User:Sainsf/sandbox: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Line 4: Line 4:
{{Short description|Molecular-scale artificial or biological device}}
{{Short description|Molecular-scale artificial or biological device}}
{{Molecular nanotechnology subfields}}
{{Molecular nanotechnology subfields}}
'''Molecular machines''' are a class of molecules typically described as an assembly of a discrete number of molecular components intended to produce mechanical movements in response to specific stimuli, mimicking macromolecular devices such as switches and motors. Naturally occurring or biological molecular machines such as [[kinesin]]s and [[ribosome]]s are responsible for vital living processes such as [[DNA replication]] and [[ATP synthase|ATP synthesis]]. For the last several decades, scientists have attempted, with varying degrees of success, to miniaturize machines found in the macroscopic world. The . In 2016 the [[Nobel Prize in Chemistry]] was awarded to [[Jean-Pierre Sauvage]], [[Fraser Stoddart|Sir J. Fraser Stoddart]], and [[Ben Feringa|Bernard L. Feringa]] for the design and synthesis of molecular machines.
'''Molecular machines''' are a class of molecules typically described as an assembly of a discrete number of molecular components intended to produce mechanical movements in response to specific stimuli, mimicking macromolecular devices such as switches and motors. Naturally occurring or biological molecular machines such as [[kinesin]]s and [[ribosome]]s (often in the form of [[Protein complex|multi-protein complex]]es) are responsible for vital living processes such as [[DNA replication]] and [[ATP synthase|ATP synthesis]]. For the last several decades, scientists have attempted, with varying degrees of success, to miniaturize machines found in the macroscopic world. The first example of an artificial molecular machine (AMM) was reported in 1994, featuring a [[rotaxane]] with a ring and two different possible binding sites. In 2016 the [[Nobel Prize in Chemistry]] was awarded to [[Jean-Pierre Sauvage]], [[Fraser Stoddart|Sir J. Fraser Stoddart]], and [[Ben Feringa|Bernard L. Feringa]] for the design and synthesis of molecular machines.

AMMs have diversified rapidly over the past few decades and their design principles, properties, and [[Characterization (materials science)|characterization]] methods have been outlined better. A major starting point for the design of AMMs is to exploit the existing modes of motion in molecules, such as rotation about single bonds or [[Cis–trans isomerism|''cis-trans'' isomerization]]. Different AMMs are produced by introducing various functionalities, such as the introduction of [[bistability]] to create switches. A broad range of AMMs has been designed, featuring different properties and applications; some of these include [[molecular motor]]s, [[molecular switch|switche]]s, and [[molecular logic gate|logic gate]]s. A wide range of applications have been demonstrated for AMMs, including those integrated into [[polymer]]ic, [[liquid crystal]], and [[crystal]]line systems for varied functions (such as [[Materials science|material]]s research, [[homogenous catalysis]] and [[surface chemistry]]).


==Terminology==
==Terminology==

Revision as of 06:52, 1 February 2023

Molecular machines are a class of molecules typically described as an assembly of a discrete number of molecular components intended to produce mechanical movements in response to specific stimuli, mimicking macromolecular devices such as switches and motors. Naturally occurring or biological molecular machines such as kinesins and ribosomes (often in the form of multi-protein complexes) are responsible for vital living processes such as DNA replication and ATP synthesis. For the last several decades, scientists have attempted, with varying degrees of success, to miniaturize machines found in the macroscopic world. The first example of an artificial molecular machine (AMM) was reported in 1994, featuring a rotaxane with a ring and two different possible binding sites. In 2016 the Nobel Prize in Chemistry was awarded to Jean-Pierre Sauvage, Sir J. Fraser Stoddart, and Bernard L. Feringa for the design and synthesis of molecular machines.

AMMs have diversified rapidly over the past few decades and their design principles, properties, and characterization methods have been outlined better. A major starting point for the design of AMMs is to exploit the existing modes of motion in molecules, such as rotation about single bonds or cis-trans isomerization. Different AMMs are produced by introducing various functionalities, such as the introduction of bistability to create switches. A broad range of AMMs has been designed, featuring different properties and applications; some of these include molecular motors, switches, and logic gates. A wide range of applications have been demonstrated for AMMs, including those integrated into polymeric, liquid crystal, and crystalline systems for varied functions (such as materials research, homogenous catalysis and surface chemistry).

Terminology

Several definitions describe a "molecular machine" as a class of molecules typically described as an assembly of a discrete number of molecular components intended to produce mechanical movements in response to specific stimuli. The expression is often more generally applied to molecules that simply mimic functions that occur at the macroscopic level.[1] A few prime requirements for a molecule to be considered a "molecular machine" are: the presence of moving parts, the ability to consume energy, and the ability to perform a task.[2] Molecular machines differ from other stimuli-responsive compounds that can produce motion (such as cis-trans isomers) in their relatively larger amplitude of movement (potentially due to chemical reactions) and the presence of a clear external stimulus to regulate the movements (as compared to random thermal motion).[1]

