a 15-mer (5'-GGTTGGTGTGGTTGG-3') quadruplex forming oligonucleotide. In addition to the Raman vibrational bands corresponding to the aptamer and blocking agent, new peaks (mainly at 1140, 1540, and 1635 cm−1) that are characteristic of the protein are observed upon binding of thrombin. These spectral changes are not observed when the aptamer-nanoparticle assembly is exposed to a nonbinding protein such as bovine serum albumin (BSA). This methodology could be further used for the development of label-free biosensors for direct detection of proteins and other molecules of interest for which aptamers are available.

1.

Introduction

Detection and tracking of biomolecules is invaluable not only in the field of medical diagnostics but also in biological and biotechnology research, drug development, environmental monitoring, forensic investigations, and biodefense. To this end, researchers of different expertise have sought to develop methodologies/techniques that are sensitive, specific, robust, high throughput (amenable to multiplexing), simple, and cost effective. Presently, most methods used to detect biomolecules that require molecular recognition events (e.g., formation of antibody-antigen or aptamer-analyte complexes) are variations of the “sandwich type” enzyme-linked immunosorbent assays (ELISA) and typically rely on the use of receptors tagged with probe molecules, which may not always be desirable.1, 2, 3 Labeling of biomolecules is a time-consuming process that can lead to the loss of biological activity. Other assays make use of nanoparticles and their tendency to form aggregates, which results in color change due to a shift in the plasmon resonance in response to binding events.4, 5 While this is a convenient method, it is limited by its susceptibility to changes in the environment, e.g., change in ionic strength or pH, which can also induce aggregation and color change. Another nanoparticle-based technique takes advantage of the ability of metal nanoparticles to quench emission from nearby fluorescent compounds. In this case, the presence of the analyte is detected by the change in the intensity of emission from the fluorescent molecule attached to the receptor.6, 7 Direct detection techniques such as mass spectrometry and surface plasmon resonance (SPR) spectroscopy are also widely used, but they are expensive and potentially limited by extensive sample preparation. For example, mass spectrometry requires isolation and purification of the analyte (usually by chromatography).8 SPR, which detects binding of the molecule by measuring the change in the refractive index at the surface of the metal functionalized with a conjugated receptor, is subject to nonspecific interactions.

Another technique that is currently under development is based on surface-enhanced Raman scattering (SERS). SERS occurs when a molecule is in close proximity to the metal surface. When combined with the resonance Raman effect, signal enhancements of up 1014 to 1015 have been observed, allowing even single molecule detection9 under very favorable conditions. This huge signal enhancement is attributed to enhanced electromagnetic fields within the immediate vicinity of the metal upon excitation of plasmon resonances by photon interaction,10 and to charge transfer processes between the metal and the adsorbed molecule.11

SERS-based assays,3, 12, 13, 14 which require molecular recognition events, have been made possible by labeling either the capture agent or the target molecule itself with a Raman-active molecule. Recently, Neuman 15 reported SERS-based direct detection of the platelet derived growth factor (PDGF) by monitoring the randomness of the SERS spectra of the PDGF aptamer, following its exposure to the target molecule PDGF. The change in the reproducibility of the pristine anti-PDGF aptamer was thought to be due to change in the conformation of the aptamer as a result of protein binding.

Aptamers are functional single-stranded oligonucleotides (DNA or RNA) generated by the process called systematic evolution of ligands by exponential enrichment (SELEX).16 They bind to their target molecules selectively and with high affinity by forming secondary structures and shapes (e.g., quadruplex, hairpin loop, and T-junction).17, 18 Several aptamers have been developed for the detection of certain biological and chemical threat agents [e.g., ricin,19 anthrax spores,20 and trinitrotoluene (TNT)21]; cancer biomarkers;22, 23, 24, 25 HIV-associated proteins;26, 27, 28, 29 food-borne pathogens;30 and other biologically important biomolecules such as insulin,31, 32 immunoglobulin E (IgE),33 and thrombin.34 They exhibit protein binding affinities that are comparable to those of corresponding antibodies.18, 35, 36, 37, 38 As in the case of antibodies, labeling of the aptamer can sometimes lead to loss of activity resulting from loss of ability to form the secondary structure that is required for binding.

In the present study, we exploit the sensitivity offered by SERS in the development of an aptamer-based direct-detection method for target molecules requiring molecular recognition using the thrombin and thrombin binding aptamer (TBA)34 pair as a model system. TBA is a 15-mer ( 5 -GGTTGGTGTGGTTGG- 3 ) oligonucleotide that binds to the fibrinogen-recognition exosite of thrombin39 after forming a chair-type G-quadruplex structure40, 41, 42, 43, 44 (Fig. 1 ). The compact size of the folded aptamer (1.1nm) 45 makes it an ideal receptor for this assay, as it brings the target molecule in close proximity to the metal surface, i.e., the analyte is well within the required distance (<10nm) for local SERS field enhancement.

Fig. 1

(a) Antiparallel quadruplex or chair conformation of the TBA. (b) Guanine quartet of the TBA formed via the Hoogsten-type guanine-guanine interaction. In an antiparallel quadruplex, the guanosine residues have an alternating syn and anti conformation.

047006_1_032004jbo1.jpg

The technique (Fig. 2 ) utilizes immobilized nanoparticles that are functionalized with an aptamer. Binding of thrombin is detected by the presence of additional Raman bands following exposure of a TBA-functionalized SERS substrate to the test sample. The method described here may reduce the number of steps required in a sandwiched-based assay, and has the potential to discriminate between specific and nonspecific binding events.

