Peptides work as both functional molecules to modulate various biological phenomena and self-assembling artificial materials. The introduction of photoresponsive units to peptides allows the spatiotemporal remote control of their structure and function upon light irradiation. This article overviews the photoresponsive peptide design, interaction with biomolecules, and applications in self-assembling materials over the last 30 years. Peptides modified with photochromic (photoisomerizable) molecules, such as azobenzene and spiropyran, reversibly photo-controlled the binding to biomolecules and nanostructure formation through self-assembly. Photocleavable molecular units irreversibly control the functions of peptides through cleavage of the main chain and deprotection by light. Photocrosslinking between peptides or between peptides and other biomolecules enhances the structural stability of peptide assemblies and complexes. These photoresponsive peptides spatiotemporally controlled the formation and dissociation of peptide assemblies, gene expressions, protein–drug interactions, protein–protein interactions, liposome deformation and motility, cytoskeleton structure and stability, and cell functions by appropriate light irradiation. These molecular systems can be applied to photo-control biological functions, molecular robots, artificial cells, and next-generation smart drug delivery materials.

Peptides, which are short protein fragments, not only act as functional molecules that modulate many biological phenomena, such as protein–protein interaction,1–3 gene expression,4–6 cytoskeleton dynamics,7 and amyloid fibril formation,8 but also attract attention as components for biomaterials,9–11 therapeutic drugs,9–14 bioimaging probes,15–17 and nanomaterials.18–20 Biomaterials composed of peptides can be mainly used for tissue engineering scaffolds, drug delivery vehicles, and three-dimensional printing bioinks.9–11 Cell-penetrating peptides and organelle-localized peptides play a crucial role in efficient drug delivery systems into cells and bioimaging in cells.21,22 Peptides that bind to metals and semiconductor materials serve as an interface for adhesion between inorganic and biological materials.23 Rationally designed peptides that self-assemble through the formation of secondary structures, such as α-helices and β-sheets, can be used to models of amyloid fibril formation, gel materials for tissue engineering, self-healing materials, nanocapsules, and nanotubes for drug delivery.23–30 These functional peptide materials can be synthesized using solid-phase peptide synthesis and ligation technology between peptides, based on the rational design.31,32 Selecting peptides with desired functions from biosynthesized random libraries is possible using phage display and other peptide display technologies.33 

Imparting stimulus-responsiveness, such as pH, temperature, light, metal ions, and enzymes, to peptides and proteins allows them to be equipped with smart features to control biological activity, drug release, and nanomaterial properties.11 In particular, the introduction of photoresponsive units to peptides and proteins allows the spatiotemporal remote control of their structure and function upon light irradiation.34–38 Light is a clean energy source with low biological toxicity that can be endowed with a range of capabilities due to its lack of waste formation, wavelength tunability, recyclability, and potential for spatiotemporal resolution. Photostimulation with such advantageous properties enables even photocontrol of the structure and function of peptides and proteins in living systems.39,40 Photochromic (photoisomerizable), photocleavable, photocrosslinking, and photodimerizable peptide materials are the most prominent design strategies for photoresponsive peptide materials (Fig. 1). This review presents an overview of each photoresponsive peptide material, especially photocontrol of peptides, regarding self-assembly, biomolecule interaction, such as proteins and nucleic acids, and cytoskeleton functions. Figure 2 summarizes the available photoresponsive molecular tools.41–46 Azobenzene, which is most prominently used as a photoswitch, isomerizes from the trans state to the cis state by ultraviolet (UV) light irradiation, and returns to the trans state by visible light irradiation or heat [Fig. 2(a)].43,44 Similarly, spiropyran isomerizes to merocyanine by UV light irradiation and reverts to spiropyran by visible light irradiation or heat.45,46 The isomerization between uncharged spiropyran and zwitterionic merocyanine induces large structural changes in the peptides. Photochromic diarylethenes and fulgides are advantageous by being able to isomerize only with light of different wavelengths and not thermally.41,42,45 Photocleavable molecular units, which are typified by o-nitrobenzyl groups, irreversibly control the functions of peptides by main chain cleavage and light deprotection.41,47 Photocrosslinking of peptides or peptides and other biomolecules increases the structural stability of peptide assemblies and complexes and allows the profiling of peptide-biomolecular interactions.48 

FIG. 1.

Overview of photoresponsive peptide materials.

FIG. 1.

Overview of photoresponsive peptide materials.

Close modal
FIG. 2.

Examples of molecular units for photoresponsive peptide materials. (a) Photochromic, (b) photocleavable, (c) photocrosslinking, and (d) photodimerizable molecular units.

FIG. 2.

Examples of molecular units for photoresponsive peptide materials. (a) Photochromic, (b) photocleavable, (c) photocrosslinking, and (d) photodimerizable molecular units.

Close modal

Fibrous peptide assemblies are useful not only for elucidating the formation mechanism of amyloid fibrils, which are involved in protein misfolding diseases, such as Alzheimer's and prion diseases, but also as self-assembling nanomaterials that can be used for self-healing materials, artificial extracellular matrices, drug delivery, and nanoelectronics. By introducing photochromic units, peptide nanofibers acquire the ability to switch their structure and function with light. Aemissegger and co-workers developed a β-hairpin peptide that incorporated a photochromic azobenzene linker as a turn element [Fig. 3(a)].49,50 The DPro-Gly segment of 12-residue peptide, derived from protein GB1, was replaced by an azobenzene linker, 3-[(3-aminomethyl)-phenylazo]phenylacetic acid (AMPP). The thermodynamically favored trans-form of azobenzene-containing peptide self-assembled into oligomer, whereas the photoisomerization to cis-form induced well-defined β-hairpin monomer formation. The turn mimic that is based on azobenzene photoswitch can serve as a photoregulatory element to control β-hairpin formation in peptide hormones or larger proteins. Waldauer et al. determined hexapeptide (Ac-CHGGCK-NH2) crosslinked with azobenzene at the two Cys residues by molecular dynamics simulations.51 The peptide that crosslinked with azobenzene enables reversible photo-triggered aggregation to form amyloid-like fibril in the trans state, whereas soluble monomer in the cis state. Doran et al. incorporated an AMPP into the putative turn region of the amyloid β 42 (Aβ42) chain as a photoswitch linker [Fig. 3(b)].52 The trans state of the azobenzene-containing Aβ42 has formed fibrous assemblies with cytotoxicity that is similar to wild-type Aβ42. In contrast, photoisomerization to the cis state afforded amorphous aggregate that exhibits no cytotoxicity. β-Turn nucleation has been proposed as a rate-determining step in the self-assembly pathway of Aβ fibril, but these results demonstrated insight into Aβ self-assembly that β-turn intermediates are not strictly required for Aβ fibril or cytotoxic oligomer formation.

FIG. 3.

Photoisomerization of azobenzene controls peptide assemblies. (a) β-hairpin structure controlled by azobenzene-containing peptide photoisomerization. Reproduced with permission from Aemissegger et al., J. Am. Chem. Soc. 127, 2929–2936 (2005). Copyright 2005 American Chemical Society. (b) Cross-β amyloid fibril formation by photoisomerization of azobenzene-containing Amyloid-β mimic peptide. Reproduced with permission from Doran et al., ACS Chem. Neurosci. 3, 211 − 220 (2012). Copyright 2012 American Chemical Society. (c) Photocontrol of hydrogel viscoelasticity comprising of azobenzene-containing (RADA)4 peptide. Reproduced with permission from Doran et al., Polym. Chem. 5, 241–248 (2014). Copyright 2014 Royal Society of Chemistry. (d) Guest-release photocontrol from artificial viral capsid self-assembled from an azobenzene-containing β-annulus peptide. FCS showed the release of fluorescent-labeled dextrans from artificial viral capsid by UV irradiation. Reproduced with permission from Matsuura and Fujita, Int. J. Mol. Sci. 22, 4028 (2021). Copyright 2021 Authors, licensed under a Creative Commons Attribution (CC BY) License.

FIG. 3.

Photoisomerization of azobenzene controls peptide assemblies. (a) β-hairpin structure controlled by azobenzene-containing peptide photoisomerization. Reproduced with permission from Aemissegger et al., J. Am. Chem. Soc. 127, 2929–2936 (2005). Copyright 2005 American Chemical Society. (b) Cross-β amyloid fibril formation by photoisomerization of azobenzene-containing Amyloid-β mimic peptide. Reproduced with permission from Doran et al., ACS Chem. Neurosci. 3, 211 − 220 (2012). Copyright 2012 American Chemical Society. (c) Photocontrol of hydrogel viscoelasticity comprising of azobenzene-containing (RADA)4 peptide. Reproduced with permission from Doran et al., Polym. Chem. 5, 241–248 (2014). Copyright 2014 Royal Society of Chemistry. (d) Guest-release photocontrol from artificial viral capsid self-assembled from an azobenzene-containing β-annulus peptide. FCS showed the release of fluorescent-labeled dextrans from artificial viral capsid by UV irradiation. Reproduced with permission from Matsuura and Fujita, Int. J. Mol. Sci. 22, 4028 (2021). Copyright 2021 Authors, licensed under a Creative Commons Attribution (CC BY) License.

Close modal

Supramolecular hydrogels that are formed by peptide self-assembly are attractive materials for controlled drug release and regenerative medicine. Photoresponsive hydrogels have been developed by incorporating azobenzene photoswitches into the side chains or terminals of short self-assembling peptides.53–55 Doran et al. achieved reversible optical control of hydrogel viscoelasticity by incorporating the azobenzene turn mimetic (AMPP) into the center of a hydrogel-forming peptide (RADA)4 [Fig. 3(c)].56 The trans state formed a rigid self-supporting gel via an extended or bent β-sheet structure, while photoisomerization to the cis state reduced hydrogel rigidity, which was caused by disrupting the well-ordered assembly structure. Such photo-responding hydrogel of self-assembled peptides can be applied to peptide materials that allow cultured cell harvesting and encapsulated drug release by light irradiation.

A 24-residue β-annulus peptide (INHVGGTGGAIMAPVAVTRQLVGS), derived from tomato bushy stunt virus, self-assembled into hollow nanocapsules with sizes of 30–50 nm in water.57 A photochromic azobenzene linker, AMPP, replaced the putative turn region (Pro14-Val15-Ala16 sequence) of β-annulus peptide, which plays a pivotal role in artificial viral capsid formation [Fig. 3(d)].58 The trans state of β-annulus-azo peptide self-assembled into capsids with diameters of 30–50 nm at 25 μM close to the critical aggregation concentration, whereas the cis-rich state after UV irradiation formed micrometer-sized aggregates. Circular dichroism (CD) spectra revealed increased contents of the random coil by photoisomerization from the trans to the cis-form. Fluorescence correlation spectroscopy (FCS) showed that fluorescein isothiocyanate (FITC)-labeled dextran was encapsulated in the artificial viral capsid that comprises trans-rich β-annulus-azo peptide. The FCS curve shifted toward shorter diffusion times after UV irradiation, indicating the release of FITC-labeled dextran from the capsid. However, FITC-labeled dextran re-encapsulation in the capsid was not observed after sequential UV and visible-light irradiations, due to aggregate formation of the cis-rich β-annulus-azo peptide. This proof-of-concept provides a guideline to develop a novel photoinduced drug delivery system using the azobenzene-containing artificial viral capsids.