This definition generally applies to synthetic molecular machines, which have historically gained inspiration from the naturally occurring biological molecular machines (also referred to as "nanomachines"). Biological machines are considered to be nanoscale devices (such as molecular proteins) in a living system that convert various forms of energy to mechanical work in order to drive crucial biological processes such as intracellular transport, muscle contractions, ATP generation and cell division.[3][4]

History

Biological molecular machines have been known and studied for years given their vital role in sustaining life, and have served as inspiration for synthetically designed systems with similar useful functionality.[3][4] The advent of conformational analysis, or the study of conformers to analyze complex chemical structures, in the 1950s gave rise to the idea of understanding and controlling relative motion within molecular components for further applications. This led to the design of "proto-molecular machines" featuring conformational changes such as cog-wheeling of the aromatic rings in triptycenes.[6] By 1980, scientists could achieve desired conformations using external stimuli and utilize this for different applications. A major example is the design of a photoresponsive crown ether containing an azobenzene unit, which could switch between cis and trans isomers on exposure to light and hence tune the cation-binding properties of the ether.[7] In his seminal 1959 lecture There's Plenty of Room at the Bottom, Richard Feynman alluded to the idea and applications of molecular devices designed artificially by manipulating matter at the atomic level.[5] This was further substantiated by Eric Drexler during the 1970s, who developed ideas based on molecular nanotechnology such as nanoscale "assemblers",[8] though their feasibility was disputed.[9]

The first example of an artificial molecular machine (a switchable molecular shuttle). The positively charged ring (blue) is initially positioned over the benzidine unit (green), but shifts to the biphenol unit (red) when the benzidine gets protonated (purple)
The first example of an artificial molecular machine (a switchable molecular shuttle). The positively charged ring (blue) is initially positioned over the benzidine unit (green), but shifts to the biphenol unit (red) when the benzidine gets protonated (purple)[10]

Though these events served as inspiration for the field, the actual breakthrough in practical approaches to synthesize artificial molecular machines (AMMs) took place in 1991 with the invention of a "molecular shuttle" by Sir Fraser Stoddart.[10] Building upon the assembly of mechanically linked molecules such as catenanes and rotaxanes as developed by Jean-Pierre Sauvage in the early 1980s,[11][12] this shuttle features a rotaxane with a ring that can move across an "axle" between two ends or possible binding sites (hydroquinone units). This design realized the well-defined motion of a molecular unit across the length of the molecule for the first time.[6] In 1994, an improved design allowed control over the motion of the ring by pH variation or electrochemical methods, making it the first example of an AMM. Here the two binding sites are a benzidine and a biphenol unit; the cationic ring typically prefers staying over the benzidine ring, but moves over to the biphenol group when the benzidine gets protonated at low pH or if it gets electrochemically oxidized.[13] In 1998, a study could capture the rotary motion of a decacyclene molecule on a copper-base metallic surface using a scanning tunneling microscope.[14] Over the following decade, a broad variety of AMMs responding to various stimuli were invented for different applications.[15][16] In 2016, the Nobel Prize in Chemistry was awarded to Sauvage, Stoddart, and Bernard L. Feringa for the design and synthesis of molecular machines.[17][18]

Artificial molecular machines

Over the past few decades AMMs have diversified rapidly and their design principles,[2] properties,[19] and characterization methods[20] have been outlined more clearly. A major starting point for the design of AMMs is to exploit the existing modes of motion in molecules.[2] For instance, single bonds can be visualized as axes of rotation, as can be metallocene complexes. Bending or V-like shapes can be achieved by incorporating double bonds, that can undergo cis-trans isomerization in response to certain stimuli (typically irradiation with a suitable wavelength), as seen in numerous designs consisting of stilbene and azobenzene units. Similarly, ring-opening and -closing reactions such as those seen for spiropyran and diarylethene can also produce curved shapes. Another common mode of movement is the circumrotation of rings relative to one another as observed in mechanically interlocked molecules (primarily catenanes). While this type of rotation can not be accessed beyond the molecule itself (because the rings are confined within one another), rotaxanes can overcome this as the rings can undergo translational movements along a dumbbell-like axis. Another line of AMMs consists of biomolecules such as DNA and proteins as part of their design, making use of phenomena like protein folding and unfolding.[21][22]

Some common types of motion seen in some simple components of artificial molecular machines. a) Rotation around single bonds and in sandwich-like metallocenes. b) Bending due to cis-trans isomerization. c) Translational motion of a ring along the dumbbell-like rotaxane axis. d) Rotation of interlocked rings in a catenane
Some modes of motion present in some simple building blocks of artificial molecular machines. a) Rotation around single bonds and in sandwich-like metal complexes. b) Bending as a consequence of double-bond isomerization. c) Translational motion of a ring along the dumbbell of a rotaxane. d) Circumrotation motion of rings in a catenane.

AMM designs have diversified significantly since the early days of the field. A major route is the introduction of bistability to produce molecular switches, featuring two distinct configurations for the molecule to convert between. This has been perceived as a step forward from the original molecular shuttle which consisted of two identical sites for the ring to move between without any preference, in a manner analogous to the ring flip in an unsubstituted cyclohexane. If these two sites are different from each other in terms of features like electron density, this can give rise to weak or strong recognition sites as in biological systems -- such AMMs have found applications in catalysis and drug delivery. This switching behavior has been further optimized to acquire useful work that gets lost when a typical switch returns to its original state. Inspired from the use of kinetic control to produce work in natural processes, molecular motors are designed to have a continuous energy influx to keep them away from equilibrium to deliver work.

