Single-molecule experiment

Last updated
Single polymer molecules (0.4 nm thick chains) recorded under aqueous media at different pH using an AFM. Drastic change of polymer chain conformation is observed at a small change of pH. Single-Molecule-Under-Water-AFM-Tapping-Mode.jpg
Single polymer molecules (0.4 nm thick chains) recorded under aqueous media at different pH using an AFM. Drastic change of polymer chain conformation is observed at a small change of pH.

A single-molecule experiment is an experiment that investigates the properties of individual molecules. Single-molecule studies may be contrasted with measurements on an ensemble or bulk collection of molecules, where the individual behavior of molecules cannot be distinguished, and only average characteristics can be measured. Since many measurement techniques in biology, chemistry, and physics are not sensitive enough to observe single molecules, single-molecule fluorescence techniques (that have emerged since the 1990s for probing various processes on the level of individual molecules) caused a lot of excitement, since these supplied many new details on the measured processes that were not accessible in the past. Indeed, since the 1990s, many techniques for probing individual molecules have been developed. [2]

Contents

The first single-molecule experiments were patch clamp experiments performed in the 1970s, but these were limited to studying ion channels. Today, systems investigated using single-molecule techniques include the movement of myosin on actin filaments in muscle tissue and the spectroscopic details of individual local environments in solids. Biological polymers' conformations have been measured using atomic force microscopy (AFM). Using force spectroscopy, single molecules (or pairs of interacting molecules), usually polymers, can be mechanically stretched, and their elastic response recorded in real time.

History

In the gas phase at ultralow pressures, single-molecule experiments have been around for decades, but in the condensed phase only since 1989 with the work by W. E. Moerner and Lothar Kador. [3] One year later, Michel Orrit and Jacky Bernard were able to show also the detection of the absorption of single molecules by their fluorescence. [4]

Many techniques have the ability to observe one molecule at a time, most notably mass spectrometry, where single ions are detected. In addition, one of the earliest means of detecting single molecules, came about in the field of ion channels with the development of the patch clamp technique by Erwin Neher and Bert Sakmann (who later went on to win the Nobel prize for their seminal contributions). However, the idea of measuring conductance to look at single molecules placed a serious limitation on the kind of systems which could be observed.

Fluorescence is a convenient means of observing one molecule at a time, mostly due to the sensitivity of commercial optical detectors, capable of counting single photons. However, spectroscopically, the observation of one molecule requires that the molecule is in an isolated environment and that it emits photons upon excitation, which owing to the technology to detect single photons by use of photomultiplier tubes (PMT) or avalanche photodiodes (APD), enables one to record photon emission events with great sensitivity and time resolution.

More recently, single-molecule fluorescence is the subject of intense interest for biological imaging, through the labeling of biomolecules such as proteins and nucleotides to study enzymatic function which cannot easily be studied on the bulk scale, due to subtle time-dependent movements in catalysis and structural reorganization. The most studied protein has been the class of myosin/actin enzymes found in muscle tissues. Through single-molecule techniques the step mechanism has been observed and characterized in many of these proteins.

In 1997, single-molecule detection was demonstrated with surface-enhanced Raman spectroscopy (SERS) by Katrin Kneipp, H. Kneipp, Y. Wang, L.T. Perelman and others [5] at MIT and independently by S. Nie and S. R. Emory at Indiana University. [6] The MIT team used non-resonance Raman excitation and surface enhancement with silver nanoclusters to detect single cresyl violet molecules, while the team at Indiana University used resonance Raman excitation and surface enhancement with silver nanoparticles to detect single rhodamine 6G molecules.