Photoisomerization of uncharged spiropyran and zwitterionic merocyanine is a photoswitch system that causes more drastic structural changes than azobenzene.46 Sendai et al. developed a photoreconfigurable supramolecular nanotube comprising spiropyran/merocyanine-modified GroEL, which is one of the molecular chaperones [Fig. 4(a)].59 GroEL mutant containing 14 Cys residues at each cavity entrance was modified with spiropyran/merocyanine-maleimide. The merocyanine-modified GroEL was self-assembled into micrometer-long nanotubes by the coordination of Mg2+ with merocyanine. The nanotubes dissociated to monomeric GroEL when the merocyanine-modified GroEL was photoisomerized into the spiropyran form by visible light irradiation. Conversely, the nanotubes were reconstructed when the spiropyran-modified GroEL was exposed to UV light to convert to merocyanine form. The GroEL cavity can encapsulate macromolecules or nanoparticles, thus the spiropyran/merocyanine-modified GroEL nanotube allows the development of a spatiotemporal control system of guest delivery. Liu et al. demonstrated that spiropyran-modified tetrapeptide Fmoc-KKSPKF-NH2 self-assembled into β-sheet nanofibers to form a hydrogel in acidic water under visible light exposure, while dissociated when the light source was removed [Fig. 4(b)].60 The spiropyran-modified peptide thermally isomerized into a more stable protonated merocyanine form and the nanofiber was dissociated due to increased electrostatic repulsion in the absence of continuous visible light irradiation. This light-driven dissipative self-assembly of peptides enables simple spatiotemporal control of hydrogel materials without fuel waste production.

FIG. 4.

Protein/peptide assemblies controlled by spiropyran/merocyanine photoisomerization. (a) Spiropyran/merocyanine-modified GroEL photoisomerization controls protein nanotube. Reproduced with permission from Sendai et al., J. Am. Chem. Soc. 135, 11509–11512 (2013). Copyright 2013 American Chemical Society. (b) Light-driven hydrogel self-assembled from spiropyran/merocyanine-modified peptide. Reproduced with permission from Liu et al., Chem. Commun. 57, 13776–13779 (2021). Copyright 2021 Royal Society of Chemistry. (c) Peptide nanofibers controlled by photoisomerization of spiropyran/merocyanine induced dramatic morphological changes in giant liposomes. Reproduced with permission from Liang et al., Front. Mol. Biosci. 10, 1137885 (2023). Copyright 2023 Authors, licensed under a Creative Commons Attribution (CC BY) License.

FIG. 4.

Protein/peptide assemblies controlled by spiropyran/merocyanine photoisomerization. (a) Spiropyran/merocyanine-modified GroEL photoisomerization controls protein nanotube. Reproduced with permission from Sendai et al., J. Am. Chem. Soc. 135, 11509–11512 (2013). Copyright 2013 American Chemical Society. (b) Light-driven hydrogel self-assembled from spiropyran/merocyanine-modified peptide. Reproduced with permission from Liu et al., Chem. Commun. 57, 13776–13779 (2021). Copyright 2021 Royal Society of Chemistry. (c) Peptide nanofibers controlled by photoisomerization of spiropyran/merocyanine induced dramatic morphological changes in giant liposomes. Reproduced with permission from Liang et al., Front. Mol. Biosci. 10, 1137885 (2023). Copyright 2023 Authors, licensed under a Creative Commons Attribution (CC BY) License.

Close modal

Cytoskeletons in eukaryotic cells, such as microtubules and actin filaments, dynamically control cellular morphology by reversible constituent protein polymerization/depolymerization. We developed an artificial cytoskeleton-like system, in which reversible polymerization/depolymerization of spiropyran/merocyanine-modified peptide nanofiber in giant unilamellar vesicles (GUVs) dramatically changed the morphology of GUVs with diameters that are near cell size [Fig. 4(c)].61 Spiropyran-modified peptide FKFECSPKFE formed β-sheet nanofibers, whereas merocyanine-form photoisomerization completely dissociated the nanofibers. Spherical GUVs that encapsulate the merocyanine-peptide FKFECMCKFE dramatically changed into worm-like vesicles by the photoisomerization to the spiropyran-form. UV and visible light irradiation reversibly controlled the dramatic morphological changes in GUVs. This photo-controlled system, which mimics the polymerization and depolymerization of the natural cytoskeleton, can be used as a component of molecular robots and artificial cell systems.

Moreover, several studies focused on photocontrol of self-assembly of peptides modified with various photoswitches. Poloni et al. revealed that photoisomerization of tryptophan zipper peptide modified with overcrowded alkene photoswitch generated two thermally stable isomers and controlled the formation of amyloid-like fibrils via β-hairpin structure.62 Marafon et al. developed photoisomerizable unsaturated β-amino acid, (Z/E)-3-aminoprop-2-enoic acid, to enable photo-control of self-assembly of peptide fordmars.63 Nakamura et al. showed the photo-triggered out-of-equilibrium pattern formation using peptide-type nanofiber modified with benzoylhydrazone group as photoswitch in self-sorting double network hydrogel.64 Light irradiation using a photomask followed by thermal incubation induced the spatially controlled peptide nanofiber condensation via metastable nanofiber photodecomposition in the irradiated areas and supply of monomers from the nonirradiated areas. Sun et al. developed a light-triggered platform that allows spatiotemporal control of self-assembly from nanoparticles into nanofibers in living cells for controlling cellular behaviors.65 A photoswitch 3-methylene-2-(quinolin-8-yl) isoindolin-1-one was modified with a peptide having an integrin binding sequence (RGD) and self-assembling sequence (SGKLVFF). Peptide conjugate photoisomerization caused a morphological change of the peptide assemblies on the cell surface from nanoparticles to nanofibers, which induced cell adhesion. Ji et al. showed that the morphology and gel properties of diphenylalanine assemblies can be modulated by light through coassembly with photochromic bipyridine derivatives.66,67

As described above, structural control of peptide assemblies by photoisomerization has greatly helped not only in controlling biological functions but also in developing novel drug delivery systems and bottom-up nanotechnology.

Peptide–DNA interactions are abundant in living cells, and controlling these interactions allows the control of various intracellular signals such as gene expression. Sequence-specific DNA-binding proteins, known as transcription factors, primarily regulate gene expression in eukaryotes. Basic leucine zipper (bZIP) transcription factors, such as yeast GCN4, are dimeric proteins that recognize dyadic and mostly palindromic DNA sites,68 and studies investigated the control of GCN4-bZIP binding to DNA by various external stimuli.69–73 Caamaño et al. developed a peptide conjugate, wherein two basic regions of bZIP protein were linked through azobenzene as photoswitch [Fig. 5(a)].74 The cis isomer bound to duplex DNA containing the recognition site with a high affinity of 60–70 times more efficiently than the trans isomer, due to the geometric proximity of peptides. CD spectra revealed an increased α-helix content of the cis isomer upon DNA binding, whereas the ellipticity increase was smaller for the trans isomer. However, the reversibility of DNA binding was incomplete, because the DNA binding inhibits the cis-to-trans photoisomerization and thermal isomerization process. Woolley et al. developed a GCN4-bZIP bridged with an azobenzene between Cys residues within the zipper domain to achieve the reversible photo-control of DNA-binding of GCN4-bZIP [Fig. 4(b)].75 The trans-form destabilizes the helical structure of the coiled-coil region of GCN4-bZIP, whereas photoisomerization to the cis-form increases helical content and substantially enhances DNA binding. Since the thermal relaxation to the trans state and concomitant dissociation of the protein−DNA complex readily occurred, the photo-control of DNA-binding was observed to reversible. Furthermore, Guerrero et al. achieved photo-control of DNA-binding of HDH-3, an 18-residue miniature engrailed homeodomain, by cross-linking azobenzene through two Cys residues [Fig. 4(c)].76 The trans-form of the azobenzene-bridged HDH-3 bound to the target DNA with higher affinity (Kd = 7.5 nM) than the cis-form after UV irradiation (Kd = 140 nM). The remarkable affinity of the trans-form with only 18 amino acid residues was most likely a consequence of the cross-linker-induced preorganization of its recognition helix.

FIG. 5.

Peptide-DNA interactions controlled by photoisomerization. (a) Photocontrol of DNA-binding of two basic regions of bZIP protein linked through azobenzene. Reproduced with permission from Caamaño et al., Angew. Chem. Int. Ed. 39, 3104–3107 (2000). Copyright 2000 Wiley-VCH. (b) Reversible photocontrol of DNA binding of azobenzene-bridged GCN4-bZIP. Reproduced with permission from Woolley et al., Biochemistry 45, 6075–6084 (2006). Copyright 2006 American Chemical Society. (c) DNA-binding photocontrol of azobenzene-bridged miniature helix peptide. Reproduced with permission from Guerrero et al., J. Am. Chem. Soc. 127, 15624–15629 (2005). Copyright 2005 American Chemical Society. (d) DNA-binding photocontrol of azobenzene-bridged coiled-coil proteins in living cells. Reproduced with permission from Zhang et al., Angew. Chem. Int. Ed. 49, 3943–3946 (2010). Copyright 2010 Wiley-VCH. (e) Diarylethene-bridged helix peptides to photoregulate DNA-binding. QCM sensorgrams of diarylethene-bridged helix peptides binding to DNA immobilized on plates. Reproduced with permission from Fujimoto et al., Chem. Eur. J. 18, 9834–9840 (2012). Copyright 2012 Wiley-VCH.

FIG. 5.

Peptide-DNA interactions controlled by photoisomerization. (a) Photocontrol of DNA-binding of two basic regions of bZIP protein linked through azobenzene. Reproduced with permission from Caamaño et al., Angew. Chem. Int. Ed. 39, 3104–3107 (2000). Copyright 2000 Wiley-VCH. (b) Reversible photocontrol of DNA binding of azobenzene-bridged GCN4-bZIP. Reproduced with permission from Woolley et al., Biochemistry 45, 6075–6084 (2006). Copyright 2006 American Chemical Society. (c) DNA-binding photocontrol of azobenzene-bridged miniature helix peptide. Reproduced with permission from Guerrero et al., J. Am. Chem. Soc. 127, 15624–15629 (2005). Copyright 2005 American Chemical Society. (d) DNA-binding photocontrol of azobenzene-bridged coiled-coil proteins in living cells. Reproduced with permission from Zhang et al., Angew. Chem. Int. Ed. 49, 3943–3946 (2010). Copyright 2010 Wiley-VCH. (e) Diarylethene-bridged helix peptides to photoregulate DNA-binding. QCM sensorgrams of diarylethene-bridged helix peptides binding to DNA immobilized on plates. Reproduced with permission from Fujimoto et al., Chem. Eur. J. 18, 9834–9840 (2012). Copyright 2012 Wiley-VCH.