Fig. 2

Scheme for direct detection of thrombin. (a) Nanoparticles are immobilized on a glass surface previously treated with (3-aminopropyl) trimethoxysilane. (b) Nanoparticles are functionalized with the thiolated thrombin binding aptamer. (c) The exposed metal surface is blocked by 6-mercaptohexanol. (d) The target molecule (thrombin in this case) is introduced. SERS spectrum is obtained at each step.

047006_1_032004jbo2.jpg

2.

Experimental Section

The general protocol for the assay is presented in Fig. 2. First, the silver nanoparticles are immobilized on a glass surface previously treated with 3-aminopropyl trimethoxysilane. The immobilized nanoparticles are functionalized with a thiolated thrombin binding aptamer and the surface is blocked with 6-mercaptohexanol to prevent nonspecific binding. A solution containing thrombin is then introduced. SERS spectra are obtained for each step to monitor the relevant binding events. Binding of thrombin is detected by the presence of additional Raman bands following exposure of TBA-functionalized SERS substrate to the test sample.

2.1.

Materials

Monobasic and dibasic potassium phosphate (KPi) and potassium chloride were obtained from Fisher Scientific. Phosphate buffered saline (PBS) was purchased from USB Corporation, Silver nitrate (AgNO3) , sodium citrate, (3-aminopropyl) trimethoxysilane (APTMS), 6-mercaptohexanol (MH), bovine serum albumin (BSA), and thrombin were purchased from Sigma Chemical Company (Saint Louis, Missouri). Thrombin was dialyzed (Spectrapor Biotech membrane 3000 MWCO) against PBS for 3h prior to use. The thiolated thrombin binding aptamer ( 5 -HS- (CH2)5CH2 -GTTGGTGTGGTTGG- 3 ) was ordered from Sigma Chemical Company. It was supplied as a disulfide form and was used without hydrolysis. Glass cover slips (25CIR-1) were purchased from Fisher Scientific Incorporated (Pittsburgh, Pennsylvania). All glasswares used were cleaned with aqua regia.

2.2.

Silver Nanoparticle Preparation

Silver nanoparticles (NPs) were prepared following the procedure of Lee and Meisel.46 Briefly, 20mg of AgNO3 was dissolved in 100mL of Milli-Q water and the solution was brought to boiling, after which 20mL of 0.1% sodium citrate was added drop-wise. The mixture was then refluxed for one hour. The resulting solution was greenish yellow in color. The plasmon resonance was centered at about 412nm as measured by a Cary UV-Vis spectrophotometer (Varian, Palo Alto, California). Functionalization of the nanoparticles resulted in the formation of a new broad absorption band centered at about 660nm , which allows for SERS measurements at our laser excitation wavelength (647nm) . The nanoparticles have an average diameter of about 40nm as measured by a Nanosight (Amesbury, United Kingdom) LM 20 system and atomic force microscopy.

2.3.

Glass Substrate Pretreatment

Glass cover slips were cleaned by sonicating them subsequently in acetone, 1M NaOH, and Milli-Q water for one hour each. They were further rinsed with Milli-Q water three times and dried under nitrogen flow. The cover slips were then treated with freshly prepared 5mM APTMS (in toluene) for 10min and blown dry with nitrogen.

2.4.

Substrate Preparation

The stock solution of silver nanoparticles was dialyzed against Milli-Q water for 3h and was diluted with Milli-Q water to get a 0.2nM solution. A 10μL aliquot of this solution was deposited on an APTMS- treated glass cover slip and allowed to air dry at room temperature. The cover slip was then washed with Milli-Q water and blown dry with nitrogen flow. A 10μL aliquot of 10μM thrombin binding aptamer in PBS (with 100mM KCl added to induce the formation of the quadruplex structure) was added and incubated for 12h . It was rinsed with Milli-Q water and blown dry with nitrogen. A 0.1% aqueous solution of 6-mercaptohexanol was then added to block the surface not occupied by the aptamer. The substrate was rinsed with PBS, and 10μL of 10μM thrombin solution was added and incubated at 4°C for at least one hour. For the negative control, bovine serum albumin, a nonbinding protein, is used instead of the target molecule.

2.5.

Surface-Enhanced Raman Scattering Measurements

SERS spectra were acquired using a custom-built Raman system based on a Till Photonics microscope equipped with a 60× , 1.45-NA oil objective, via a SpectraPro 2300i Acton spectrometer with a Princeton Instruments Pixis100 back-illuminated charge-coupled device (CCD) camera, and 647-nm excitation wavelength from an Ar–Kr Innova 70C Coherent laser. An integration time of 60s and laser power of 100μW were used for all SERS measurements. The spectra were obtained by taking an average of five measurements from five different clusters/aggregates of nanoparticles. The baseline was corrected using a third-order polynomial fit.

3.

Results and Discussion

The use of aptamers as molecular receptors has been attracting a lot of attention for a number of reasons. First, they are less susceptible to denaturation and degradation than the corresponding antibodies. Second, they can be synthesized with high purity and reproducibility and are also easily engineered. Third, they are much smaller than the corresponding antibodies, which is significant in applications where a smaller size of the capture ligand is desired. This is represented in the present case, where the target molecule needs to be in close proximity to the metallic nanoparticle for significant SERS enhancement.

The methodology reported here offers several potential benefits: 1. it does not require biotin-streptavidin binding, as the aptamer can be attached covalently to the SERS substrate via thiolchemistry; 2. the SERS substrate functionalized with aptamer is robust, owing to the stability of the aptamer; and 3. it is also compact and could be further miniaturized to make it suitable for field applications using portable Raman spectrometers.