Nanomanipulators such as the atomic force microscope are also suited to single-molecule experiments of biological significance, since they work on the same length scale of most biological polymers. Besides, atomic force microscopy (AFM) is appropriate for the studies of synthetic polymer molecules. AFM provides a unique possibility of 3D visualization of polymer chains. For instance, AFM tapping mode is gentle enough for the recording of adsorbed polyelectrolyte molecules (for example, 0.4 nm thick chains of poly(2-vinylpyridine)) under liquid medium. The location of two-chain-superposition correspond in these experiments to twice the thickness of single chain (0.8 nm in the case of the mentioned example). At the application of proper scanning parameters, conformation of such molecules remain unchanged for hours that allows the performance of experiments under liquid media having various properties. [1] Furthermore, by controlling the force between the tip and the sample high resolution images can be obtained. [7] [8] Optical tweezers have also been used to study and quantify DNA-protein interactions. [7] [8]

About the experiments

Concept

Single-molecule fluorescence spectroscopy uses the fluorescence of a molecule for obtaining information on its environment, structure, and position. The technique affords the ability of obtaining information otherwise not available due to ensemble averaging (that is, a signal obtained when recording many molecules at the same time represents an average property of the molecules' dynamics). The results in many experiments of individual molecules are two-state trajectories.

Single-channel recording

Two examples of data (~10 s each) from a single-channel recording experiment using the patch clamp technique. The upper trace is at a lower concentration, while the lower trace is recorded at an agonist concentration near this ion channel's EC50. Channel openings are represented by downward deflections, in this case deflections of approximately 5 pA. SingleMoleculeData IonChannel TwoConcentrations.png
Two examples of data (~10 s each) from a single-channel recording experiment using the patch clamp technique. The upper trace is at a lower concentration, while the lower trace is recorded at an agonist concentration near this ion channel's EC50. Channel openings are represented by downward deflections, in this case deflections of approximately 5 pA.

As in the case of single molecule fluorescence spectroscopy, the technique known as single channel recording can be used to obtain specific kinetic information—in this case about ion channel function—that is not available when ensemble recording, such as whole-cell recording, is performed. [9] Specifically, ion channels alternate between conducting and non-conducting classes, which differ in conformation. Therefore, the functional state of ion channels can be directly measured with sufficiently sensitive electronics, provided that proper precautions are taken to minimize noise. In turn, each of these classes may be divided into one or more kinetic states with direct bearing on the underlying function of the ion channel. Performing these types of single molecule studies under systematically varying conditions (e.g. agonist concentration and structure, permeant ion and/or channel blocker, mutations in the ion channel amino acids), can provide information regarding the interconversion of various kinetic states of the ion channel. [10] In a minimal model for an ion channel, there are two states: open and closed. However, other states are often needed in order to accurately represent the data, including multiple closed states as well as inactive and/or desensitized states, which are non-conducting states that can occur even in the presence of stimulus. [9]

Biomolecule labeling

Single fluorophores can be chemically attached to biomolecules, such as proteins or DNA, and the dynamics of individual molecules can be tracked by monitoring the fluorescent probe. Spatial movements within the Rayleigh limit can be tracked, along with changes in emission intensity and/or radiative lifetime, which often indicate changes in local environment. For instance, single-molecule labeling has yielded a vast quantity of information on how kinesin motor proteins move along microtubule strands in muscle cells. Single-molecule imaging in live cells reveals interesting information about protein dynamics under its physiological environment. Several biophysical parameters about protein dynamics can be quantified such as diffusion coefficient, mean squared displacements, residence time, the fraction of bound and unbound molecules, and target-search mechanism of protein binding to its target site in the live cell. [11]

Single-molecule fluorescence resonance energy transfer (FRET)

Main article smFRET.

In single-molecule fluorescence resonance energy transfer, the molecule is labeled in (at least) two places. A laser beam is focused on the molecule exciting the first probe. When this probe relaxes and emits a photon, it has a chance of exciting the other probe. The efficiency of the absorption of the photon emitted from the first probe in the second probe depends on the distance between these probes. Since the distance changes with time, this experiment probes the internal dynamics of the molecule.