Close modal

The reversible DNA-binding photocontrol could pave the way for studies of cellular processes through transcription control. Zhang et al. revealed that the introduction of a designed azobenzene-bridged dominant negative peptide (XAFosW) controlled the activity of a coiled-coil protein, the AP-1 transcription factor, in living cells [Fig. 4(d)].77 The photoisomerization to the cis-form enhanced helicity and induced strand exchange of coiled-coil to dissociate AP-1 transcription factor (cFos/cJun) from DNA. The AP-1 activity in living cells was inhibited with the XAFosW based on UV light irradiation.

Photocontrol using other DNA-binding motifs is being considered. Murawska et al. developed an AT-hook binding motif peptide incorporated azobenzene linker as a minor groove binder and a zink-finger peptide incorporated azobenzene linker as a major groove binder.78 Photochemical isomerization in each case affected DNA binding, as determined in fluorescent displacement assays on model DNA strands, which provides promising tools for DNA modulation. Photochromic molecules, other than azobenzene, have also been used for peptide-DNA interaction photocontrol. Azobenzenes have an undesirable property for biological use, such as thermal isomerization between the two photoisomers, but diarylethene never isomerizes without photoirradiation. Fujimoto et al. developed peptides possessing diarylethene-bridged and DNA-binding regions and achieved DNA-binding photocontrol without thermal isomerization [Fig. 4(e)].79 The open-form of the diarylethene-bridged peptide qk̂Dae folded into stable α-helices, whereas the closed-form after UV irradiation destabilized the helical structures. Quartz-crystal microbalance (QCM) analysis revealed that the open isomer is strongly associated with a target DNA compared with the closed one.

Specific peptides have a great affinity to target proteins to inhibit or activate their activities. Photochromic molecule conjugation to the ligand peptides is a promising approach to photocontrol target protein activities by changing the interactions between the peptides and proteins. Two groups (Liu et al.80 and Hamachi et al.81) developed the concept using ribonuclease S, an RNA hydrolyzing enzyme, which consists of S-peptide and S-protein fragments. The reconstitution efficiency of the mutated S-peptide to S-protein changed in the cis- and trans-states by incorporating an azobenzene moiety to S-peptide, thereby enabling the reversible photocontrol of enzymatic activity of the reconstituted ribonuclease S by light irradiation.81 Currently, incorporating photochromic molecules can control various types of protein activities, as shown below.

Volgraf et al. developed azobenzene-conjugated peptide agonists of ionotropic glutamate receptors that are crucial mediators of excitatory information transfer in the central nervous system to photocontrol the ion channel activities.82,83 In their first approach, the azobenzene-conjugated agonist covalently tethered to the mutated cysteine residue of the target ionotropic glutamate receptor, iGluR6, to open/close the channel,82 as a similar strategy to photocontrol the potassium channel.84 In their second approach, non-tethered azobenzene-conjugated agonists that can be used for wild-type ionotropic glutamate receptors, iGluR5 and iGluR6, were developed [Fig. 6(a)].83 The azobenzene-conjugated glutamate analog showed channel activation in the trans-form compared to the cis-form in the whole-cell voltage clamp assay in HEK293 cells that express iGluR6. Further, the system was used to reversibly activate iGluR6 channels in cultured rat hippocampal neurons by switching wavelengths of light between 380 and 500 nm [Fig. 6(a)]. Mafy et al. developed an azobenzene-conjugated peptide inhibitor of centromere-associated protein E (CENP-E), which is a mitotic kinesin used for chromosome transportation [Fig. 6(b)].85,86 The design was based on the known CENP-E inhibitor that prevents chromosome alignment and mitotic progression. The azobenzene-conjugated peptide inhibitor showed ∼10-fold higher inhibition CENP-E activity in the trans-form compared to the cis-form. This system was used to reversibly photocontrol the CENP-E-dependent chromosome congression and mitotic progression by switching light wavelengths (365 and 510 nm). Not only the azobenzene moiety but also the thioxylation (O/S exchange) of the peptide bond could be used for the cis- and trans-photoisomerization toward photocontrol of the binding to target proteins.87 

FIG. 6.

Photoisomerization controls peptide-protein interactions. (a) Photochromic glutamate analog for glutamate receptor agonist and whole-cell current clamp recording. Reproduced with permission from Volgraf et al., J. Am. Chem. Soc. 129, 260 − 261 (2007). Copyright 2007 American Chemical Society. (b) Photoswitchable CEMP-E inhibitor (a mitotic kinesin) regulating mitotic progression. Reproduced with permission from Mafy et al., J. Am. Chem. Soc. 142, 1763 − 1767 (2020). Copyright 2020 American Chemical Society. (c) Insulin secretion control by a photoswitchable incretin. Reproduced with permission from Broichhagen et al., Angew. Chem. Int. Ed. 54, 15565–15569 (2015). Copyright 2015 Wiley-VCH. (d) Photoswitchable MLL1-WDR5 interaction inhibitors control MLL1 methyltransferase activity. Reproduced with permission from Albert et al., Chem. Sci. 8, 4612–4618 (2017). Copyright 2017 Royal Society of Chemistry. (e) Photoswitching peptide inhibitors to β-appendage of AP2 regulating clathrin-mediated endocytosis. Reproduced with permission from Nevola et al., Angew. Chem. Int. Ed. 52, 7704 − 7708 (2013). Copyright 2013 Wiley-VCH. (f) Photoswitchable histone methyltransferase MLL1 inhibitor controlling hematopoiesis in zebrafish. Reproduced with permission from Albert et al., ACS Cent. Sci. 8, 57−66 (2022). Copyright 2022 American Chemical Society.

FIG. 6.

Photoisomerization controls peptide-protein interactions. (a) Photochromic glutamate analog for glutamate receptor agonist and whole-cell current clamp recording. Reproduced with permission from Volgraf et al., J. Am. Chem. Soc. 129, 260 − 261 (2007). Copyright 2007 American Chemical Society. (b) Photoswitchable CEMP-E inhibitor (a mitotic kinesin) regulating mitotic progression. Reproduced with permission from Mafy et al., J. Am. Chem. Soc. 142, 1763 − 1767 (2020). Copyright 2020 American Chemical Society. (c) Insulin secretion control by a photoswitchable incretin. Reproduced with permission from Broichhagen et al., Angew. Chem. Int. Ed. 54, 15565–15569 (2015). Copyright 2015 Wiley-VCH. (d) Photoswitchable MLL1-WDR5 interaction inhibitors control MLL1 methyltransferase activity. Reproduced with permission from Albert et al., Chem. Sci. 8, 4612–4618 (2017). Copyright 2017 Royal Society of Chemistry. (e) Photoswitching peptide inhibitors to β-appendage of AP2 regulating clathrin-mediated endocytosis. Reproduced with permission from Nevola et al., Angew. Chem. Int. Ed. 52, 7704 − 7708 (2013). Copyright 2013 Wiley-VCH. (f) Photoswitchable histone methyltransferase MLL1 inhibitor controlling hematopoiesis in zebrafish. Reproduced with permission from Albert et al., ACS Cent. Sci. 8, 57−66 (2022). Copyright 2022 American Chemical Society.

Close modal

The introduction of photochromic molecules to the backbone of peptides is an attractive approach for inducing dynamic structural changes in peptides. Azobenzene-based backbones were frequently used for this purpose due to their synthetic simplicity and commercial availability. Hodson and Trauner developed incretin mimetics that contain the azobenzene backbone to photocontrol the binding to the target glucagon-like peptide-1 receptor (GLP-1R) resulting in insulin secretion [Fig. 6(c)].88 They introduced azobenzene moiety as a bridge between two α-helixes of the original GLP-1R agonist and evaluated the effects of the peptide (LirAzo) on the cellular GLP-1R signaling of Ca2+ and cyclic AMP (cAMP) followed by insulin secretion. The trans-LirAzo induced the increment of Ca2+ influx compared to cis-LirAzo, whereas cis-LirAzo induced cAMP generation compared to trans-LirAzo. With different signaling, cis-LirAzo augmented glucose-induced insulin secretion compared to trans-LirAzo, whereas trans-LirAzo demonstrated a significant protective effect against glucolipotoxicity compared to cis-LirAzo. The azobenzene backbone strategy was used for photoswitchable inhibition of mixed-lineage leukemia 1 (MLL1) methyltransferase core complex [Fig. 6(d)].89 The azobenzene-conjugated peptide inhibitors were designed for the photocontrol of the methylation activity because the binding of MLL1 and WD40-repeat protein 5 (WDR5) is crucial for the methylation activity. A peptide that demonstrated a higher inhibition activity in the trans-form compared to the cis-form and the original WDR5-interacting peptide was found by optimizing the position of the azobenzene that was introduced to the WDR5-interacting peptide. The in vitro assay showed that the trans-form demonstrated a stronger MLL1 methyltransferase activity inhibition compared to the cis-form [Fig. 6(d)]. The system was used for suppressing MLL1-target gene transcription and inhibiting leukemia cell growth by incorporating octa-arginine (a common cell-penetrating peptide) into the peptide. In another example, reversible photocontrol of cell adhesion was achieved by introducing an azobenzene backbone between a cell adhesive RGD peptide and alkanethiols that was used for substrate immobilization.90 

Light-driven change of secondary peptide structures is also useful to control the interaction with target proteins. Hence, stapled peptides that are linked by azobenzene moieties were developed. Kneissl et al. designed α-helical peptides stapled with azobenzene moiety for photoswitching the secondary peptide structures by cis- and trans-isomerization that are targeted to an anti-apoptotic protein, Bcl-xL.91 Nevola et al. used a similar strategy for the photocontrol of protein–protein interaction of the AP2 complex that is involved in the clathrin-mediated endocytosis [Fig. 6(e)].92 The design was based on the C-terminal β-arrestin peptide that binds to the β-appendage of AP2 with an α-helical structure. Several peptides that adopt a helical structure and bind to β-appendage of AP2, preferably in the trans-form or cis-form were obtained, by stapling the peptide using an azobenzene crosslinker by the reaction with cysteine residues that were introduced at different positions. These peptides provided “stop” and “go” signals to the clathrin-mediated endocytosis by photocontrolling the binding to β-appendage of AP2. In particular, one azobenzene-stapled peptide inhibited the clathrin-mediated endocytosis of transferrin efficiently in the trans-form by slowing down the membrane traffic, as observed in the light-dependent increased/decreased numbers of clathrin-coated pits [Fig. 6(e)]. The azobenzene-stapled peptide strategy was also used to photocontrol the interaction between S-peptide and S-protein.93 Azobenzene-conjugated cyclic peptides were recently developed as photoswitching inhibitors of the histone methyltransferase MLL1 complex that has various hematopoietic roles, such as a chromatin-modifier and a potent oncogenic driver toward hematopoietic control in zebrafish [Fig. 6(f)].94 The design was based on a cyclic peptide, MM-401, the potent inhibitor of MLL1 complex, including MLL1, WDR5, RbBP5, Ash2L, and DPY30. One of the designed azobenzene-tethered peptides strongly bound to WDR5 as a potential MLL1 methylation activity inhibitor in the cis-form than the trans-form. This system was used to photocontrol the hematopoiesis of zebrafish. Peptide treatment in the cis-form resulted in a lack of responsivity upon tail touch in the larvae and induced abnormal developmental phenotypes (curved body axis and heart edema), whereas the trans-form demonstrated no apparent effects [Fig. 6(f)]. Additionally, larvae incubation with the trans-form followed by light irradiation (520 nm for 30 s) and subsequent incubation for 18 h showed comparable effects to the cis-form treatment. Another unique approach is the cross-linking of two enzyme sites using azobenzene moiety to photocontrol the enzymatic activity. Schierling et al. utilized this approach to photocontrol the activity of the homodimeric restriction endonuclease PvuI, where the dramatic activity change was observed by introducing two azobenzene moieties.95 

These examples demonstrated that incorporating photochromic molecules into peptides is a promising approach to photocontrol peptide interactions to target proteins in vitro and in vivo.