We have previously reported a Raman and SERS characterization study of both the thiolated and unthiolated analogs of the thrombin binding aptamer (TBA).47 We have shown that the formation of the quadruplex structure by TBA, indicated by the diagnostic 1480-cm1 band ( C8N7H deformation, Fig. 1)48, 49, 50 under favorable conditions, i.e., presence of K+ and incubation at 4°C , can be conveniently monitored by Raman spectroscopy. We also showed that the Raman and SERS spectra of the aptamer are dramatically different. This observation is not uncommon in SERS studies, and can be attributed to different enhancement factors of vibrational modes due to the fact that chemical bonds that are closer to the metallic nanoparticle surface receive stronger enhancement.

For this study, we have used a thiolated analog of the aptamer to facilitate binding of the aptamer to the nanoparticle. To show that thiolation does not prevent formation of the quadruplex structure, we previously measured the Raman spectra of both the thiolated and unthiolated analogs under the same conditions.47 Raman spectra of the two analogs both exhibit the characteristic 1480-cm1 band of the quadruplex structure, indicating that thiolation does not inhibit the formation of the quadruplex structure by the aptamer, which is responsible for its binding to thrombin. To ensure that the aptamer is in its quadruplex form before immobilization, it was prepared in PBS containing K+ ions.

Figures 3 and 3 present the individual SERS spectra of the thiolated aptamer (TBA) and the blocking agent mercaptohexanol (MH), respectively. It can be observed that the characteristic 1480-cm1 band of the quadruplex aptamer structure is missing in the SERS spectrum. This may suggest that the aptamer is in its unfolded form, but the absence of the band around 1496cm1 , which corresponds to the C8N7 vibration of guanine when it is not hydrogen bonded,50, 51 indicates otherwise. It is very likely that the guanine tetrads do not interact with the metal, since they are not exposed to it, i.e., their location in the quadruplex configuration does not allow for it. Another very noticeable feature of the SERS spectrum of TBA is the presence of a strong band around 1398cm1 . This can be attributed mostly to the CH2 deformation,52 which is shifted from the 1425-cm1 band of the spontaneous Raman spectrum,47 with some contribution from the ring-stretching mode of the guanine (1392cm1) .50 The intensity of the peak suggests there is proximity of this moiety to the metal surface.

Fig. 3

SERS spectra of (a) immobilized TBA, (b) blocking agent (MH), and (c) immobilized TBA together with MH. The spectra are normalized with respect to the most intense peak for each trace.

047006_1_032004jbo3.jpg

Figure 3 shows the SERS spectrum of TBA together with MH. It can be seen that the major SERS peaks corresponding to MH (1080 and 1435cm1 , C–C stretch and C–H deformation, respectively) added to the substrate already functionalized with the aptamer are detected together with those peaks corresponding to the aptamer (1005, 1245, and 1398cm1 ). The addition of the target protein (thrombin) resulted in additional bands in the spectra [Fig. 4 ], namely 1635cm1 (amide 1, CO stretch coupled with C–N–H bend),53, 54 1540cm1 (amide 2, N–H bend coupled with C–N stretch),55, 56 and 1140cm1 (C–N stretch),54 which are protein characteristic vibrational peaks (Table 1 ). In addition, there is a three-fold increase in the intensity of the C–H peak upon addition of thrombin. These observations strongly suggest binding of the target molecule. It may be noted that the amide 2 band is not normally Raman active. However, Raman enhancement near metallic nanoparticles can result in the appearance of vibrational modes that are normally forbidden by selection rules, as well as in the disappearance of certain vibrational bands. This is due to several factors, which include the proximity of the molecular bond to the metal surface and different orientation of the molecule with respect to the metal surface following surface adsorption or binding to capture agent. This may explain the appearance of this band at 1540cm1 . Similar observations have been reported in the literature. For example, Hu 56 reported the appearance of an amide 2 band (centered at 1526cm1 ) in the SERS spectrum of lysozyme, Kim 57 observed the presence of this band (at 1520cm1 ) in the SERS spectrum of Gly-L-Phe dipeptide, and Podstawka 58, 59 observed this band from 1504to1554cm1 in the SERS spectra of a series of homo- and heterodipeptides.

Fig. 4

(a) SERS of immobilized TBA together with MH (solid line) and TBA with MH and the target protein, thrombin (dashed line). The spectra normalized with respect to the 1398-cm1 TBA peak. (b) SERS spectra of immobilized TBA with MH (solid line) and immobilized TBA with MH and the nonbinding BSA (dashed line). The spectra are normalized with respect to the most intense peak (2920cm1) for both.

047006_1_032004jbo4.jpg

Table 1

SERS frequencies of TBA, TBA+MH , and TBA+MH+thrombin and their corresponding vibrational modes. The observed additional peaks on TBA+MH and TBA+MH+thrombin spectra are listed in bold for clarity.