Various energy sources are employed to drive molecular machines today, but this was not the case during the early years of AMM development. Though the movements in AMMs were regulated relative to the random thermal motion generally seen in molecules, they could not be controlled or manipulated as desired. This led to the addition of stimuli-responsive moieties in AMM design, so that externally applied non-thermal sources of energy could drive molecular motion and hence allow control over the properties. Chemical energy (or "chemical fuels") was an attractive option at the beginning, given the broad array of reversible chemical reactions (heavily based on acid-base chemistry) to switch molecules between different states.[23] However, this comes with the issue of practically regulating the delivery of the chemical fuel and the removal of waste generated to maintain the efficiency of the machine as in biological systems. Though some AMMs have found ways to circumvent this, more recently waste-free reactions such based on electron transfers or isomerization have gained attention (such as redox-responsive viologens). Eventually, several different forms of energy (electric, magnetic, optical and so on) have become the primary energy sources used to power AMMs, even producing autonomous systems such as light-driven motors.

Types

Various AMMs have been designed with a broad range of functions and applications, several of which have been tabulated below along with indicative images:[19]

Type Details Image
Molecular balance A molecule that can interconvert between two and more conformational or configurational states in response to the dynamic of multiple intra- and intermolecular driving forces,[24][25] such as hydrogen bonding, solvophobic or hydrophobic effects,[26] π interactions,[27] and steric and dispersion interactions.[28] The distinct conformers of a molecular balance can show different interactions with the same molecule, such that analyzing the ratio of the conformers and the energies for these interactions can enable quantification of different properties (such as CH-π or arene-arene interactions, see image).[29][30] An example of a molecular balance
Molecular hinge A molecular hinge is a molecule that can typically rotate in a crank-like motion around a rigid axis, such as a double bond or aromatic ring, to switch between reversible configurations.[31] Such configurations must have distinguishable geometries; for instance, azobenzene groups in a linear molecule may undergo cis-trans isomerization[32] when irradiated with ultraviolet light, triggering a reversible transition to a bent or V-shaped conformation (see image).[33][34][35] Molecular hinges have been adapted for applications such as nucleotide base recognition,[36] peptide modifications,[37] and visualizing molecular motion.[38] An example of a molecular hinge that can undergo cis-trans isomerization about a double bond
Molecular logic gate A molecule that performs a logical operation on one or more logic inputs and produces a single logic output.[39] Modelled on logic gates, these molecules have slowly replaced the conventional silicon-based machinery. Several applications have come forth, such as water quality examination, food safety examination, metal ion detection, and pharmaceutical studies.[40][41] The first example of a molecular logic gate was reported in 1993, featuring a receptor (see image) where the emission intensity could be treated as a tunable output if the concentrations of protons and sodium ions were to be considered as inputs.[42] The first reported molecular logic gate
Molecular motor A molecule that is capable of directional rotary motion around a single or double bond.[43][44] Carbon nanotube nanomotors have also been produced.[45] Molecular dynamics simulation of a synthetic molecular rotor composed of three molecules in a nanopore (outer diameter 6.7 nm) at 250 K
Molecular necklace A class of mechanically interlocked molecules derived from catenanes where a large macrocycle backbone connects at least three small rings in the shape of a necklace. A molecular necklace consisting of a large macrocycle threaded by n-1 rings (hence comprising n rings) is represented as [n]MN.[46] The first molecular necklace was synthesized in 1992, featuring several α-cyclodextrins on a single polyethylene glycol chain backbone; the authors connected this to the idea of a "molecular abacus" proposed by Stoddart and coworkers around the same time.[47] Several interesting applications have emerged for these molecules, such as antibacterial activity,[48] desulfurization of fuels,[49] and piezoelectricity.[50] An example of a molecular necklace
Molecular propeller An example of a molecular propeller pumping water molecules due to its hydrophobic surface
Molecular shuttle An example of a rotaxane-based molecular shuttle
Molecular switch A molecule that can be reversibly shifted between two or more stable states in response to certain stimuli. This change of states influences the properties of the molecule according to the state it occupies at the moment. Unlike a molecular motor, any mechanical work done due to the motion in a switch is generally undone once the the molecule returns to its original state unless it is part of a larger motor-like system.[51] An example of a molecular switch
Molecular tweezers An example of molecular tweezers binding a fullerene
Nanocar A nanocar with C60 fullerenes as wheels

Biological molecular machines

A ribosome performing the elongation and membrane targeting stages of protein translation. The ribosome is green and yellow, the tRNAs are dark blue, and the other proteins involved are light blue. The produced peptide is released into the endoplasmic reticulum.