Cytoskeletons, such as microtubules and actin filaments, are crucial cell components. Controlling the structures and functions of the cytoskeletons by light is a promising approach for manipulating cell functions because of their important functions in cells, such as structural support, cell division, cell migration, and intracellular transport. The major strategy of controlling the cytoskeletons by light is photochromic molecule conjugation to the cytoskeleton-targeted drugs or the design of photochromic molecules mimicking the drugs, which was developed by Borowiak et al. Their first report designed photostatins, which show reversible cistrans isomerization by blue and green light, based on a microtubule-destabilizing drug, colchicine.96 Photostatins show a stronger microtubule-disrupting effect in cis-state compared to trans-state, which can be used to control microtubule dynamics in cells. They later reported the detailed analysis and in vivo applications of photostatins.97 Afterward, various types of photoswitches were developed for microtubule98–106 and actin filament control,107–109 mainly by photochromic molecule conjugation such as azobenzenes to the cytoskeleton-targeted drugs. Additionally, a tubulin-specific photoswitchable fluorescent probe was developed by conjugating the spiropyran moiety to colchicine unit.110 

The cytoskeletons, in addition to cellular applications, can be used as components of dynamic nanomaterials. Microtubules are frequently used for combining with motor proteins (kinesin and dynein) to construct ATP-driven motile nanomaterials. The control of movement, stability, and assembling properties of microtubules are important for constructing sophisticated motile systems. Keya et al. developed light-responsive control of microtubule swarming by conjugating photoresponsive DNA-containing azobenzene groups to the outer microtubule surface.111 In this case, microtubule swarm formation was promoted in the trans-state and inhibited in the cis-state of the azobenzene moiety. New types of mechanical robots include the reversible association/dissociation of microtubule swarms by light. Later, the mechanical properties of microtubules112 and cargo delivery113 were controlled using the photoresponsive DNA-conjugated microtubules.

Recently, peptides that are conjugated with photoisomerization moiety were developed to control the structures and dynamics of actin filaments and microtubules. Borowiak et al. developed light-dependent actin filament effectors by conjugating azobenzene to actin filament-stabilizing marine depsipeptide jasplakinolide [Fig. 7(a)].107 Among the designed molecules, termed optojasps, optojasp-1 with the shortest linker demonstrated cytotoxicity in cis-form compared to trans-form. Additionally, the cis-form of optojasp-1 increased actin nucleation and caused large aggregate formation, such as actin filament stabilizer, including phalloidin, whereas the trans-form demonstrated no such effect. The irradiation of 390 nm of light to the optojasp-1-treated cells in the trans-state for 5 h caused actin filament aggregation [Fig. 7(a), bottom left]. The irradiation of 390 nm of light and the subsequent irradiation of 475 nm of light for 1 h followed by incubation under a dark environment for 12 h recovered the regular actin filament structures [Fig. 7(a), bottom right]. Thus, these structural changes of actin filaments were reversibly controlled by light irradiation. This system was also used to photocontrol the cellular dynamics related to actin filaments, including cell shapes, motility, and division. Later, the cryo-EM structures of actin filaments with optojasps were obtained.108 Additionally, new types of optojasps that can be activated with longer wavelengths in the visible range (e.g., 440–477 nm) were developed.109 These molecular tools help understand and control the dynamics of actin filaments.

FIG. 7.

Control of cytoskeletons by photochromic peptides. (a) Control of actin filament dynamics by azobenzene-conjugated depsipeptide. Reproduced with permission from Borowiak et al., J. Am. Chem. Soc. 142, 9240 − 9249 (2020). Copyright 2020 American Chemical Society. (b) Peptide-based photoswitch for the control of microtubule dynamics. Reproduced with permission from Kirchner et al., Nat. Commun. 13, 6066 (2022). Copyright 2022 Springer Nature. (c) Spiropyran-conjugated TP reversibly photocontrol microtubule structures. Reproduced with permission from Inaba et al., ChemBioChem 24, e202200782 (2023). Copyright 2023 Wiley-VCH.

FIG. 7.

Control of cytoskeletons by photochromic peptides. (a) Control of actin filament dynamics by azobenzene-conjugated depsipeptide. Reproduced with permission from Borowiak et al., J. Am. Chem. Soc. 142, 9240 − 9249 (2020). Copyright 2020 American Chemical Society. (b) Peptide-based photoswitch for the control of microtubule dynamics. Reproduced with permission from Kirchner et al., Nat. Commun. 13, 6066 (2022). Copyright 2022 Springer Nature. (c) Spiropyran-conjugated TP reversibly photocontrol microtubule structures. Reproduced with permission from Inaba et al., ChemBioChem 24, e202200782 (2023). Copyright 2023 Wiley-VCH.

Close modal

Kirchner et al. used a peptide-derived photoswitchable motif, plinabulin (a tubulin polymerization inhibitor), as a photochromic molecule to optically control microtubules [Fig. 7(b)].114 Plinabulin consists of a cyclic dipeptide core with the arylidene residue [Fig. 7(b)]. They focused on the use of plinabulin as a photochromic molecule based on the E/Z-photoisomerization of arylidene substituents in other examples. The reversible E/Z-photoisomerization of plinabulin was observed, which revealed the dramatically different cytotoxicity against HT-29 human colon cancer cells between Z-form (thermodynamically stable form, IC50 = 0.47 nM) and E-form (metastable form, IC50 = 92 nM). The Z-form of plinabulin induced abnormal tubulin aggregation and mitotic arrest in several cells, whereas the E-form demonstrated regular cell division as mitotic spindles were observed. The derivatives and the shorter version of plinabulin showed different light sensitivity, reversibility, and cytotoxicity, indicating that the chemical structure modification is useful to optically controlling microtubule dynamics leading to cell manipulation.

Spiropyran was conjugated to Tau-derived peptide (TP), which was bound to the inner pocket of microtubules, to reversibly alter microtubule configurations by light [Fig. 7(c)].115 Spiropyran was conjugated to the cysteine residues that were introduced to the different TMR-labeled TP positions. The peptide with spiropyran at the N-terminus (TMR-1-TP-SP) and the merocyanine form (TMR-1-TP-MC), among the four peptides with different positions of spiropyran, showed the largest difference in microtubule structures prepared with GTP, as TMR-1-TP-SP stabilized the microtubule structures but not TMR-1-TP-MC [Fig. 7(c)]. The structural change of microtubules by TMR-1-TP-SP/TMR-1-TP-MC was reversible and repetitive by UV light (365 nm) and visible light (505 nm) irradiation. The photocontrol of microtubule structures using a reversible photoisomerization system is useful for the spatiotemporal manipulation of cell fates.

Photocleavage reactions, including deprotection of the protecting group and cleavage of the peptide backbone, are a powerful tool to activate or deactivate functional peptides and proteins with high spatiotemporal resolution. Most photocleavage of peptide backbones is based on intramolecular cyclization of the photoexcited side chain group followed by amide bond cleavage [Fig. 2(b)]. Peptide backbone that incorporated 2-nitrophenylalanine and developed by Peters et al. is readily cleaved by UV irradiation via intramolecular cyclization.116 Furthermore, they revealed that genetically engineered protein containing 2-nitrophenylalanine was site-specifically photocleaved. Taniguchi et al. developed a photo-triggered production system of Aβ1 − 42 via photocleavage of the protecting group (6-nitroveratryloxycarbonyl) followed by O−N intramolecular acyl migration.117 Katayama et al. created photocleavable peptides with a bromocoumarin linker having a high photolytic efficiency.118 Shigenaga et al. developed photocleavable peptides that can release a functional peptide after the stimulus-induced removal of a phenolic protective group and subsequent cleavage of peptide backbone through nucleophilic attack of phenolic hydroxyl group to an adjacent peptide bond.119–121 The system was used for the photo-responding release of nucleocytoplasmic shuttle peptide to control intracellular localization with light.119 Mangubat-Medina et al. developed a new photocaging method using histidine-directed backbone modification to selectively modify peptides and proteins at the amide N–H bond, which allows the photorelease of the backbone modification and function restoration.122 

Photocleavage123–128 of the peptide backbone or photodeprotection129–133 from the side chain described in Sec. III A can trigger spatiotemporal-controlled peptide self-assembly. Photo-induced elimination of units that inhibit the self-assembly of peptides can initiate the self-assembly with controlled timing. Bosques and Imperiali pioneered such a photo-induced peptide assembling system using peptide having a fibril inhibitory unit and amyloidogenic unit linked by a photocleavable amino acid [Fig. 8(a)].123 The peptide was not assembled due to electrostatic repulsion among polycationic N,N-dimethylethylenediamine units, while afforded fibrous assemblies after photolysis and subsequent incubation for several hours. Haines et al. developed a light-activated peptide hydrogel formation system using peptide photo-deprotection [Fig. 8(b)].129 MAX7CNB, a 20-residue peptide, is protected with α-carboxy-2-nitrobenzyl group at Cys residue and did not self-assemble due to the unfold conformation. UV irradiation released the protecting group and triggered peptide folding to produce amphiphilic β-hairpins that self-assemble into viscoelastic hydrogel material. The formed gel surface was non-cytotoxic for NIH 3T3 fibroblast cells, conducive to cell adhesion, and allowed cell migration. Recently, Xiang et al. developed a novel photo-triggered peptide hydrogel photocaged by a positively charged dipeptide.124 The photoactivated hydrogel was suitable for two-dimensional (2D) and 3D cell cultures, and its photo-controllable mechanical strength could regulate the spreading of stem cells on its surface.

FIG. 8.