SERS frequency, cm−1 AssignmentReferences
TBATBA +MH TBA+ MH+ thrombin
860860860sugar vib ( C3 endo),64
100510051005deoxyribose vib65
1080 (w)1080 (w)1080 (w) PO2 sym str in the TGT loop of TBA51
10801080C–C str in MH
1140C–N str of thrombin54
124512451245dT, N–H def, and C–N str of guanine66
127812781278ring str and C–H def of thymine67
1370 (w)1370 (w)1370 (w)dT, dG C2 endo/syn66
139813981398deoxyribosyl (C5H2) def52, 65, 66, 68
14351435C–H def of MH
1540Amide 2 of protein55, 56
158015801580 C2N3 of guanine
1635Amide I of protein53, 54
1660 (w)1660 (w)1660 (w) C6O6 (DNA)
292029202920deoxyribosyl C–H str47

These spectral changes that were observed with the addition of thrombin to the aptamer functionalized substrate were not detected when the same experiment was conducted using a nonbinding protein (BSA) [Fig. 4]. The spectra of TBA with MH and of TBA with MH and BSA are almost superimposable, suggesting that no binding has occurred. However, it can also be seen that there is a noticeable change in the spectrum of TBA with MH, specifically the relative intensities of their major peaks. One possible explanation for this observation is the variation in the amounts of aptamer and mercaptohexanol that were bound to the nanoparticle. The intensity ratio of 1398-cm11080-cm1 peak is noticeably higher in this set of spectra, suggesting an increased amount of aptamer adsorbed on the nanoparticle compared to the previous experiment (using thrombin). It is likely that this increase in the amount of aptamer molecules bound to the nanoparticle surface resulted in a higher relative intensity of the C–H peak. It is also possible that the higher concentration of bound aptamer led to a change in its orientation with respect to the nanoparticle surface, i.e., from flat or slightly bent to perpendicular, which in turn results in the observed higher relative intensity of the C–H peak.60 It has been shown both experimentally and theoretically that the C–H stretching mode is sensitive to molecule orientation on the nanoparticle surface.61, 62 In addition to the already mentioned changes in the spectrum of BSA-NP complex, we noted an additional band around 1518cm1 . One possible explanation for the appearance of this peak is nonspecific binding of BSA or other components (e.g., buffer components) of the test sample to the nanoparticle, resulting from insufficient coverage of the metal surface by both TBA and MH. This observation may indicate that nonspecific binding can be detected and discriminated against based on the different spectral response. While the sensitivity is not yet comparable with commonly used label-based and sandwiched-type detection methods such as ELISA, it is comparable with other direct detection techniques available in the literature, such as those reported by Jung,63 a surface acoustic wave (SAW)-based detection of thrombin (1μM) using an aptamer-functionalized sensor chip; and by Neumann,15 a SERS-based direct detection of target molecules (PDGF and cocaine), wherein the randomness of the SERS spectra of the specific aptamer following exposure to the corresponding target molecule ( 1μM PDGF and 20μM cocaine) is monitored. We believe that further improvements (e.g., preconcentration of the sample on the SERS substrate) to the current technique will enhance its sensitivity.

4.

Conclusions

We demonstrate a simple and selective SERS-based method for direct detection of biomolecules based on molecular recognition, which require fewer steps than traditional sandwiched-based assays. Binding of the target molecule to the molecular receptor (aptamer) is manifested by the appearance of the molecule’s characteristic vibrational bands (C–N stretching, amide 1 and amide 2 bands), in addition to the vibrational peaks corresponding to the receptor and blocking agent. It is important to note that interfering species that are able to bind may produce a noticeably different spectral response, thus further improving the selectivity.

Since aptamers are available not only for biomolecules but also for small molecules such as ricin, trinitrotoluene, and polycyclic aromatic hydrocarbons, this technique can be useful for the detection not only of medical biomarkers but also of environmental pollutants, as well as biological and chemical threat agents. In addition, the SERS substrates can be functionalized with different aptamers, allowing for multiplexed screening and detection. The simplicity of the developed label-free detection technique could be further utilized for on-chip and real-time detection by implementing it on a nano- and microscale platform.

Acknowledgments

This work is supported by funding from the National Science Foundation (NSF) and the Laboratory Directed Research and Development Program of Lawrence Livermore National Laboratory (LLNL). Cho acknowledges (LLNL-JRNL) for a Lawrence Scholar Program predoctoral fellowship (LLNL-JRNL-421298-DRAFT). The Center for Biophotonics Science and Technology is a designated NSF Science and Technology Center, managed by the University of California, Davis, under Cooperative Agreement number PHY 0120999. We would like to thank Rick Moerschell formerly of Biorad Incorporated for useful discussions.

References

1. 

Y. C. Cao, R. Jin, J. M. Nam, C. S. Thaxton, and C. A. Mirkin, “Raman dye-labeled nanoparticle probes for proteins,” J. Am. Chem. Soc., 125 (48), 14676 –14677 (2003). https://doi.org/10.1021/ja0366235 0002-7863 Google Scholar

2. 

J. M. Nam, C. S. Thaxton, and C. A. Mirkin, “Nanoparticle-based bio-bar codes for the ultrasensitive detection of proteins,” Science, 301 (5641), 1884 –1886 (2003). https://doi.org/10.1126/science.1088755 0036-8075 Google Scholar

3. 

X. X. Han, L. J. Cai, J. Guo, C. X. Wang, W. D. Ruan, W. Y. Han, W. Q. Xu, B. Zhao, and Y. Ozaki, “Fluorescein isothiocyanate linked immunoabsorbent assay based on surface-enhanced resonance Raman scattering,” Anal. Chem., 80 (8), 3020 –3024 (2008). https://doi.org/10.1021/ac702497t 0003-2700 Google Scholar

4. 

W. Zhao, M. A. Brook, and Y. Li, “Design of gold nanoparticle-based colorimetric biosensing assays,” ChemBioChem, 9 (15), 2363 –2371 (2008). https://doi.org/10.1002/cbic.200800282 1439-4227 Google Scholar

5. 

V. Pavlov, Y. Xiao, B. Shlyahovsky, and I. Willner, “Aptamer-functionalized Au nanoparticles for the amplified optical detection of thrombin,” J. Am. Chem. Soc., 126 (38), 11768 –11769 (2004). https://doi.org/10.1021/ja046970u 0002-7863 Google Scholar

6. 