Versus ensemble experiments

When looking at data related to individual molecules, one usually can construct propagators, and jumping time probability density functions, of the first order, the second order and so on, whereas from bulk experiments, one usually obtains the decay of a correlation function. [12] From the information contained in these unique functions (obtained from individual molecules), one can extract a relatively clear picture on the way the system behaves; e.g. its kinetic scheme, or its potential of activity, or its reduced dimensions form. [13] [14] In particular, one can construct (many properties of) the reaction pathway of an enzyme when monitoring the activity of an individual enzyme. [15] Additionally, significant aspects regarding the analysis of single molecule data—such as fitting methods and tests for homogeneous populations—have been described by several authors. [9] On the other hand, there are several issues with the analysis of single molecule data including construction of a low noise environment and insulated pipet tips, filtering some of the remaining unwanted components (noise) found in recordings, and the length of time required for data analysis (pre-processing, unambiguous event detection, plotting data, fitting kinetic schemes, etc.).

Impact

Single-molecule techniques impacted optics, electronics, biology, and chemistry. In the biological sciences, the study of proteins and other complex biological machinery was limited to ensemble experiments that nearly made impossible the direct observation of their kinetics. For example, it was only after single molecule fluorescence microscopy was used to study kinesin-myosin pairs in muscle tissue that direct observation of the walking mechanisms were understood. These experiments, however, have for the most part been limited to in vitro studies, as useful techniques for live cell imaging have yet to be fully realized. The promise of single molecule in vivo imaging, [16] however, brings with it an enormous potential to directly observe bio-molecules in native processes. These techniques are often targeted for studies involving low-copy proteins, many of which are still being discovered. These techniques have also been extended to study areas of chemistry, including the mapping of heterogeneous surfaces. [17]

See also

Related Research Articles

<span class="mw-page-title-main">Spectroscopy</span> Study involving matter and electromagnetic radiation

Spectroscopy is the field of study that measures and interprets electromagnetic spectra. In narrower contexts, spectroscopy is the precise study of color as generalized from visible light to all bands of the electromagnetic spectrum.

<span class="mw-page-title-main">Raman spectroscopy</span> Spectroscopic technique

Raman spectroscopy is a spectroscopic technique typically used to determine vibrational modes of molecules, although rotational and other low-frequency modes of systems may also be observed. Raman spectroscopy is commonly used in chemistry to provide a structural fingerprint by which molecules can be identified.

Force spectroscopy is a set of techniques for the study of the interactions and the binding forces between individual molecules. These methods can be used to measure the mechanical properties of single polymer molecules or proteins, or individual chemical bonds. The name "force spectroscopy", although widely used in the scientific community, is somewhat misleading, because there is no true matter-radiation interaction.

In physics and physical chemistry, time-resolved spectroscopy is the study of dynamic processes in materials or chemical compounds by means of spectroscopic techniques. Most often, processes are studied after the illumination of a material occurs, but in principle, the technique can be applied to any process that leads to a change in properties of a material. With the help of pulsed lasers, it is possible to study processes that occur on time scales as short as 10−16 seconds. All time-resolved spectra are suitable to be analyzed using the two-dimensional correlation method for a correlation map between the peaks.

Coherent anti-Stokes Raman spectroscopy, also called Coherent anti-Stokes Raman scattering spectroscopy (CARS), is a form of spectroscopy used primarily in chemistry, physics and related fields. It is sensitive to the same vibrational signatures of molecules as seen in Raman spectroscopy, typically the nuclear vibrations of chemical bonds. Unlike Raman spectroscopy, CARS employs multiple photons to address the molecular vibrations, and produces a coherent signal. As a result, CARS is orders of magnitude stronger than spontaneous Raman emission. CARS is a third-order nonlinear optical process involving three laser beams: a pump beam of frequency ωp, a Stokes beam of frequency ωS and a probe beam at frequency ωpr. These beams interact with the sample and generate a coherent optical signal at the anti-Stokes frequency (ωprpS). The latter is resonantly enhanced when the frequency difference between the pump and the Stokes beams (ωpS) coincides with the frequency of a Raman resonance, which is the basis of the technique's intrinsic vibrational contrast mechanism.