Peptide assemblies induced by photocleavage. (a) Temporal control of peptide fibrillization by photocleavage. Reproduced with permission from Bosques and Imperiali, J. Am. Chem. Soc. 125, 7530–7531 (2003). Copyright 2003 American Chemical Society. (b) Light-activated hydrogel formation via β-hairpin peptide folding and self-assembly triggered by the photo-deprotection. Reproduced with permission from Haines et al., J. Am. Chem. Soc. 127, 17025–17029 (2005). Copyright 2005 American Chemical Society. (c) Spatiotemporal control of peptide nanofiber growth triggered by photocleavage of a DNA-conjugated β-sheet forming peptide. Reproduced with permission from Furutani et al., Chem. Commun. 51, 8020–8022 (2015). Copyright 2015 The Royal Society of Chemistry. (d) Light-induced propulsion of a giant liposome driven by peptide nanofibre growth. The tracking trajectories of giant liposomes by UV light irradiation. Reproduced with permission from Inaba et al., Sci. Rep. 8, 6243 (2018). Copyright 2018 Authors, licensed under a Creative Commons Attribution (CC BY) License. (e) Light-induced peptide nanofiber growth derived negative phototaxis of nucleo-spheres. The tracking trajectories of nucleo-spheres by UV light irradiation. Reproduced with permission from Inaba et al., ACS Appl. Bio Mater. 4, 5425–5434 (2021). Copyright 2021 American Chemical Society.

FIG. 8.

Peptide assemblies induced by photocleavage. (a) Temporal control of peptide fibrillization by photocleavage. Reproduced with permission from Bosques and Imperiali, J. Am. Chem. Soc. 125, 7530–7531 (2003). Copyright 2003 American Chemical Society. (b) Light-activated hydrogel formation via β-hairpin peptide folding and self-assembly triggered by the photo-deprotection. Reproduced with permission from Haines et al., J. Am. Chem. Soc. 127, 17025–17029 (2005). Copyright 2005 American Chemical Society. (c) Spatiotemporal control of peptide nanofiber growth triggered by photocleavage of a DNA-conjugated β-sheet forming peptide. Reproduced with permission from Furutani et al., Chem. Commun. 51, 8020–8022 (2015). Copyright 2015 The Royal Society of Chemistry. (d) Light-induced propulsion of a giant liposome driven by peptide nanofibre growth. The tracking trajectories of giant liposomes by UV light irradiation. Reproduced with permission from Inaba et al., Sci. Rep. 8, 6243 (2018). Copyright 2018 Authors, licensed under a Creative Commons Attribution (CC BY) License. (e) Light-induced peptide nanofiber growth derived negative phototaxis of nucleo-spheres. The tracking trajectories of nucleo-spheres by UV light irradiation. Reproduced with permission from Inaba et al., ACS Appl. Bio Mater. 4, 5425–5434 (2021). Copyright 2021 American Chemical Society.

Close modal

We developed a spatiotemporal control system of peptide nanofiber growth by photocleavage of a DNA-conjugated β-sheet forming peptide linked through a photoresponsive amino acid residue [Fig. 8(c)].125 Single-strand DNA (dA20) was used not only as the β-sheet disturbing unit because of its electrostatic repulsion but also addressing unit on complementary dT20-immobilized materials by its hybridization. UV light irradiation to the conjugate caused the cleavage of the peptide bond to generate the phenolic intermediate and subsequent intramolecular cyclization and then releases free FKFEFKFE peptide, which self-assembles to form nanofibers. The photo-triggered nanofiber growth selectively occurred on the complementary DNA (dT20)-immobilized glass substrate, while the nanofiber formation was minimally observed on the non-complementary DNA substrate even after UV irradiation. Afterward, we demonstrated that the photo-triggered peptide nanofiber growth on asymmetric giant liposomes promoted the translational motion, as a functional mimicking system of “actin comet tail” [Fig. 8(d)].126 An improved DNA-peptide conjugate that is photo-cleaved faster was immobilized onto phase-separated giant liposomes using streptavidin–biotin interaction and DNA hybridization. UV light irradiation greatly enhanced the translational movement of the conjugate–modified giant liposomes and the movement was sustained even after UV light irradiation ceased, whereas the translational motion was suppressed without UV light irradiation. Force generation due to peptide nanofiber growth and the Marangoni effect due to the surface tension gradient between the nanofiber-forming side and the opposite side are the possible driving forces of propulsion.

In nature, various phototactic microorganisms move toward (positive phototaxis) and away (negative phototaxis) from the light source. Recently, we demonstrated that DNA microspheres equipped with the photo-triggered nanofiber growth system showed negative phototaxis [Fig. 8(e)].127 DNA-peptide conjugate that was connected by a photocleavage site was immobilized on a biotin-modified nucleo-sphere that was self-assembled from DNA three-way junctions bearing self-complementary sticky-ends. UV light irradiation induced the asymmetric peptide nanofiber growth on the nucleo-sphere surface and the directional movement away from the light source. Peptide nanofiber growth could be induced only on the light-irradiated side and not the opposite side because the interior of the nucleo-spheres is filled with DNA and has a light-scattering property, resulting in negative phototaxis. The artificial phototactic system based on peptide nanofiber formation will not only be useful for understanding phototactic microorganisms in nature but may also result in future applications for microrobot construction.

Peptide photocleavage was used to photocontrol the biological functions, such as ligand binding to targets,134–145 cell adhesion/death,146–152 and intracellular localization.153–155 The photocleavage reaction is irreversible in contrast to the reversible structural change of photochromic molecules, but peptide cleavage possibly induces large structural changes. Examples of the photocleavage strategy to photocontrol the biological functions in vitro and in vivo are described below.

Incorporation of the photochromic moieties into the peptides is widely used to photocontrol the binding to target proteins. In particular, peptide ligands were caged with photocleavage moieties to decrease the binding affinity to target proteins, whereas the photocleavage reaction recovered the binding property to target proteins.134,135 In another approach, the peptide-based ligands with photocleavage moiety inhibited the activity of kinase136 and antibody,137 where the photocleavage reactions released the peptides and restored targeted protein activities. Additionally, the photocleavage technique was used to analyze the binding kinetics of peptides to target proteins.138 A similar concept was used to construct photocaged antibody fragments (nanobody) to photocontrol the binding to the antigen [Fig. 9(a)].139 The photocaged tyrosine variants, in this example, were incorporated into the key residue for the binding to target GFP in an anti-GFP nanobody. Compared to the parent nanobody, the photocaged nanobody exhibited ∼10 000-fold impaired binding affinity. The light-induced binding of sfGFP to the photocaged nanobody in living cells was performed using this system [Fig. 9(a)]. The caging strategy was used for light-activated covalent inhibitors to photocontrol cell cycle and apoptosis [Fig. 9(b)].140 MG132, a covalent proteasome inhibitor peptide, was photocaged at the reactive aldehyde to inhibit the activity. Irradiation with 405 nm of light released the de-caged MG132 that bound to the proteasome, resulting in cell cycle arrest in metaphase and apoptosis [Fig. 9(b)]. The photocaging strategy of protein-binding peptides was used for other cellular applications, such as kinase photoactivation in living cells141 and enzymatic hydrogel patterning used for the 3D invasion of primary human mesenchymal stem cells.142 

FIG. 9.

Peptide photocleavage controls biological functions. (a) Light-activatable binding of nanobody to its antigen. Reproduced with permission from Jedlitzke et al., Angew. Chem. Int. Ed. 59, 1506−1510 (2020). Copyright 2020 Wiley-VCH. (b) Light-activatable binding of covalent inhibitor, MG132, to proteasome and cell apoptosis. Reproduced with permission from Uhl et al., Angew. Chem. Int. Ed. 60, 1187−1196 (2021). Copyright 2021 Wiley-VCH. (c) Light-induced ligand exchange with a photocleavable peptide ligand. Reproduced with permission from Celie et al., J. Am. Chem. Soc. 131, 12298−12304 (2009). Copyright 2009 American Chemical Society. (d) Light-induced DNA binding of peptide dimer. Reproduced with permission from Jiménez-Balsa et al., Angew. Chem. Int. Ed. 51, 8825−8829 (2012). Copyright 2012 Wiley-VCH. (e) Photolytic RGDS peptide removal from peptide amphiphile nanofibers and effects on the cell adhesion. Reproduced with permission from Sur et al., ACS Nano 6, 10776−10785 (2012). Copyright 2012 American Chemical Society. (f) Light-induced activation of cell adhesion of caged cyclic RGD peptide on hydrogels and patterning of in vivo cell adhesion. Reproduced with permission from Lee et al., Nat. Mater. 14, 352−360 (2015). Copyright 2015 Springer Nature.

FIG. 9.

Peptide photocleavage controls biological functions. (a) Light-activatable binding of nanobody to its antigen. Reproduced with permission from Jedlitzke et al., Angew. Chem. Int. Ed. 59, 1506−1510 (2020). Copyright 2020 Wiley-VCH. (b) Light-activatable binding of covalent inhibitor, MG132, to proteasome and cell apoptosis. Reproduced with permission from Uhl et al., Angew. Chem. Int. Ed. 60, 1187−1196 (2021). Copyright 2021 Wiley-VCH. (c) Light-induced ligand exchange with a photocleavable peptide ligand. Reproduced with permission from Celie et al., J. Am. Chem. Soc. 131, 12298−12304 (2009). Copyright 2009 American Chemical Society. (d) Light-induced DNA binding of peptide dimer. Reproduced with permission from Jiménez-Balsa et al., Angew. Chem. Int. Ed. 51, 8825−8829 (2012). Copyright 2012 Wiley-VCH. (e) Photolytic RGDS peptide removal from peptide amphiphile nanofibers and effects on the cell adhesion. Reproduced with permission from Sur et al., ACS Nano 6, 10776−10785 (2012). Copyright 2012 American Chemical Society. (f) Light-induced activation of cell adhesion of caged cyclic RGD peptide on hydrogels and patterning of in vivo cell adhesion. Reproduced with permission from Lee et al., Nat. Mater. 14, 352−360 (2015). Copyright 2015 Springer Nature.

Close modal

Photocleavable peptides were used for light-induced ligand exchange, mainly for ligand binding in situ analysis.143,144 In particular, the photocleavable peptide ligand was developed for analyzing the ligand binding to MHC class I protein crystals [Fig. 9(c)].143 UV light-induced cleavage of the peptide in the MHC class I protein crystals released the lytic fragment and allowed for new full-length peptide binding. The in crystallo ligand exchange could be used for the high-throughput structural determination of the protein–ligand complexes. Not only the binding of peptide ligands but also the binding of DNA was also controlled using the photocleavable peptides. The N-terminal positively charged region of DNA-binding GCN4 transcription factor that was connected to negatively charged eight Glu residues through a photocleavable amino acid was designed [Fig. 9(d)].145 The dimerized peptide demonstrated a weak binding to target DNA due to the electrostatic repulsion between the negatively charged Glu moiety and DNA. The Glu moiety was released and the DNA-binding property was restored upon UV light irradiation.