M. Sha, S. Penn, G. Freeman, and W. Doering, “Detection of human viral RNA via a combined fluorescence and SERS molecular beacon assay,” NanoBiotechnology, 3 (1), 23 –30 (2007). https://doi.org/10.1007/s12030-007-0003-5 1551-1286 Google Scholar

7. 

B. Dubertret, M. Calame, and A. J. Libchaber, “Single-mismatch detection using gold-quenched fluorescent oligonucleotides,” Nat. Biotechnol., 19 (4), 365 –370 (2001). https://doi.org/10.1038/86762 1087-0156 Google Scholar

8. 

S. R. Kalb and J. R. Barr, “Mass spectrometric detection of ricin and its activity in food and clinical samples,” Anal. Chem., 81 (6), 2037 –2042 (2009). https://doi.org/10.1021/ac802769s 0003-2700 Google Scholar

9. 

W. E. Doering and S. Nie, “Single-molecule and single-nanoparticle SERS: examining the roles of surface active sites and chemical enhancement,” J. Phys. Chem. B, 106 (2), 311 –317 (2002). https://doi.org/10.1021/jp011730b 1089-5647 Google Scholar

10. 

M. Moskovits, “Surface-enhanced Raman spectroscopy: a brief retrospective,” J. Raman Spectrosc., 36 (6–7), 485 –496 (2005). https://doi.org/10.1002/jrs.1362 0377-0486 Google Scholar

11. 

A. Otto, “The ‘chemical’ (electronic) contribution to surface-enhanced Raman scattering,” J. Raman Spectrosc., 36 (6–7), 497 –509 (2005). https://doi.org/10.1002/jrs.1355 0377-0486 Google Scholar

12. 

K. Hering, D. Cialla, K. Ackermann, T. Dörfer, R. Möller, H. Schneidewind, R. Mattheis, W. Fritzsche, P. Rösch, and J. Popp, “SERS: a versatile tool in chemical and biochemical diagnostics,” Anal. Bioanal. Chem., 390 (1), 113 –124 (2008). https://doi.org/10.1007/s00216-007-1667-3 1618-2642 Google Scholar

13. 

M. D. Porter, R. J. Lipert, L. M. Siperko, G. Wang, and R. Narayanana, “SERS as a bioassay platform: fundamentals, design, and applications,” Chem. Soc. Rev., 37 (5), 1001 –1011 (2008). https://doi.org/10.1039/b708461g 0306-0012 Google Scholar

14. 

H. Cho, B. R. Baker, S. Wachsmann-Hogiu, C. V. Pagba, T. A. Laurence, S. M. Lane, L. P. Lee, and J. B. H. Tok, “Aptamer-based SERRS sensor for thrombin detection,” Nano Lett., 8 (12), 4386 –4390 (2008). https://doi.org/10.1021/nl802245w 1530-6984 Google Scholar

15. 

O. Neumann, D. Zhang, F. Tam, S. Lal, P. Wittung-Stafshede, and N. J. Halas, “Direct optical detection of aptamer conformational changes induced by target molecules,” Anal. Chem., 81 (24), 10002 –10006 (2009). https://doi.org/10.1021/ac901849k 0003-2700 Google Scholar

16. 

A. D. Ellington and J. W. Szostak, “In vitro selection of RNA molecules that bind specific ligands,” Nature, 346 (6287), 818 –822 (1990). https://doi.org/10.1038/346818a0 0028-0836 Google Scholar

17. 

N. de-los-Santos-Álvarez, M. A. J. Lobo-Castañón, A. J. Miranda-Ordieres, and P. Tuñón-Blanco, “Aptamers as recognition elements for label-free analytical devices,” TrAC, Trends Anal. Chem., 27 (5), 437 –446 (2008). https://doi.org/10.1016/j.trac.2008.03.003 0165-9936 Google Scholar

18. 

T. Hermann and D. J. Patel, “Adaptive recognition by nucleic acid aptamers,” Science, 287 (5454), 820 –825 (2000). https://doi.org/10.1126/science.287.5454.820 0036-8075 Google Scholar

19. 

J. X. N. S. Y. Y. Jijun Tang, “The DNA aptamers that specifically recognize ricin toxin are selected by two in vitro selection methods,” Electrophoresis, 27 (7), 1303 –1311 (2006). https://doi.org/10.1002/elps.200500489 0173-0835 Google Scholar

20. 

J. G. Bruno and J. L. Kiel, “In vitro selection of DNA aptamers to anthrax spores with electrochemiluminescence detection,” Biosens. Bioelectron., 14 (5), 457 –464 (1999). https://doi.org/10.1016/S0956-5663(99)00028-7 0956-5663 Google Scholar

21. 

E. Ehrentreich-Förster, D. Orgel, A. Krause-Griep, B. Cech, V. Erdmann, F. Bier, F. Scheller, and M. Rimmele, “Biosensor-based on-site explosives detection using aptamers as recognition elements,” Anal. Bioanal. Chem., 391 (5), 1793 –1800 (2008). https://doi.org/10.1007/s00216-008-2150-5 1618-2642 Google Scholar

22. 

B. J. Hicke, C. Marion, Y. F. Chang, T. Gould, C. K. Lynott, D. Parma, P. G. Schmidt, and S. Warren, “Tenascin-C aptamers are generated using tumor cells and purified protein,” J. Biol. Chem., 276 (52), 48644 –48654 (2001). https://doi.org/10.1074/jbc.M104651200 0021-9258 Google Scholar

23. 