Fluorescence correlation spectroscopy (FCS) is a statistical analysis, via time correlation, of stationary fluctuations of the fluorescence intensity. Its theoretical underpinning originated from L. Onsager's regression hypothesis. The analysis provides kinetic parameters of the physical processes underlying the fluctuations. One of the interesting applications of this is an analysis of the concentration fluctuations of fluorescent particles (molecules) in solution. In this application, the fluorescence emitted from a very tiny space in solution containing a small number of fluorescent particles (molecules) is observed. The fluorescence intensity is fluctuating due to Brownian motion of the particles. In other words, the number of the particles in the sub-space defined by the optical system is randomly changing around the average number. The analysis gives the average number of fluorescent particles and average diffusion time, when the particle is passing through the space. Eventually, both the concentration and size of the particle (molecule) are determined. Both parameters are important in biochemical research, biophysics, and chemistry.

Ultrafast laser spectroscopy is a category of spectroscopic techniques using ultrashort pulse lasers for the study of dynamics on extremely short time scales. Different methods are used to examine the dynamics of charge carriers, atoms, and molecules. Many different procedures have been developed spanning different time scales and photon energy ranges; some common methods are listed below.

Chemical imaging is the analytical capability to create a visual image of components distribution from simultaneous measurement of spectra and spatial, time information. Hyperspectral imaging measures contiguous spectral bands, as opposed to multispectral imaging which measures spaced spectral bands.

Lipid bilayer characterization is the use of various optical, chemical and physical probing methods to study the properties of lipid bilayers. Many of these techniques are elaborate and require expensive equipment because the fundamental nature of the lipid bilayer makes it a very difficult structure to study. An individual bilayer, since it is only a few nanometers thick, is invisible in traditional light microscopy. The bilayer is also a relatively fragile structure since it is held together entirely by non-covalent bonds and is irreversibly destroyed if removed from water. In spite of these limitations dozens of techniques have been developed over the last seventy years to allow investigations of the structure and function of bilayers. The first general approach was to utilize non-destructive in situ measurements such as x-ray diffraction and electrical resistance which measured bilayer properties but did not actually image the bilayer. Later, protocols were developed to modify the bilayer and allow its direct visualization at first in the electron microscope and, more recently, with fluorescence microscopy. Over the past two decades, a new generation of characterization tools including AFM has allowed the direct probing and imaging of membranes in situ with little to no chemical or physical modification. More recently, dual polarisation interferometry has been used to measure the optical birefringence of lipid bilayers to characterise order and disruption associated with interactions or environmental effects.

<span class="mw-page-title-main">Raman microscope</span> Laser microscope used for Raman spectroscopy

The Raman microscope is a laser-based microscopic device used to perform Raman spectroscopy. The term MOLE is used to refer to the Raman-based microprobe. The technique used is named after C. V. Raman, who discovered the scattering properties in liquids.

<span class="mw-page-title-main">Fluorescence in the life sciences</span> Scientific investigative technique

Fluorescence is used in the life sciences generally as a non-destructive way of tracking or analysing biological molecules. Some proteins or small molecules in cells are naturally fluorescent, which is called intrinsic fluorescence or autofluorescence. Alternatively, specific or general proteins, nucleic acids, lipids or small molecules can be "labelled" with an extrinsic fluorophore, a fluorescent dye which can be a small molecule, protein or quantum dot. Several techniques exist to exploit additional properties of fluorophores, such as fluorescence resonance energy transfer, where the energy is passed non-radiatively to a particular neighbouring dye, allowing proximity or protein activation to be detected; another is the change in properties, such as intensity, of certain dyes depending on their environment allowing their use in structural studies.

Single-molecule fluorescence resonance energy transfer is a biophysical technique used to measure distances at the 1-10 nanometer scale in single molecules, typically biomolecules. It is an application of FRET wherein a pair of donor and acceptor fluorophores are excited and detected at a single molecule level. In contrast to "ensemble FRET" which provides the FRET signal of a high number of molecules, single-molecule FRET is able to resolve the FRET signal of each individual molecule. The variation of the smFRET signal is useful to reveal kinetic information that an ensemble measurement cannot provide, especially when the system is under equilibrium with no ensemble/bulk signal change. Heterogeneity among different molecules can also be observed. This method has been applied in many measurements of intramolecular dynamics such as DNA/RNA/protein folding/unfolding and other conformational changes, and intermolecular dynamics such as reaction, binding, adsorption, and desorption that are particularly useful in chemical sensing, bioassays, and biosensing.