The most complex macromolecular machines are found within cells, often in the form of multi-protein complexes.[52] Important examples of biological machines include motor proteins such as myosin, which is responsible for muscle contraction, kinesin, which moves cargo inside cells away from the nucleus along microtubules, and dynein, which moves cargo inside cells towards the nucleus and produces the axonemal beating of motile cilia and flagella. "[I]n effect, the [motile cilium] is a nanomachine composed of perhaps over 600 proteins in molecular complexes, many of which also function independently as nanomachines ... Flexible linkers allow the mobile protein domains connected by them to recruit their binding partners and induce long-range allostery via protein domain dynamics."[53] Other biological machines are responsible for energy production, for example ATP synthase which harnesses energy from proton gradients across membranes to drive a turbine-like motion used to synthesise ATP, the energy currency of a cell.[54] Still other machines are responsible for gene expression, including DNA polymerases for replicating DNA, RNA polymerases for producing mRNA, the spliceosome for removing introns, and the ribosome for synthesising proteins. These machines and their nanoscale dynamics are far more complex than any molecular machines that have yet been artificially constructed.[55]

Biological machines have potential applications in nanomedicine. [56] For example, they could be used to identify and destroy cancer cells.[57][58] Molecular nanotechnology is a speculative subfield of nanotechnology regarding the possibility of engineering molecular assemblers, biological machines which could re-order matter at a molecular or atomic scale. Nanomedicine would make use of these nanorobots, introduced into the body, to repair or detect damages and infections, but these are considered to be far beyond current capabilities.[59]

Research and applications

The construction of more complex molecular machines is an active area of theoretical and experimental research. Though a diverse variety of AMMs are known today, experimental studies of these molecules are inhibited by the lack of methods to construct these molecules.[60] In this context, theoretical modeling has emerged as a pivotal tool to understand the self-assembly or disassembly processes in these systems.[61][62]

A wide range of applications have been demonstrated for AMMs, including those integrated into polymeric,[63][64] liquid crystal,[65][66] and crystalline[67][68] systems for varied functions.. Homogenous catalysis is a prominent example, especially in areas like asymmetric synthesis, utilizing noncovalent interactions and biomimetic allosteric catalysis.[69][70] AMMs have been pivotal in the design of several stimuli-responsive smart materials, such as 2D and 3D self-assembled materials and nanoparticle-based systems, for versatile applications ranging from 3D printing to drug delivery.[71][72] AMMs are gradually moving from the conventional solution-phase chemistry to surfaces and interfaces. For instance, AMM-immobilized surfaces (AMMISs) are a novel class of functional materials consisting of AMMs attached to inorganic surfaces forming features like self-assembled monolayers; this gives rise to tunable properties such as fluorescence, aggregation and drug-release activity.[73] Most of these applications remain at the proof-of-concept level, and need major modifications to be adapted to the industrial scale. Challenges in streamlining macroscale applications include autonomous operation, the complexity of the machines, stability in the synthesis of the machines and the working conditions.[1][74]