C. Wang, M. Zhang, G. Yang, D. Zhang, H. Ding, H. Wang, M. Fan, B. Shen, and N. Shao, “Single-stranded DNA aptamers that bind differentiated but not parental cells: subtractive systematic evolution of ligands by exponential enrichment,” J. Biotechnol., 102 (1), 15 –22 (2003). https://doi.org/10.1016/S0168-1656(02)00360-7 0168-1656 Google Scholar

24. 

J. Ruckman, L. S. Green, J. Beeson, S. Waugh, W. L. Gillette, D. D. Henninger, L. Claesson-Welsh, and N. Janjic, “2-fluoropyrimidine RNA-based aptamers to the 165-amino acid form of vascular endothelial growth factor (VEGF165). Inhibition of receptor binding and vegf-induced vascular permeability through interactions requiring the exon 7-encoded domain,” J. Biol. Chem., 273 (32), 20556 –20567 (1998). https://doi.org/10.1074/jbc.273.32.20556 0021-9258 Google Scholar

25. 

C. H. B. Chen, G. A. Chernis, V. Q. Hoang, and R. Landgraf, “Inhibition of heregulin signaling by an aptamer that preferentially binds to the oligomeric form of human epidermal growth factor receptor-3,” Proc. Natl. Acad. Sci. U.S.A., 100 (16), 9226 –9231 (2003). https://doi.org/10.1073/pnas.1332660100 0027-8424 Google Scholar

26. 

N. Sayer, J. Ibrahim, K. Turner, A. Tahiri-Alaoui, and W. James, “Structural characterization of a 2F-RNA aptamer that binds a HIV-1 SU glycoprotein, gp120,” Biochem. Biophys. Res. Commun., 293 (3), 924 –931 (2002). https://doi.org/10.1016/S0006-291X(02)00308-X 0006-291X Google Scholar

27. 

R. Yamamoto and P. K. R. Kumar, “Molecular beacon aptamer fluoresces in the presence of Tat protein of HIV-1,” Genes Cells, 5 (5), 389 –396 (2000). https://doi.org/10.1046/j.1365-2443.2000.00331.x 1356-9597 Google Scholar

28. 

O. Kensch, B. A. Connolly, H. J. r. Steinhoff, A. McGregor, R. S. Goody, and T. Restle, “HIV-1 reverse transcriptase-pseudoknot RNA aptamer interaction has a binding affinity in the low picomolar range coupled with high specificity,” J. Biol. Chem., 275 (24), 18271 –18278 (2000). https://doi.org/10.1074/jbc.M001309200 0021-9258 Google Scholar

29. 

N. Jing, R. F. Rando, Y. Pommier, and M. E. Hogan, “Ion selective folding of loop domains in a potent anti-HIV oligonucleotide,” Biochemistry, 36 (41), 12498 –12505 (1997). https://doi.org/10.1021/bi962798y 0006-2960 Google Scholar

30. 

S. Tombelli, M. Minunni, and M. Mascini, “Aptamers-based assays for diagnostics, environmental and food analysis,” Biomol. Eng., 24 (2), 191 –200 (2007). https://doi.org/10.1016/j.bioeng.2007.03.003 1389-0344 Google Scholar

31. 

W. Yoshida, E. Mochizuki, M. Takase, H. Hasegawa, Y. Morita, H. Yamazaki, K. Sode, and K. Ikebukuro, “Selection of DNA aptamers against insulin and construction of an aptameric enzyme subunit for insulin sensing,” Biosens. Bioelectron., 24 (5), 1116 –1120 (2009). https://doi.org/10.1016/j.bios.2008.06.016 0956-5663 Google Scholar

32. 

A. C. Connor, K. A. Frederick, E. J. Morgan, and L. B. McGown, “Insulin capture by an insulin-linked polymorphic region G-quadruplex DNA oligonucleotide,” J. Am. Chem. Soc., 128 (15), 4986 –4991 (2006). https://doi.org/10.1021/ja056097c 0002-7863 Google Scholar

33. 

T. Wiegand, P. Williams, S. Dreskin, M. Jouvin, J. Kinet, and D. Tasset, “High-affinity oligonucleotide ligands to human IgE inhibit binding to Fc epsilon receptor I,” J. Immunol., 157 (1), 221 –230 (1996). 0022-1767 Google Scholar

34. 

L. C. Bock, L. C. Griffin, J. A. Latham, E. H. Vermaas, and J. J. Toole, “Selection of single-stranded DNA molecules that bind and inhibit human thrombin,” Nature, 355 (6360), 564 –566 (1992). https://doi.org/10.1038/355564a0 0028-0836 Google Scholar

35. 

S. Song, L. Wang, J. Li, C. Fan, and J. Zhao, “Aptamer-based biosensors,” TrAC, Trends Anal. Chem., 27 (2), 108 –117 (2008). https://doi.org/10.1016/j.trac.2007.12.004 0165-9936 Google Scholar

36. 

T. Mairal, V. Cengiz Özalp, P. Lozano Sánchez, M. Mir, I. Katakis, and C. O’Sullivan, “Aptamers: molecular tools for analytical applications,” Anal. Bioanal. Chem., 390 (4), 989 –1007 (2008). https://doi.org/10.1007/s00216-007-1346-4 1618-2642 Google Scholar

37. 

S. Tombelli, M. Minunni, and M. Mascini, “Analytical applications of aptamers,” Biosens. Bioelectron., 20 (12), 2424 –2434 (2005). https://doi.org/10.1016/j.bios.2004.11.006 0956-5663 Google Scholar

38. 