<span class="mw-page-title-main">Reduced dimensions form</span>

In biophysics and related fields, reduced dimension forms (RDFs) are unique on-off mechanisms for random walks that generate two-state trajectories (see Fig. 1 for an example of a RDF and Fig. 2 for an example of a two-state trajectory). It has been shown that RDFs solve two-state trajectories, since only one RDF can be constructed from the data, where this property does not hold for on-off kinetic schemes, where many kinetic schemes can be constructed from a particular two-state trajectory (even from an ideal on-off trajectory). Two-state time trajectories are very common in measurements in chemistry, physics, and the biophysics of individual molecules (e.g. measurements of protein dynamics and DNA and RNA dynamics, activity of ion channels, enzyme activity, quantum dots ), thus making RDFs an important tool in the analysis of data in these fields.

The technique of vibrational analysis with scanning probe microscopy allows probing vibrational properties of materials at the submicrometer scale, and even of individual molecules. This is accomplished by integrating scanning probe microscopy (SPM) and vibrational spectroscopy. This combination allows for much higher spatial resolution than can be achieved with conventional Raman/FTIR instrumentation. The technique is also nondestructive, requires non-extensive sample preparation, and provides more contrast such as intensity contrast, polarization contrast and wavelength contrast, as well as providing specific chemical information and topography images simultaneously.

<span class="mw-page-title-main">Two-state trajectory</span>

A two-state trajectory is a dynamical signal that fluctuates between two distinct values: ON and OFF, open and closed, , etc. Mathematically, the signal has, for every either the value or .

The following outline is provided as an overview of and topical guide to biophysics:

<span class="mw-page-title-main">Infrared Nanospectroscopy (AFM-IR)</span> Infrared microscopy technique

AFM-IR or infrared nanospectroscopy is one of a family of techniques that are derived from a combination of two parent instrumental techniques. AFM-IR combines the chemical analysis power of infrared spectroscopy and the high-spatial resolution of scanning probe microscopy (SPM). The term was first used to denote a method that combined a tuneable free electron laser with an atomic force microscope equipped with a sharp probe that measured the local absorption of infrared light by a sample with nanoscale spatial resolution.

<span class="mw-page-title-main">Cho Minhaeng</span> South Korean scientist (born 1965)

Cho Minhaeng is a South Korean scientist in researching physical chemistry, spectroscopy, and microscopy. He was director of the National Creative Research Initiative Center for Coherent Multidimensional Spectroscopy and is founding director of the Center for Molecular Spectroscopy and Dynamics in the Institute for Basic Science (IBS), located in Korea University.

<span class="mw-page-title-main">Fluorescence imaging</span> Type of non-invasive imaging technique

Fluorescence imaging is a type of non-invasive imaging technique that can help visualize biological processes taking place in a living organism. Images can be produced from a variety of methods including: microscopy, imaging probes, and spectroscopy.