References

  1. ^ a b c Vincenzo, V.; Credi, A.; Raymo, F. M.; Stoddart, J. F. (2000). "Artificial Molecular Machines". Angewandte Chemie International Edition. 39 (19): 3348–3391. doi:10.1002/1521-3773(20001002)39:19<3348::AID-ANIE3348>3.0.CO;2-X.
  2. ^ a b c Cheng, C.; Stoddart, J. F. (2016). "Wholly Synthetic Molecular Machines". ChemPhysChem. 17 (12): 1780–1793. doi:10.1002/cphc.201501155.
  3. ^ a b Huang, T. J.; Juluri, B. K. (2008). "Biological and biomimetic molecular machines". Nanomedicine. 3 (1): 107–124. doi:10.2217/17435889.3.1.107.
  4. ^ a b Kinbara, K.; Aida, T. (2005). "Toward Intelligent Molecular Machines: Directed Motions of Biological and Artificial Molecules and Assemblies". Chemical Reviews. 105 (4): 1377–1400. doi:10.1021/cr030071r.
  5. ^ a b Feynman, R. (1960). "There's Plenty of Room at the Bottom" (PDF). Engineering and Science. 23 (5): 22–36.
  6. ^ a b Kay, E. R.; Leigh, D. A. (2015). "Rise of the molecular machines". Angewandte Chemie International Edition. 54 (35): 10080–10088. doi:10.1002/anie.201503375.
  7. ^ Shinkai, S.; Nakaji, T.; Nishida, Y.; Ogawa, T.; Manabe, O. (1980). "Photoresponsive crown ethers. 1. Cis-trans isomerism of azobenzene as a tool to enforce conformational changes of crown ethers and polymers". Journal of the American Chemical Society. 102 (18): 5860–5865. doi:10.1021/ja00538a026.
  8. ^ Drexler, K. E. (1981). "Molecular engineering: An approach to the development of general capabilities for molecular manipulation". Proceedings of the National Academy of Sciences. 78 (9): 5275–5278. doi:10.1073/pnas.78.9.5275.
  9. ^ Baum, R. (1 December 2003). "Drexler and Smalley make the case for and against 'molecular assemblers'". C&EN. Vol. 81, no. 48. pp. 37–42. Retrieved 16 January 2023.
  10. ^ a b Anelli, P. L.; Spencer, N.; Stoddart, J. F. (1991). "A molecular shuttle". Journal of the American Chemical Society. 113 (13): 5131–5133. doi:10.1021/ja00013a096.
  11. ^ Dietrich-Buchecker, C. O.; Sauvage, J. P.; Kintzinger, J. P. (1983). "Une nouvelle famille de molecules : les metallo-catenanes" [A new family of molecules: metallo-catenanes]. Tetrahedron Letters (in French). 24 (46): 5095–5098. doi:10.1016/S0040-4039(00)94050-4.
  12. ^ Dietrich-Buchecker, C. O.; Sauvage, J. P.; Kern, J. M. (May 1984). "Templated synthesis of interlocked macrocyclic ligands: the catenands". Journal of the American Chemical Society. 106 (10): 3043–3045. doi:10.1021/ja00322a055.
  13. ^ Bissell, R. A; Córdova, E.; Kaifer, A. E.; Stoddart, J. F. (1994). "A chemically and electrochemically switchable molecular shuttle". Nature. 369 (6476): 133–137. doi:10.1038/369133a0.
  14. ^ Gimzewski, J. K.; Joachim, C.; Schlittler, R. R.; Langlais, V.; Tang, H.; Johannsen, I. (1998). "Rotation of a Single Molecule Within a Supramolecular Bearing". Science. 281 (5376): 531–533. doi:10.1126/science.281.5376.531.
  15. ^ Balzani, V.; Credi, A.; Raymo, F. M.; Stoddart, J. F. (2000). "Artificial Molecular Machines". Angewandte Chemie International Edition. 39 (19): 3348–3391. doi:10.1002/1521-3773(20001002)39:19<3348::AID-ANIE3348>3.0.CO;2-X.
  16. ^ Erbas-Cakmak, S.; Leigh, D. A.; McTernan, C. T.; Nussbaumer, A. L. (2015). "Artificial Molecular Machines". Chemical Reviews. 115 (18): 10081–10206. doi:10.1021/acs.chemrev.5b00146.
  17. ^ Staff (5 October 2016). "The Nobel Prize in Chemistry 2016". Nobel Foundation. Retrieved 5 October 2016.
  18. ^ Chang, Kenneth; Chan, Sewell (5 October 2016). "3 Makers of 'World's Smallest Machines' Awarded Nobel Prize in Chemistry". New York Times. Retrieved 5 October 2016.
  19. ^ a b Erbas-Cakmak, Sundus; Leigh, David A.; McTernan, Charlie T.; Nussbaumer, Alina L. (2015). "Artificial Molecular Machines". Chemical Reviews. 115 (18): 10081–10206. doi:10.1021/acs.chemrev.5b00146. PMC 4585175. PMID 26346838.
  20. ^ Nogales, E.; Grigorieff, N. (2001). "Molecular Machines: putting the pieces together". The Journal of cell biology. 152 (1): F1-10. doi:10.1083/jcb.152.1.f1. PMID 11149934.
  21. ^ Mao, X.; Liu, M.; Li, Q.; Fan, C.; Zuo, X. (2022). "DNA-Based Molecular Machines". JACS Au. 2 (11): 2381–2399. doi:10.1021/jacsau.2c00292.
  22. ^ Saper, G.; Hess, H. (2020). "Synthetic Systems Powered by Biological Molecular Motors". Chemical Reviews. 120 (1): 288–309. doi:10.1021/acs.chemrev.9b00249.