M. Famulok, J. S. Hartig, and G. Mayer, “Functional aptamers and aptazymes in biotechnology, diagnostics, and therapy,” Chem. Rev., 107 (9), 3715 –3743 (2007). https://doi.org/10.1021/cr0306743 0009-2665 Google Scholar

39. 

L. R. Paborsky, S. N. McCurdy, L. C. Griffin, J. J. Toole, and L. L. Leung, “The single-stranded DNA aptamer-binding site of human thrombin,” J. Biol. Chem., 268 (28), 20808 –20811 (1993). 0021-9258 Google Scholar

40. 

K. Y. Wang, S. McCurdy, R. G. Shea, S. Swaminathan, and P. H. Bolton, “A DNA aptamer which binds to and inhibits thrombin exhibits a new structural motif for DNA,” Biochemistry, 32 (8), 1899 –1904 (1993). https://doi.org/10.1021/bi00059a003 0006-2960 Google Scholar

41. 

K. Y. Wang, S. H. Krawczyk, N. Bischofberger, S. Swaminathan, and P. H. Bolton, “The tertiary structure of a DNA aptamer which binds to and inhibits thrombin determines activity,” Biochemistry, 32 (42), 11285 –11292 (1993). https://doi.org/10.1021/bi00093a004 0006-2960 Google Scholar

42. 

R. F. Macaya, P. Schultze, F. W. Smith, J. A. Roe, and J. Feigon, “Thrombin-binding DNA aptamer forms a unimolecular quadruplex structure in solution,” Proc. Natl. Acad. Sci. U.S.A., 90 (8), 3745 –3749 (1993). https://doi.org/10.1073/pnas.90.8.3745 0027-8424 Google Scholar

43. 

P. Schultze, R. F. Macaya, and J. Feigon, “Three-dimensional solution structure of the thrombin-binding DNA aptamer d(GGTTGGTGTGGTTGG),” J. Mol. Biol., 235 (5), 1532 –1547 (1994). https://doi.org/10.1006/jmbi.1994.1105 0022-2836 Google Scholar

44. 

K. Padmanabhan, K. P. Padmanabhan, J. D. Ferrara, J. E. Sadler, and A. Tulinsky, “The structure of alpha-thrombin inhibited by a 15-mer single-stranded DNA aptamer,” J. Biol. Chem., 268 (24), 17651 –17654 (1993). 0021-9258 Google Scholar

45. 

P. Alberti and J. L. Mergny, “DNA duplex-quadruplex exchange as the basis for a nanomolecular machine,” Proc. Natl. Acad. Sci. U.S.A., 100 (4), 1569 –1573 (2003). https://doi.org/10.1073/pnas.0335459100 0027-8424 Google Scholar

46. 

P. C. Lee and D. Meisel, “Adsorption and surface-enhanced Raman of dyes on silver and gold sols,” J. Phys. Chem., 86 (17), 3391 –3395 (1982). https://doi.org/10.1021/j100214a025 0022-3654 Google Scholar

47. 

C. V. Pagba, S. M. Lane, and S. Wachsmann-Hogiu, “Raman and surface-enhanced Raman spectroscopic studies of the 15-mer DNA thrombin-binding aptamer,” J. Raman Spectrosc., 41 (3), 241 –247 (2010). 0377-0486 Google Scholar

48. 

T. Miura and G. J. Thomas, “Structural polymorphism of telomere DNA: interquadruplex and duplex-quadruplex conversions probed by Raman spectroscopy,” Biochemistry, 33 (25), 7848 –7856 (1994). https://doi.org/10.1021/bi00191a012 0006-2960 Google Scholar

49. 

J. M. Benevides, A. H. J. Wang, G. A. Van der Marel, J. H. Van Boom, and G. J. Thomas, “Crystal and solution structures of the B-DNA dodecamer d(CGCAAATTTGCG) probed by Raman spectroscopy: heterogeneity in the crystal structure does not persist in the solution structure,” Biochemistry, 27 (3), 931 –938 (1988). https://doi.org/10.1021/bi00403a014 0006-2960 Google Scholar

50. 

Y. Nishimura, M. Tsuboi, T. Sato, and K. Aoki, “Conformation-sensitive Raman lines of mononucleotides and their use in a structure analysis of polynucleotides: guanine and cytosine nucleotides,” J. Mol. Struct., 146 123 –153 (1986). https://doi.org/10.1016/0022-2860(86)80288-5 0022-2860 Google Scholar

51. 

J. A. Mondragon-Sanchez, J. Liquier, R. H. Shafer, and E. Taillandier, “Tetraplex structure formation in the thrombin-binding DNA aptamer by metal cations measured by vibrational spectroscopy,” J. Biomol. Struct. Dyn., 22 (3), 365 –373 (2004). 0739-1102 Google Scholar

52. 

K. Kneipp and J. Flemming, “Surface enhanced Raman scattering (SERS) of nucleic acids adsrobed on colloidal silver particles,” J. Mol. Struct., 145 (1–2), 173 –179 (1986). https://doi.org/10.1016/0022-2860(86)87041-7 0022-2860 Google Scholar

53. 

W. L. Peticolas, “Raman spectroscopy of DNA and proteins,” Methods Enzymol., 246 389 –416 (1995). https://doi.org/10.1016/0076-6879(95)46019-5 0076-6879 Google Scholar

54. 