References

  1. 1 2 Roiter V, Minko S (2005). "AFM Single Molecule Experiments at the Solid-Liquid Interface: In Situ Conformation of Adsorbed Flexible Polyelectrolyte Chains". Journal of the American Chemical Society. 127 (45): 15688–15689. doi:10.1021/ja0558239. PMID   16277495.
  2. Juette MF, Terry DS, Wasserman MR, Zhou Z, Altman RB, Zheng Q, Blanchard SC (June 2014). "The bright future of single-molecule fluorescence imaging". Curr Opin Chem Biol. 20: 103–111. doi:10.1016/j.cbpa.2014.05.010. PMC   4123530 . PMID   24956235.
  3. W. E. Moerner and L. Kador, Optical detection and spectroscopy of single molecules in a solid, Phys. Rev. Lett. 62, 2535 - 2538 (1989)
  4. M. Orrit and J. Bernard, Single pentacene molecules detected by fluorescence excitation in a p-terphenyl crystal, Phys. Rev. Lett. 65, 2716–2719 (1990)
  5. Kneipp, Katrin; Wang, Yang; Kneipp, Harald; Perelman, Lev T.; Itzkan, Irving; Dasari, Ramachandra R.; Feld, Michael S. (1997-03-03). "Single Molecule Detection Using Surface-Enhanced Raman Scattering (SERS)". Physical Review Letters. 78 (9): 1667–1670. doi:10.1103/PhysRevLett.78.1667. ISSN   0031-9007.
  6. Nie, Shuming; Emory, Steven R. (1997-02-21). "Probing Single Molecules and Single Nanoparticles by Surface-Enhanced Raman Scattering". Science. 275 (5303): 1102–1106. doi:10.1126/science.275.5303.1102. ISSN   0036-8075.
  7. 1 2 Murugesapillai D, et al. (2014). "DNA bridging and looping by HMO1 provides a mechanism for stabilizing nucleosome-free chromatin". Nucleic Acids Res. 42 (14): 8996–9004. doi: 10.1093/nar/gku635 . PMC   4132745 . PMID   25063301.
  8. 1 2 Murugesapillai D, et al. (2016). "Single-molecule studies of high-mobility group B architectural DNA bending proteins". Biophys Rev. 9 (1): 17–40. doi: 10.1007/s12551-016-0236-4 . PMC   5331113 . PMID   28303166.
  9. 1 2 3 B. Sakmann and E. Neher, Single-Channel Recording, ISBN   9780306414190 (1995).
  10. Picones, A; Keung, E; Timpe, LC (2001). "Unitary conductance variation in Kir2.1 and in cardiac inward rectifier potassium channels". Biophys J. 81 (4): 2035–2049. doi:10.1016/S0006-3495(01)75853-5.
  11. Podh, Nitesh Kumar; Paliwal, Sheetal; Dey, Partha; Das, Ayan; Morjaria, Shruti; Mehta, Gunjan (5 November 2021). "In-vivo Single-Molecule Imaging in Yeast: Applications and Challenges". Journal of Molecular Biology. 433 (22): 167250. doi:10.1016/j.jmb.2021.167250. PMID   34537238. S2CID   237573437.
  12. O. Flomenbom, J. Klafter, and A. Szabo, What can one learn from two-state single molecule trajectories? Archived January 14, 2012, at the Wayback Machine , Biophys. J. 88, 3780–3783 (2005); arXiv : q-bio/0502006
  13. O. Flomenbom, and R. J. Silbey, Utilizing the information content in two-state trajectories Archived January 14, 2012, at the Wayback Machine , Proc. Natl. Acad. Sci. USA 103, 10907–10910 (2006).
  14. O. Flomenbom, and R. J. Silbey, Toolbox for analyzing finite two-state trajectories, Phys. Rev. E 78, 066105 (2008); arXiv:0802.1520.
  15. O. Flomenbom, K. Velonia, D. Loos, et al., Stretched exponential decay and correlations in the catalytic activity of fluctuating single lipase molecules Archived January 14, 2012, at the Wayback Machine , Proc. Natl. Acad. Sci. US 102, 2368–2372 (2005).
  16. Zhan, Hong; Stanciauskas, Ramunas; Stigloher, Christian; Keomanee-Dizon, Kevin; Jospin, Maelle; Bessereau, Jean-Louis; Pinaud, Fabien (2014). "In vivo single-molecule imaging identifies altered dynamics of calcium channels in dystrophin-mutant C. elegans". Nature Communications. 5: ncomms5974. Bibcode:2014NatCo...5.4974Z. doi:10.1038/ncomms5974. PMC   4199201 . PMID   25232639.
  17. Walder, R.; Nelson, N.; Schwartz, D. K. (2011). "Super-resolution surface mapping using the trajectories of molecular probes". Nature Communications. 2: 515. Bibcode:2011NatCo...2..515W. doi: 10.1038/ncomms1530 . PMID   22044994.