  23. ^ Biagini, C.; Di Stefano, S. (2020). "Abiotic Chemical Fuels for the Operation of Molecular Machines". Angewandte Chemie International Edition. 59 (22): 8344–8354. doi:10.1002/anie.201912659.
  24. ^ Paliwal, S.; Geib, S.; Wilcox, C. S. (1994). "Molecular Torsion Balance for Weak Molecular Recognition Forces. Effects of "Tilted-T" Edge-to-Face Aromatic Interactions on Conformational Selection and Solid-State Structure". Journal of the American Chemical Society. 116 (10): 4497–4498. doi:10.1021/ja00089a057.
  25. ^ Mati, Ioulia K.; Cockroft, Scott L. (2010). "Molecular balances for quantifying non-covalent interactions" (PDF). Chemical Society Reviews. 39 (11): 4195–4205. doi:10.1039/B822665M. hdl:20.500.11820/7ce18ff7-1196-48a1-8c67-3bc3f6b46946. PMID 20844782. S2CID 263667.
  26. ^ Y., Lixu; A., Catherine; Cockroft, S. L. (2015). "Quantifying Solvophobic Effects in Nonpolar Cohesive Interactions". Journal of the American Chemical Society. 137 (32): 10084–10087. doi:10.1021/jacs.5b05736. hdl:20.500.11820/604343eb-04aa-4d90-82d2-0998898400d2. ISSN 0002-7863. PMID 26159869.
  27. ^ L., Ping; Z., Chen; Smith, M. D.; Shimizu, K. D. (2013). "Comprehensive Experimental Study of N-Heterocyclic π-Stacking Interactions of Neutral and Cationic Pyridines". The Journal of Organic Chemistry. 78 (11): 5303–5313. doi:10.1021/jo400370e. PMID 23675885.
  28. ^ Hwang, J.; Li, P.; Smith, M. D.; Shimizu, K. D. (2016). "Distance-Dependent Attractive and Repulsive Interactions of Bulky Alkyl Groups". Angewandte Chemie International Edition. 55 (28): 8086–8089. doi:10.1002/anie.201602752. PMID 27159670.
  29. ^ Carroll, W. R.; Zhao, C.; Smith, M. D.; Pellechia, P. J.; Shimizu, K. D. (2011). "A Molecular Balance for Measuring Aliphatic CH−π Interactions". Organic Letters. 13 (16): 4320–4323. doi:10.1021/ol201657p.
  30. ^ Carroll, W. R.; Pellechia, P.; Shimizu, K. D. (2008). "A Rigid Molecular Balance for Measuring Face-to-Face Arene−Arene Interactions". Organic Letters. 10 (16): 3547–3550. doi:10.1021/ol801286k.
  31. ^ Kassem, Salma; van Leeuwen, Thomas; Lubbe, Anouk S.; Wilson, Miriam R.; Feringa, Ben L.; Leigh, David A. (2017). "Artificial molecular motors" (PDF). Chemical Society Reviews. 46 (9): 2592–2621. doi:10.1039/C7CS00245A. PMID 28426052.
  32. ^ Bandara, H. M. Dhammika; Burdette, S. C. (2012). "Photoisomerization in different classes of azobenzene". Chemical Society Reviews. 41 (5): 1809–1825. doi:10.1039/c1cs15179g. PMID 22008710.
  33. ^ Wang, J.; Jiang, Q.; Hao, X.; Yan, H.; Peng, H.; Xiong, B.; Liao, Y.; Xie, X. (2020). "Reversible photo-responsive gel–sol transitions of robust organogels based on an azobenzene-containing main-chain liquid crystalline polymer". RSC Advances. 10 (7): 3726–3733. Bibcode:2020RSCAd..10.3726W. doi:10.1039/C9RA10161F.
  34. ^ Hada, M.; Yamaguchi, D.; Ishikawa, T.; Sawa, T.; Tsuruta, K.; Ishikawa, K.; Koshihara, S.-y.; Hayashi, Y.; Kato, T. (13 September 2019). "Ultrafast isomerization-induced cooperative motions to higher molecular orientation in smectic liquid-crystalline azobenzene molecules". Nature Communications. 10 (1): 4159. Bibcode:2019NatCo..10.4159H. doi:10.1038/s41467-019-12116-6. ISSN 2041-1723. PMC 6744564. PMID 31519876.
  35. ^ Garcia-Amorós, J.; Reig, M.; Cuadrado, A.; Ortega, M.; Nonell, S.; Velasco, D. (2014). "A photoswitchable bis-azo derivative with a high temporal resolution". Chemical Communications. 50 (78): 11462–11464. doi:10.1039/C4CC05331A. PMID 25132052.
  36. ^ Hamilton, A. D.; Van Engen, D. (1987). "Induced fit in synthetic receptors: nucleotide base recognition by a molecular hinge". Journal of the American Chemical Society. 109 (16): 5035–5036. doi:10.1021/ja00250a052.
  37. ^ Dumy, P.; Keller, M.; Ryan, D. E.; Rohwedder, B.; Wöhr, T.; Mutter, M. (1997). "Pseudo-Prolines as a Molecular Hinge: Reversible Induction of cis Amide Bonds into Peptide Backbones". Journal of the American Chemical Society. 119 (5): 918–925. doi:10.1021/ja962780a.
  38. ^ Ai, Y.; Chan, M. H.-Y.; Chan, A. K.-W.; Ng, M.; Li, Y.; Yam, V. W.-W. (2019). "A platinum(II) molecular hinge with motions visualized by phosphorescence changes". Proceedings of the National Academy of Sciences. 116 (28): 13856–13861. doi:10.1073/pnas.1908034116.