Cell adhesion photocontrol using the photocleavable peptides immobilized on the substrates was widely explored. One strategy includes light-induced cell detachment or decreased cell adhesion.146–148 Sur et al. developed self-assembling peptide amphiphiles connected with cell adhesion epitope RGDS through photocleavable units as a light-responsive synthetic matrix that mimics natural extracellular matrixes [Fig. 9(e)].147 The self-assembling nanofiber matrix demonstrated increased cell spreading and mature focal adhesion, whereas UV light treatment inhibited the cell spreading due to the release of RGDS peptides. The opposite strategy is the light-induced cell adhesion using caged RGD peptide on the substrate surface.149–151 Lee et al. used this strategy for in vivo control of cell adhesion [Fig. 9(f)].151 They designed poly(ethylene glycol) di-acrylate hydrogels that present caged cyclic RGD peptides where the carboxylic group of the Asp residue was photocaged. UV light-induced cell adhesion was observed due to active cyclic RGD peptide presentation using the hydrogels as substrates. Hydrogel implantation and transdermal UV light irradiation induced cell adhesion on the exposure spot [Fig. 9(f)], demonstrating the spatiotemporal control of cell adhesion in vivo. The immobilization of photocleavable peptides on the substrates was also used for peptide cytotoxicity analysis.152 

Photocrosslinking of peptides at the suitable position is expected to stabilize the secondary structure, promote self-assembly, and reinforce mechanical rigidity. Rughani et al. revealed that diene polymerization remarkably enhanced the mechanical rigidity of hydrogel self-assembled from an amphiphilic β-hairpin peptide that was modified with diene groups at the Lys side chain [Fig. 10(a)].156 Rheological experiments reveal that irradiation caused intrafibrillar cross-linking and a 2.5-fold increase in mechanical rigidity. This peptide hydrogel is expected to be used as an injectable material whose mechanical properties can be easily modulated post-delivery in vivo. Ding et al. revealed that ruthenium-complex-catalyzed photocrosslinking dramatically enhanced the mechanical stability of small tyrosine-containing peptide hydrogels by 104-fold [Fig. 10(b)].157 The mechanical stability was enhanced due to the formation of a densely entangled fibrous network of peptide dimers through a dityrosine linkage. Peptide-based hydrogels frequently have low mechanical stability with storage moduli of 10–1000 Pa that greatly hinder their practical application, but the storage modulus of this crosslinked peptide hydrogel was approximately 100 kPa, which is one of the highest reported so far for hydrogels that are made of small peptide molecule. Recently, Pugliese et al. created a peptide hydrogel with pH-switchable on–off luminescence by ruthenium-complex-catalyzed photocrosslinking of a tyrosine-containing peptide, which is potentially applied for biomedical imaging, pH sensing, photonics, soft electronics, and bioprinting.158 Furthermore, photocrosslinking has become a powerful tool in chemical biology for identifying and mapping stable or transient interactions between biomacromolecules. Smith and co-workers used mass spectra of peptide photocrosslinking to explore the initial oligomeric structure of amyloid fibrils [Fig. 10(c)].159,160 The comparison of the abilities of three probes, including phenyl trifluoromethyldiazirine, benzophenone, and phenylazide, revealed that phenyl trifluoromethyldiazirine gave more accurate mass spectra results of amyloid nanostructures due to the higher carbene intermediate reactivity.

FIG. 10.

Photocrosslinking of peptide assemblies. (a) Intrafibrillar photopolymerization of β-hairpin peptide-based hydrogel. Reproduced with permission from Rughani et al., Macromolecules 43, 7924–7930 (2010). Copyright 2010 American Chemical Society. (b) Ruthenium-complex-catalyzed photocrosslinking of tyrosine-containing peptide hydrogels. Reproduced with permission from Ding et al., Langmuir 29, 13299–13306 (2013). Copyright 2013 American Chemical Society. (c) Photocrosslinking of amyloid-like nanostructures with diazirine derivative. Reproduced with permission from Preston et al., ACS Chem. Biol. 9, 761–768 (2014). Copyright 2014 American Chemical Society.

FIG. 10.

Photocrosslinking of peptide assemblies. (a) Intrafibrillar photopolymerization of β-hairpin peptide-based hydrogel. Reproduced with permission from Rughani et al., Macromolecules 43, 7924–7930 (2010). Copyright 2010 American Chemical Society. (b) Ruthenium-complex-catalyzed photocrosslinking of tyrosine-containing peptide hydrogels. Reproduced with permission from Ding et al., Langmuir 29, 13299–13306 (2013). Copyright 2013 American Chemical Society. (c) Photocrosslinking of amyloid-like nanostructures with diazirine derivative. Reproduced with permission from Preston et al., ACS Chem. Biol. 9, 761–768 (2014). Copyright 2014 American Chemical Society.

Close modal

Photocrosslinking is a useful approach to covalently conjugate peptides to target biomolecules, and the reactivity was widely explored.161,162 In particular, a peptide probe that captured proteins that recognize histone post-translational modification was developed by conjugating a benzophenone moiety to the N-terminal fragment of trimethylated histone H3 at lysine-4 [Fig. 11(a)].163 The in vitro analysis revealed the covalent binding of the probe to a targeted protein, plant homeodomain (PHD) finger of ING2 by UV light irradiation followed by conjugation with rhodamine azide. Endogenous ING2 and other histone-binding partners in cell lysates, in addition to the targeted protein, were captured using the probe. Additionally, the photocrosslinking method was applied to a site-specific conjugation of peptides/proteins to antibodies.164,165 In the approach by Park et al., a Fc-binding peptide with a benzophenone moiety was developed for the conjugation to targeted immunoglobulin G (IgG).165 Efficient photocrosslinking to Her2-targeting IgG (trastuzumab) and FcIII-fused engineered Pseudomonasexotoxin endotoxin A (PE24) was observed by optimizing the positions of benzophenone introduced to the Fc-binding peptide called FcIII [Fig. 11(b)]. The constructed trastuzumab-PE24 conjugate exhibited cytotoxicity to Her2-overexpressing cell lines, probably by protein synthesis inhibition. The photocrosslinking method would be useful to conjugate various cargos into antibodies for therapeutic and diagnostic applications.

FIG. 11.

Affinity-based photocrosslinking of peptides to biomolecules. (a) Photo-affinity capture of proteins that recognize trimethylated lysine-4 of histone H3. Reproduced with permission from Li and Kapoor, J. Am. Chem. Soc. 132, 2504–2505 (2010). Copyright 2010 American Chemical Society. (b) Site-specific IgG conjugation by peptide-directed photocrosslinking. Reproduced with permission from Park et al., Bioconjugate Chem. 29, 3240–3244 (2018). Copyright 2018 American Chemical Society. (c) Light-induced microtubule stabilization by the incorporation of diazirine-conjugated Tau-derived peptide (DA-TP) and subsequent UV light irradiation. CLSM images of GTP microtubules. Reproduced with permission from Watari et al., Chem. Commun. 58, 9190−9193 (2022). Copyright 2022 The Royal Society of Chemistry.

FIG. 11.

Affinity-based photocrosslinking of peptides to biomolecules. (a) Photo-affinity capture of proteins that recognize trimethylated lysine-4 of histone H3. Reproduced with permission from Li and Kapoor, J. Am. Chem. Soc. 132, 2504–2505 (2010). Copyright 2010 American Chemical Society. (b) Site-specific IgG conjugation by peptide-directed photocrosslinking. Reproduced with permission from Park et al., Bioconjugate Chem. 29, 3240–3244 (2018). Copyright 2018 American Chemical Society. (c) Light-induced microtubule stabilization by the incorporation of diazirine-conjugated Tau-derived peptide (DA-TP) and subsequent UV light irradiation. CLSM images of GTP microtubules. Reproduced with permission from Watari et al., Chem. Commun. 58, 9190−9193 (2022). Copyright 2022 The Royal Society of Chemistry.

Close modal

Photoisomerization is a useful approach for modulating microtubule structures, as summarized in Sec. II D. One of the issues of photoisomerization is the reversible property that may induce moderate structural change of microtubules. The photocrosslinking technology, as another approach, was used for the light-induced microtubule stabilization [Fig. 11(c)].166 Diazirine was used as a photocrosslinking agent that formed a carbene upon UV light (365 nm) irradiation to form a covalent bond between TP and microtubules. TP containing diazirine at the N-terminus (DA-TP) was synthesized and TMR was subsequently conjugated (DA-TP-TMR). The binding of DA-TP-TMR to microtubules and covalent bonding formation upon UV light irradiation was confirmed by confocal laser scanning microscopy (CLSM) and SDS-PAGE. DA-TP itself stabilized GTP-microtubule structures, such as taxol. Interestingly, UV light irradiation to DA-TP-encapsulated microtubules induced longer and more rigid microtubule formation [Fig. 11(c)]. The results indicate that forming the covalent bond between DA-TP and microtubules by UV light irradiation generated long, rigid, and stable microtubules. Additionally, the binding of DA-TP-TMR to intracellular microtubules of HepG2 cells was confirmed by CLSM. UV light irradiation to the DA-TP-TMR-incorporated cells and subsequent incubation caused abnormal cell shapes, nuclear defects, and cell death. Thus, the binding of DA-TP-TMR to intracellular microtubules and the UV light-induced microtubule stabilization caused cell death. The peptide-based microtubule stabilization would be a new approach for manipulating cell structures and functions.

Photodimerization reactions can also be useful in controlling the biological function and self-assembly behavior of peptides although the examples are relatively less.167–171 Ding et al. developed pH-responding polypeptide nanogels that are stabilized by photodimerization between the cinnamyloxy groups [Fig. 12(a)].168 Diblock copolymer that comprise of poly(ethylene glycol monomethyl ether) (PEG) and poly(L-glutamic acid-co-γ-cinnamyl-L-glutamate) self-assembled into polymer micelles with PEG shell and γ-cinnamyl-L-glutamate core. The photodimerization through [2 + 2] cycloaddition between the cinnamyloxy groups afforded the cross-linked nanogels of approximately 30 nm in size. The release of the encapsulated antibiotic rifampicin in the cross-linked nanogel was controlled by nanogel swelling in response to pH. Bullen et al. showed that photodimerization of transcription factor GCN4 through [4 + 4] cycloaddition between the anthracenes induced sequence-specific and strong DNA-binding [Fig. 12(b)].169 The anthracene photodimerization only occurs after the conjugates are preorganized on the correct DNA sequence, thus it made the process very specific and controllable. Joseph et al. developed an Aβ42 amyloid nucleating core peptide modified with photodimerizable 4-methylcoumarin [Fig. 12(c)].170 The peptide conjugate self-assembled into 1D nanofibers via nanoparticle. In contrast, the photodimerization of two 4-methylcoumarin moieties that are included in γ-cyclodextrin resulted in morphological conversion to free-standing 2D nanosheets. The structural and temporal control to differentiate the pathway was used to tune the mechanical strength of hierarchical hydrogel materials.

FIG. 12.

Photodimerizable peptide materials. (a) Photodimerized pH-responsive polypeptide nanogel. Reproduced with permission from Ding et al., J. Mater. Chem. 21, 11383–11391 (2011). Copyright 2011 The Royal Society of Chemistry. (b) Light-induced sequence-selective DNA binding of peptide by anthracene photodimerization. Reproduced with permission from Bullen et al., Chem. Commun. 51, 8130–8133 (2015). Copyright 2015 The Royal Society of Chemistry. (c) 2D peptide nanosheet formed by 4-methylcoumarin photodimerization in γ-cyclodextrin. Reproduced with permission from Joseph et al., ACS Appl. Mater. Interfaces 11, 28213–28220 (2019). Copyright 2019 American Chemical Society.