M. C. Chen, R. C. Lord, and R. Mendelsohn, “Laser-excited Raman spectroscopy of biomolecules. V. Conformational changes associated with the chemical denaturation of lysozyme,” J. Am. Chem. Soc., 96 (10), 3038 –3042 (1974). https://doi.org/10.1021/ja00817a003 0002-7863 Google Scholar

55. 

Y. P. Zhang, R. N. A. H. Lewis, R. S. Hodges, and R. N. McElhaney, “FTIR spectroscopic studies of the conformation and amide hydrogen exchange of a peptide model of the hydrophobic transmembrane .alpha.-helixes of membrane proteins,” Biochemistry, 31 (46), 11572 –11578 (1992). https://doi.org/10.1021/bi00161a041 0006-2960 Google Scholar

56. 

J. Hu, R. S. Sheng, Z. S. Xu, and Y. Zeng, “Surface enhanced Raman spectroscopy of lysozyme,” Spectrochim. Acta, Part A, 51 (6), 1087 –1096 (1995). https://doi.org/10.1016/0584-8539(94)00225-Z 0584-8539 Google Scholar

57. 

S. K. Kim, M. S. Kim, and S. W. Suh, “Surface-enhanced Raman scattering (SERS) of aromatic amino acids and their glycyl dipeptides in silver sol,” J. Raman Spectrosc., 18 (3), 171 –175 (1987). https://doi.org/10.1002/jrs.1250180305 0377-0486 Google Scholar

58. 

E. Podstawka, Y. Ozaki, and L. M. Proniewicz, “Part I: surface-enhanced Raman spectroscopy investigation of amino acids and their homodipeptides adsorbed on colloidal silver,” Appl. Spectrosc., 58 (5), 570 –580 (2004). https://doi.org/10.1366/000370204774103408 0003-7028 Google Scholar

59. 

E. Podstawka, Y. Ozaki, and L. M. Proniewicz, “Part II: surface-enhanced Raman spectroscopy investigation of methionine containing heterodipeptides adsorbed on colloidal silver,” Appl. Spectrosc., 58 (5), 581 –590 (2004). https://doi.org/10.1366/000370204774103417 0003-7028 Google Scholar

60. 

A. Barhoumi, D. Zhang, F. Tam, and N. J. Halas, “Surface-enhanced Raman spectroscopy of DNA,” J. Am. Chem. Soc., 130 (16), 5523 –5529 (2008). https://doi.org/10.1021/ja800023j 0002-7863 Google Scholar

61. 

M. Moskovits and J. S. Suh, “Surface selection rules for surface-enhanced Raman spectroscopy: calculations and application to the surface-enhanced Raman spectrum of phthalazine on silver,” J. Phys. Chem., 88 (23), 5526 –5530 (1984). https://doi.org/10.1021/j150667a013 0022-3654 Google Scholar

62. 

X. Gao, J. P. Davies, and M. J. Weaver, “Test of surface selection rules for surface-enhanced Raman scattering: the orientation of adsorbed benzene and monosubstituted benzenes on gold,” J. Phys. Chem., 94 (17), 6858 –6864 (1990). https://doi.org/10.1021/j100380a059 0022-3654 Google Scholar

63. 

A. Jung, T. M. A. Gronewold, M. Tewes, E. Quandt, and P. Berlin, “Biofunctional structural design of SAW sensor chip surfaces in a microfluidic sensor system,” Sens. Actuators B, 124 (1), 46 –52 (2007). https://doi.org/10.1016/j.snb.2006.11.040 0925-4005 Google Scholar

64. 

W. L. Peticolas and E. Evertsz, “Conformation of DNA in vitro and in vivo from laser Raman scattering,” Methods Enzymol., 211 335 –352 (1992). https://doi.org/10.1016/0076-6879(92)11019-F 0076-6879 Google Scholar

65. 

“Characterization of DNA structures by laser Raman spectroscopy,” Biopolymers, 23 (2), 235 –256 (1984). https://doi.org/10.1002/bip.360230206 0006-3525 Google Scholar

66. 

C. Krafft, J. M. Benevides, and G. J. Thomas Jr., “Secondary structure polymorphism in Oxytricha nova telomeric DNA,” Nucleic Acids Res., 30 (18), 3981 –3991 (2002). https://doi.org/10.1093/nar/gkf517 0305-1048 Google Scholar

67. 

C. Otto, T. J. J. van den Tweel, F. F. M. de Mul, and J. Greve, “Surface-enhanced Raman spectroscopy of DNA bases,” J. Raman Spectrosc., 17 (3), 289 –298 (1986). https://doi.org/10.1002/jrs.1250170311 0377-0486 Google Scholar

68. 

J. M. Benevides and G. J. Thomas Jr., “Characterization of DNA structures by Raman spectroscopy: high-salt and low-salt forms of double helical poly(dG-dC) in H2O and D2O solutions and application to B, Z and A-DNA*,” Nucleic Acids Res., 11 (16), 5747 –5761 (1983). https://doi.org/10.1093/nar/11.16.5747 0305-1048 Google Scholar
©(2010) Society of Photo-Optical Instrumentation Engineers (SPIE)
Cynthia V. Pagba, Stephen M. Lane, Hansang Cho, and Sebastian Wachsmann-Hogiu "Direct detection of aptamer-thrombin binding via surface-enhanced Raman spectroscopy," Journal of Biomedical Optics 15(4), 047006 (1 July 2010). https://doi.org/10.1117/1.3465594
Published: 1 July 2010
Lens.org Logo
CITATIONS
Cited by 45 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Surface enhanced Raman spectroscopy

Molecules

Nanoparticles

Proteins

Raman spectroscopy

Metals

Receptors

Back to Top