  39. ^ Erbas-Cakmak, S.; Kolemen, S.; Sedgwick, A. C.; Gunnlaugsson, T.; James, T. D.; Yoon, J.; Akkaya, E. U. (2018). "Molecular logic gates: the past, present and future". Chemical Society Reviews. 47 (7): 2228–2248. doi:10.1039/C7CS00491E.
  40. ^ de Silva, A. P. (2011). "Molecular Logic Gate Arrays". Chemistry - An Asian Journal. 6 (3): 750–766. doi:10.1002/asia.201000603.
  41. ^ Liu, L.; Liu, P.; Ga, L.; Ai, J. (2021). "Advances in Applications of Molecular Logic Gates". ACS Omega. 6 (45): 30189–30204. doi:10.1021/acsomega.1c02912.
  42. ^ de Silva, P. A.; Gunaratne, N. H. Q.; McCoy, C. P. (1993). "A molecular photoionic AND gate based on fluorescent signalling". Nature. 364 (6432): 42–44. doi:10.1038/364042a0.
  43. ^ Lancia, F.; Ryabchun, A.; Katsonis, N. (2019). "Life-like motion driven by artificial molecular machines". Nature Reviews Chemistry. 3 (9): 536–551. doi:10.1038/s41570-019-0122-2.
  44. ^ Mickler, M.; Schleiff, E.; Hugel, T. (2008). "From Biological towards Artificial Molecular Motors". ChemPhysChem. 9 (11): 1503–1509. doi:10.1002/cphc.200800216.
  45. ^ Fennimore, A. M.; Yuzvinsky, T. D.; Han, Wei-Qiang; Fuhrer, M. S.; Cumings, J.; Zettl, A. (24 July 2003). "Rotational actuators based on carbon nanotubes". Nature. 424 (6947): 408–410. Bibcode:2003Natur.424..408F. doi:10.1038/nature01823. PMID 12879064. S2CID 2200106.
  46. ^ Zhang, Z.; Zhao, J.; Guo, Z.; Zhang, H.; Pan, H.; Wu, Q.; You, W.; Yu, W.; Yan, X. (2022). "Mechanically interlocked networks cross-linked by a molecular necklace". Nature Communications. 13 (1): 1393. doi:10.1038/s41467-022-29141-7.
  47. ^ Harada, A.; Li, J.; Kamachi, M. (1992). "The molecular necklace: a rotaxane containing many threaded α-cyclodextrins". Nature. 356 (6367): 325–327. doi:10.1038/356325a0.
  48. ^ Wu, G.-Y.; Shi, X.; Phan, H.; Qu, H.; Hu, Y.-X.; Yin, G.-Q.; Zhao, X.-L.; Li, X.; Xu, L.; Yu, Q.; Yang, H.-B. (2020). "Efficient self-assembly of heterometallic triangular necklace with strong antibacterial activity". Nature Communications. 11 (1): 3178. doi:10.1038/s41467-020-16940-z.
  49. ^ Li, S.-L.; Lan, Y.-Q.; Sakurai, H.; Xu, Q. (2012). "Unusual Regenerable Porous Metal-Organic Framework Based on a New Triple Helical Molecular Necklace for Separating Organosulfur Compounds". Chemistry - A European Journal. 18 (51): 16302–16309. doi:10.1002/chem.201203093.
  50. ^ Seo, J.; Kim, B.; Kim, M.-S.; Seo, J.-H. (2021). "Optimization of Anisotropic Crystalline Structure of Molecular Necklace-like Polyrotaxane for Tough Piezoelectric Elastomer". ACS Macro Letters. 10 (11): 1371–1376. doi:10.1021/acsmacrolett.1c00567.
  51. ^ Kassem, S.; van Leeuwen, T.; Lubbe, A. S.; Wilson, M. R.; Feringa, B. L.; Leigh, D. A. (2017). "Artificial molecular motors". Chemical Society Reviews. 46 (9): 2592–2621. doi:10.1039/C7CS00245A.
  52. ^ Donald, Voet (2011). Biochemistry. Voet, Judith G. (4th ed.). Hoboken, NJ: John Wiley & Sons. ISBN 9780470570951. OCLC 690489261.
  53. ^ Satir, P.; Christensen, S. T. (2008). "Structure and function of mammalian cilia". Histochemistry and Cell Biology. 129 (6): 687–693. doi:10.1007/s00418-008-0416-9.
  54. ^ Kinbara, Kazushi; Aida, Takuzo (2005-04-01). "Toward Intelligent Molecular Machines: Directed Motions of Biological and Artificial Molecules and Assemblies". Chemical Reviews. 105 (4): 1377–1400. doi:10.1021/cr030071r. ISSN 0009-2665. PMID 15826015.
  55. ^ Bu Z, Callaway DJ (2011). "Proteins MOVE! Protein dynamics and long-range allostery in cell signaling". Protein Structure and Diseases. Advances in Protein Chemistry and Structural Biology. Vol. 83. Academic Press. pp. 163–221. doi:10.1016/B978-0-12-381262-9.00005-7. ISBN 9780123812629. PMID 21570668.
  56. ^ Amrute-Nayak, M.; Diensthuber, R. P.; Steffen, W.; Kathmann, D.; Hartmann, F. K.; Fedorov, R.; Urbanke, C.; Manstein, D. J.; Brenner, B.; Tsiavaliaris, G. (2010). "Targeted Optimization of a Protein Nanomachine for Operation in Biohybrid Devices". Angewandte Chemie. 122 (2): 322–326. Bibcode:2010AngCh.122..322A. doi:10.1002/ange.200905200. PMID 19921669.