FIG. 12.

Photodimerizable peptide materials. (a) Photodimerized pH-responsive polypeptide nanogel. Reproduced with permission from Ding et al., J. Mater. Chem. 21, 11383–11391 (2011). Copyright 2011 The Royal Society of Chemistry. (b) Light-induced sequence-selective DNA binding of peptide by anthracene photodimerization. Reproduced with permission from Bullen et al., Chem. Commun. 51, 8130–8133 (2015). Copyright 2015 The Royal Society of Chemistry. (c) 2D peptide nanosheet formed by 4-methylcoumarin photodimerization in γ-cyclodextrin. Reproduced with permission from Joseph et al., ACS Appl. Mater. Interfaces 11, 28213–28220 (2019). Copyright 2019 American Chemical Society.

Close modal

This review outlined that the photocontrol of self-assembled materials and interactions with biomolecules by photoresponsive peptides has made significant progress over the past 30 years. Photoresponsive peptides were demonstrated to control the formation and dissociation of peptide assemblies, gene expression, protein-drug interactions, protein–protein interactions, deformation and motility of liposome, structure and stability of cytoskeletons, and cell functions. Many of the building units for incorporating photochromic, photocleavable, and photocrosslinking molecular units into peptides are available or can be readily synthesized. Photocontrol of biological systems through “optogenetics” using photoactive proteins has recently made remarkable progress,172 but photoresponsive peptides based on synthetic chemistry have the advantage of diversity in design and ease of preparation.

Developing photoresponsive peptides that respond faster and have higher spatiotemporal resolution is a future challenge. For the purpose, the optimization of structures of photoresponsive peptides is important. In addition to high-throughput screening methods, such as phage display, recent development of artificial intelligence-based technologies will help the generation of new photoresponsive peptides with superior properties such as high ON/OFF photoresponse, targetability, and self-assembly. The low stability of peptides due to proteolytic degradation may be problematic for the biological applications. Constraints of peptide conformations by cyclization and stapling and use of D-amino acids, β-peptides and peptoids are useful approaches to overcome the problem. They will help develop technologies for photo-controlling various biological systems, such as signal transduction, neuronal activation, molecular transport through channel proteins, cell division and differentiation, and internalization into cells. The use of photosensitive peptides to perform phototherapy may become possible for diseases, such as cancer, protein misfolding, and neurological. Photoresponsive peptides may also enable other in vivo applications such as tissue engineering, diagnostics, and immunomodulation. The challenge in using photoresponsive peptides in vivo is to develop molecules that respond to near-infrared light with high biopermeability. From the perspective of materials chemistry, creating smart molecular machines, molecular robots, and artificial cells is expected by incorporating photoresponsive peptides into multi-component molecular integration devices.173 For material applications of photoresponsive peptides, it is important to improve the stability and durability, understand structure-function relationships, and incorporate into other materials. Such integrated artificial systems that contain photoresponsive peptides will also be useful as next-generation smart DDS materials and can also help develop medicine.

The authors have no conflicts to disclose.