  57. ^ Patel, G. M.; Patel, G. C.; Patel, R. B.; Patel, J. K.; Patel, M. (2006). "Nanorobot: A versatile tool in nanomedicine". Journal of Drug Targeting. 14 (2): 63–7. doi:10.1080/10611860600612862. PMID 16608733. S2CID 25551052.
  58. ^ Balasubramanian, S.; Kagan, D.; Jack Hu, C. M.; Campuzano, S.; Lobo-Castañon, M. J.; Lim, N.; Kang, D. Y.; Zimmerman, M.; Zhang, L.; Wang, J. (2011). "Micromachine-Enabled Capture and Isolation of Cancer Cells in Complex Media". Angewandte Chemie International Edition. 50 (18): 4161–4164. doi:10.1002/anie.201100115. PMC 3119711. PMID 21472835.
  59. ^ Freitas, Robert A. Jr.; Havukkala, Ilkka (2005). "Current Status of Nanomedicine and Medical Nanorobotics" (PDF). Journal of Computational and Theoretical Nanoscience. 2 (4): 471. Bibcode:2005JCTN....2..471K. doi:10.1166/jctn.2005.001.
  60. ^ Golestanian, Ramin; Liverpool, Tanniemola B.; Ajdari, Armand (2005-06-10). "Propulsion of a Molecular Machine by Asymmetric Distribution of Reaction Products". Physical Review Letters. 94 (22): 220801. arXiv:cond-mat/0701169. Bibcode:2005PhRvL..94v0801G. doi:10.1103/PhysRevLett.94.220801. PMID 16090376. S2CID 18989399.
  61. ^ Drexler, K. Eric (1999-01-01). "Building molecular machine systems". Trends in Biotechnology. 17 (1): 5–7. doi:10.1016/S0167-7799(98)01278-5. ISSN 0167-7799.
  62. ^ Tabacchi, G.; Silvi, S.; Venturi, M.; Credi, A.; Fois, E. (2016). "Dethreading of a Photoactive Azobenzene-Containing Molecular Axle from a Crown Ether Ring: A Computational Investigation". ChemPhysChem. 17 (12): 1913–1919. doi:10.1002/cphc.201501160. hdl:11383/2057447. PMID 26918775. S2CID 9660916.
  63. ^ Ikejiri, S.; Takashima, Y.; Osaki, M.; Yamaguchi, H.; Harada, A. (2018). "Solvent-Free Photoresponsive Artificial Muscles Rapidly Driven by Molecular Machines". Journal of the American Chemical Society. 140 (49): 17308–17315. doi:10.1021/jacs.8b11351.
  64. ^ Iwaso, K.; Takashima, Y.; Harada, A. (2016). "Fast response dry-type artificial molecular muscles with [c2]daisy chains". Nature Chemistry. 8 (6): 625–632. doi:10.1038/nchem.2513.
  65. ^ Orlova, T.; Lancia, F.; Loussert, C.; Iamsaard, S.; Katsonis, N.; Brasselet, E. (2018). "Revolving supramolecular chiral structures powered by light in nanomotor-doped liquid crystals". Nature Nanotechnology. 13 (4): 304–308. doi:10.1038/s41565-017-0059-x.
  66. ^ Hou, J.; Long, G.; Zhao, W.; Zhou, G.; Liu, D.; Broer, D. J.; Feringa, B. L.; Chen, J. (2022). "Phototriggered Complex Motion by Programmable Construction of Light-Driven Molecular Motors in Liquid Crystal Networks". Journal of the American Chemical Society. 144 (15): 6851–6860. doi:10.1021/jacs.2c01060.
  67. ^ Terao, F.; Morimoto, M.; Irie, M. (2012). "Light-Driven Molecular-Crystal Actuators: Rapid and Reversible Bending of Rodlike Mixed Crystals of Diarylethene Derivatives". Angewandte Chemie International Edition. 51 (4): 901–904. doi:10.1002/anie.201105585.
  68. ^ Vogelsberg, C. S.; Garcia-Garibay, M. A. (2012). "Crystalline molecular machines: function, phase order, dimensionality, and composition". Chemical Society Reviews. 41 (5): 1892–1910. doi:10.1039/c1cs15197e.
  69. ^ van Dijk, L.; Tilby, M. J.; Szpera, R.; Smith, O. A.; Bunce, H. A. P.; Fletcher, S. P. (2018). "Molecular machines for catalysis". Nature Reviews Chemistry. 2 (3): 0117. doi:10.1038/s41570-018-0117.
  70. ^ Neal, E. A.; Goldup, S. M. (2014). "Chemical consequences of mechanical bonding in catenanes and rotaxanes: isomerism, modification, catalysis and molecular machines for synthesis". Chemical Communications. 50 (40): 5128–5142. doi:10.1039/C3CC47842D.
  71. ^ Corra, S.; Curcio, M.; Baroncini, M.; Silvi, S.; Credi, A. (2020). "Photoactivated Artificial Molecular Machines that Can Perform Tasks". Advanced Materials. 32 (20): 1906064. doi:10.1002/adma.201906064.
  72. ^ Moulin, E.; Faour, L.; Carmona‐Vargas, C. C.; Giuseppone, N. (2020). "From Molecular Machines to Stimuli‐Responsive Materials". Advanced Materials. 32 (20): 1906036. doi:10.1002/adma.201906036.
  73. ^ Zhang, Q.; Qu, D.-H. (2016). "Artificial Molecular Machine Immobilized Surfaces: A New Platform To Construct Functional Materials". ChemPhysChem. 17 (12): 1759–1768. doi:10.1002/cphc.201501048.
  74. ^ Aprahamian, I. (2020). "The Future of Molecular Machines". ACS Central Science. 6 (3): 347–358. doi:10.1021/acscentsci.0c00064.