Kazunori Matsuura: Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Hiroshi Inaba: Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
A. D.
Cunningham
,
N.
Qvit
, and
D.
Mochly-Rosen
, “
Peptides and peptidomimetics as regulators of protein–protein interactions
,”
Curr. Opin. Struct. Biol.
44
,
59
66
(
2017
).
2.
H.
Wang
,
R. S.
Dawber
,
P.
Zhang
,
M.
Walko
,
A. J.
Wilson
, and
X.
Wang
, “
Peptide-based inhibitors of protein–protein interactions: Biophysical, structural and cellular consequences of introducing a constraint
,”
Chem. Sci.
12
,
5977
5993
(
2021
).
3.
X.
Wang
,
D.
Ni
,
Y.
Liu
, and
S.
Lu
, “
Rational design of peptide-based inhibitors disrupting protein–protein interactions
,”
Front. Chem.
9
,
682675
(
2021
).
4.
M. E.
Vázquez
,
A. M.
Caamaño
, and
J. L.
Mascareñas
, “
From transcription factors to designed sequence-specific DNA-binding peptides
,”
Chem. Soc. Rev.
32
,
338
349
(
2003
).
5.
E.
Pazos
,
J.
Mosquera
,
M. E.
Vázquez Sentís
, and
J. L.
Mascareñas
, “
DNA recognition by synthetic constructs
,”
ChemBioChem
12
,
1958
1973
(
2011
).
6.
V. K.
Khavinson
,
I.
Grigor'evna Popovich
,
N. S.
Linkova
,
E. S.
Mironova
, and
A. R.
Ilina
, “
Peptide regulation of gene expression: A systematic review
,”
Molecules
26
,
7053
(
2021
).
7.
H.
Inaba
and
K.
Matsuura
, “
Modulation of microtubule properties and functions by encapsulation of nanomaterials using a Tau-derived peptide
,”
Bull. Chem. Soc. Jpn.
94
,
2100
2112
(
2021
).
8.
I. W.
Hamley
, “
The amyloid beta peptide: A chemist's perspective. Role in Alzheimer's and fibrillization
,”
Chem. Rev.
112
,
5147
5192
(
2012
).
9.
I. W.
Hamley
, “
Small bioactive peptides for biomaterials design and therapeutics
,”
Chem. Rev.
117
,
14015
14041
(
2017
).
10.
G.
Qi
,
Y.
Gao
,
L.
Wang
, and
H.
Wang
, “
Self-assembled peptide-based nanomaterials for biomedical imaging and therapy
,”
Adv. Mater.
30
,
1703444
(
2018
).
11.
V.
Devika
,
P. J.
Sreelekshmi
,
N.
Rajeev
,
S. A.
Lakshmi
,
A.
Chandran
,
G. B.
Gouthami
, and
S.
Sadanandan
, “
Recent advances in peptides-based stimuli-responsive materials for biomedical and therapeutic applications: A review
,”
Mol. Pharm.
19
,
1999
2021
(
2022
).
12.
A.
Henninot
,
J. C.
Collins
, and
J. M.
Nuss
, “
The current state of peptide drug discovery: Back to the future?
,”
J. Med. Chem.
61
,
1382
1414
(
2018
).
13.
M.
Muttenthaler
,
G. F.
King
,
D. J.
Adams
, and
P. F.
Alewood
, “
Trends in peptide drug discovery
,”
Nat. Rev. Drug Discovery
20
,
309
325
(
2021
).
14.
L.
Wang
,
N.
Wang
,
W.
Zhang
,
X.
Cheng
,
Z.
Yan
,
G.
Shao
,
X.
Wang
,
R.
Wang
, and
C.
Fu
, “
Therapeutic peptides: Current applications and future directions
,”
Signal Transduction Targeted Ther.
7
,
48
(
2022
).
15.
S.
Lee
,
J.
Xie
, and
X.
Chen
, “
Peptide-based probes for targeted molecular imaging
,”
Biochemistry
49
,
1364
1376
(
2010
).
16.
S.
Lee
,
J.
Xie
, and
X.
Chen
, “
Peptides and peptide hormones for molecular imaging and disease diagnosis
,”
Chem. Rev.
110
,
3087
3111
(
2010
).
17.
P.
Zhang
,
Y.
Cui
,
C. F.
Anderson
,
C.
Zhang
,
Y.
Li
,
R.
Wang
, and
H.
Cui
, “
Peptide-based nanoprobes for molecular imaging and disease diagnostics
,”
Chem. Soc. Rev.
47
,
3490
3529
(
2018
).
18.
R. V.
Ulijn
and
A. M.
Smith
, “
Designing peptide based nanomaterials
,”
Chem. Soc. Rev.
37
,
664
675
(
2008
).
19.
R.
de la Rica
and
H.
Matsui
, “
Applications of peptide and protein-based materials in bionanotechnology
,”
Chem. Soc. Rev.
39
,
3499
3509
(
2010
).
20.
H.
Inaba
and
K.
Matsuura
, “
Peptide nanomaterials designed from natural supramolecular systems
,”
Chem. Rec.
19
,
843
858
(
2019
).
21.
D. M.
Copolovici
,
K.
Langel
,
E.
Eriste
, and
Ü.
Langel
, “
Cell-penetrating peptides: Design, synthesis, and applications
,”
ACS Nano
8
,
1972
1994
(
2014
).
22.
K.
Desale
,
K.
Kuche
, and
S.
Jain
, “
Cell-penetrating peptides (CPPs): An overview of applications for improving the potential of nanotherapeutics
,”
Biomater. Sci.
9
,
1153
1188
(
2021
).
23.
B.
Ramakers
,
J.
Hest
, and
D.
Löwik
, “
Molecular tools for the construction of peptide-based materials
,”
Chem. Soc. Rev.
43
,
2743
2756
(
2014
).
24.
K.
Matsuurua
, “
Rational design of self-assembled proteins and peptides for nano-and micro-sized architectures
,”
RSC Adv.
4
,
2942
2953
(
2014
).
25.
E. D.
Santis
and
M. G.
Ryadnov
, “
Peptide self-assembly for nanomaterials: The old new kid on the block
,”
Chem. Soc. Rev.
44
,
8288
8300
(
2015
).
26.
S.
Lou
,
X.
Wang
,
Z.
Yu
, and
L.
Shi
, “
Peptide tectonics: Encoded structural complementarity dictates programmable self-assembly
,”
Adv. Sci.
6
,
1802043
(
2019
).
27.
A.
Levin
,
T. A.
Hakala
,
L.
Schnaider
,
G. J. L.
Bernardes
,
E.
Gazit
, and
T. P. J.
Knowles
, “
Biomimetic peptide self-assembly for functional materials
,”
Nat. Rev. Chem.
4
,
615
634
(
2020
).
28.
N. J.
Sinha
,
M. G.
Langenstein
,
D. J.
Pochan
,
C. J.
Kloxin
, and
J. G.
Saven
, “
Peptide design and self-assembly into targeted nanostructure and functional materials
,”
Chem. Rev.
121
,
13915
13935
(
2021
).
29.
Y.
Ruan
,
M.
Sohail
,
J.
Zhao
,
F.
Hu
,
Y.
Li
,
P.
Wang
, and
L.
Zhang
, “
Applications of material-binding peptides: A review
,”
ACS Biomater. Sci. Eng.
8
,
4738
4750
(
2022
).
30.
Y.
Huo
,
J.
Hu
,
Y.
Yin
,
P.
Liu
,
K.
Cai
, and
W.
Ji
, “
Self-assembling peptide-based functional biomaterials
,”
ChemBioChem
24
,
e202200582
(
2023
).
31.
J. M.
Palomo
, “
Solid-phase peptide synthesis: An overview focused on the preparation of biologically relevant peptides
,”
RSC Adv.
4
,
32658
32672
(
2014
).
32.
D. M. M.
Jaradat
, “
Thirteen decades of peptide synthesis: Key developents in solid phase peptide synthesis and amide bond formation utilized in peptide ligation
,”
Amino Acids
50
,
39
68
(
2018
).
33.
W.
Jaroszewicz
,
J.
Morcinek-Orłowska
,
K.
Pierzynowska
,
L.
Gaffke
, and
G.
Węgrzyn
, “
Phage display and other peptide display technologies
,”
FEMS Microbiol. Rev.
46
,
fuab052
(
2022
).
34.
G. A.
Woolley
, “
Photocontrolling peptide α helices
,”
Acc. Chem. Res.
38
,
486
493
(
2005
).
35.
G.
Mayer
and
A.
Heckel
, “
Biologically active molecules with a ‘light switch
,’”
Angew. Chem. Int. Ed.
45
,
4900
4921
(
2006
).
36.
W.
Szymański
,
J. M.
Beierle
,
H. A. V.
Kistemaker
,
W. A.
Velema
, and
B. L.
Feringa
, “
Reversible photocontrol of biological systems by the incorporation of molecular photoswitches
,”
Chem. Rev.
113
,
6114
6178
(
2013
).
37.
Z.
Yu
and
S.
Hecht
, “
Remote control over folding by light
,”
Chem. Commun.
52
,
6639
6653
(
2016
).
38.
D.
Dattler
,
G.
Fuks
,
J.
Heiser
,
E.
Moulin
,
A.
Perrot
,
X.
Yao
, and
N.
Giuseppone
, “
Design of collective motions from synthetic molecular switches, rotors, and motors
,”
Chem. Rev.
120
,
310
433
(
2020
).
39.
A.
Gautier
,
C.
Gauron
,
M.
Volovitch
,
D.
Bensimon
,
L.
Jullien
, and
S.
Vriz
, “
How to control proteins with light in living systems
,”
Nat. Chem. Biol.
10
,
533
541
(
2014
).
40.
K.
Hüll
,
J.
Morstein
, and
D.
Trauner
, “
In vivo photopharmacology
,”
Chem. Rev.
118
,
10710
10747
(
2018
).
41.
C.
Brieke
,
F.
Rohrbach
,
A.
Gottschalk
,
G.
Mayer
, and
A.
Heckel
, “
Light-controlled tools
,”
Angew. Chem. Int. Ed.
51
,
8446
8476
(
2012
).
42.
I. M.
Welleman
,
M. W. H.
Hoorens
,
B. L.
Feringa
,
H. H.
Boersma
, and
W.
Szymański
, “
Photoresponsive molecular tools for emerging applications of light in medicine
,”
Chem. Sci.
11
,
11672
11691
(
2020
).
43.
A. A.
Beharrya
and
G. A.
Woolley
, “
Azobenzene photoswitches for biomolecules
,”
Chem. Soc. Rev.
40
,
4422
4437
(
2011
).
44.
R. J.
Mart
and
R. K.
Allemann
, “
Azobenzene photocontrol of peptides and proteins
,”
Chem. Commun.
52
,
12262
12277
(
2016
).
45.
V.
Peddie
and
A. D.
Abell
, “
Photocontrol of peptide secondary structure through non-azobenzene photoswitches
,”
J. Photochem. Photobiol., C
40
,
1
20
(
2019
).
46.
R.
Klajn
, “
Spiropyran-based dynamic materials
,”
Chem. Soc. Rev.
43
,
148
184
(
2014
).
47.
J.
Liu
,
W.
Kang
, and
W.
Wang
, “
Photocleavage-based photoresponsive drug delivery
,”
Photochem. Photobiol.
98
,
288
302
(
2022
).
48.
N. D.
Pham
,
R. B.
Parker
, and
J. J.
Kohler
, “
Photocrosslinking approaches to interactome mapping
,”
Curr. Opin. Chem. Biol.
17
,
90
101
(
2013
).
49.
A.
Aemissegger
,
V.
Kräutler
,
W. F. V.
Gunsteren
, and
D.
Hilvert
, “
A photoinducible β-hairpin
,”
J. Am. Chem. Soc.
127
,
2929
2936
(
2005
).
50.
A.
Aemissegger
and
D.
Hilvert
, “
Synthesis and application of an azobenzene amino acid as a light-switchable turn element in polypeptides
,”
Nat. Protoc.
2
,
161
167
(
2007
).
51.
S. A.
Waldauer
,
S.
Hassan
,
B.
Paoli
,
P. M.
Donaldson
,
R.
Pfister
,
P.
Hamm
,
A.
Caflisch
, and
R.
Pellarin
, “
Photocontrol of reversible amyloid formation with a minimal-design peptide
,”
J. Phys. Chem. B
116
,
8961
8973
(
2012
).
52.
T. M.
Doran
,
E. A.
Anderson
,
S. E.
Latchney
,
L. A.
Opanashuk
, and
B. L.
Nilsson
, “
An azobenzene photoswitch sheds light on turn nucleation in amyloid-β self-assembly
,”
ACS Chem. Neurosci.
3
,
211
220
(
2012
).
53.
X.
Li
,
Y.
Gao
,
Y.
Kuang
, and
B.
Xu
, “
Enzymatic formation of a photoresponsive supramolecular hydrogel
,”
Chem. Commun.
46
,
5364
5366
(
2010
).
54.
W.
Li
,
I.
Park
,
S.-K.
Kang
, and
M.
Lee
, “
Smart hydrogels from laterally-grafted peptide assembly
,”
Chem. Commun.
48
,
8796
8798
(
2012
).
55.
Y.
Huang
,
Z.
Qiu
,
Y.
Xu
,
J.
Shi
,
H.
Lin
, and
Y.
Zhang
, “
Supramolecular hydrogels based on short peptides linked with conformational switch
,”
Org. Biomol. Chem.
9
,
2149
2155
(
2011
).
56.
T. M.
Doran
,
D. M.
Ryan
, and
B. L.
Nilsson
, “
Reversible photocontrol of self-assembled peptide hydrogel viscoelasticity
,”
Polym. Chem.
5
,
241
248
(
2014
).
57.
K.
Matsuura
, “
Synthetic approaches to construct viral capsid-like spherical nanomaterials
,”
Chem. Commun.
54
,
8944
8959
(
2018
).
58.
K.
Matsuura
and
S.
Fujita
, “
A photoresponsive artificial viral capsid self-assembled from an azobenzene-containing β-annulus peptide
,”
Int. J. Mol. Sci.
22
,
4028
(
2021
).
59.
T.
Sendai
,
S.
Biswas
, and
T.
Aida
, “
Photoreconfigurable supramolecular nanotube
,”
J. Am. Chem. Soc.
135
,
11509
11512
(
2013
).
60.
M.
Liu
,
C.
Creemer
,
T.
Reardona
, and
J.
Parquette
, “
Light-driven dissipative self-assembly of a peptide hydrogel
,”
Chem. Commun.
57
,
13776
13779
(
2021
).
61.
Y.
Liang
,
S.
Ogawa
,
H.
Inaba
, and
K.
Matsuura
, “
Dramatic morphological changes in liposomes induced by peptide nanofibers reversibly polymerized and depolymerized by the photoisomerization of spiropyran
,”
Front. Mol. Biosci.
10
,
1137885
(
2023
).
62.
C.
Poloni
,
M. C. A.
Stuart
,
P.
van der Meulen
,
W.
Szymanski
, and
B. L.
Feringa
, “
Light and heat control over secondary structure and amyloid-like fiber formation in an overcrowded-alkene-modified Trp zipper
,”
Chem. Sci.
6
,
7311
7318
(
2015
).
63.
G.
Marafon
,
M.
Crisma
, and
A.
Moretto
, “
Intrinsically photoswitchable α/β peptides toward two-state foldamers
,”
Angew. Chem. Int. Ed.
57
,
10217
10220
(
2018
).
64.
K.
Nakamura
,
W.
Tanaka
,
K.
Sada
,
R.
Kubota
,
T.
Aoyama
,
K.
Urayama
, and
I.
Hamachi
, “
Phototriggered spatially controlled out-of-equilibrium patterns of peptide nanofibers in a self-sorting double network hydrogel
,”
J. Am. Chem. Soc.
143
,
19532
19541
(
2021
).
65.
S.
Sun
,
H.-W.
Liang
,
H.
Wang
, and
Q.
Zou
, “
Light-triggered self-assembly of peptide nanoparticles into nanofibers in living cells through molecular conformation changes and H-bond interactions
,”
ACS Nano
16
,
18978
18989
(
2022
).
66.
W.
Ji
,
C.
Yuan
,
P.
Chakraborty
,
P.
Makam
,
S.
Bera
,
S.
Rencus-Lazar
,
J.
Li
,
X.
Yan
, and
E.
Gazit
, “
Coassembly-induced transformation of dipeptide amyloid-like structures into stimuli-responsive supramolecular materials
,”
ACS Nano
14
,
7181
7190
(
2020
).
67.
W.
Ji
,
Y.
Tang
,
P.
Makam
,
Y.
Yao
,
R.
Jiao
,
K.
Cai
,
G.
Wei
, and
E.
Gazit
, “
Expanding the structural diversity and functional scope of diphenylalanine-based peptide architectures by hierarchical coassembly
,”
J. Am. Chem. Soc.
143
,
17633
17645
(
2021
).
68.
W.
Keller
,
P.
König
, and
T. J.
Richmond
, “
Crystal structure of a bZIP/DNA complex at 2.2 A: Determinants of DNA specific recognition
,”
J. Mol. Biol.
254
,
657
667
(
1995
).
69.
J.
Rodriguez
,
J.
Mosquera
,
S.
Learte-Aymamí
,
M. E.
Vázquez
, and
J. L.
Mascareñas
, “
Stimuli-responsive DNA binding by synthetic systems
,”
Acc. Chem. Res.
53
,
2286
2298
(
2020
).
70.
B.
Cuenoud
and
A.
Schepartz
, “
Altered specificity of DNA-binding proteins with transition metal dimerization domains
,”
Science
259
,
510
513
(
1993
).
71.
M.
Ueno
,
A.
Murakami
,
K.
Makino
, and
T.
Morii
, “
Arranging quaternary structure of peptides by cyclodextrin-guest inclusion complex: Sequence-specific DNA binding by a peptide dimer with artificial dimerization module
,”
J. Am. Chem. Soc.
115
,
12575
12576
(
1993
).
72.
J.
Mosquera
,
A.
Jiménez-Balsa
,
V. I.
Dodero
,
M. E.
Vázquez
, and
J. L.
Mascareñas
, “
Stimuli-responsive selection of target DNA sequences by synthetic bZIP peptides
,”
Nat. Commun.
4
,
1874
(
2013
).
73.
P.
Novo
,
M. D.
García
,
C.
Peinador
, and
E.
Pazos
, “
Reversible control of DNA binding with cucurbit[8]uril-induced supramolecular 4,4′-bipyridinium–peptide dimers
,”
Bioconjugate Chem.
32
,
507
511
(
2021
).
74.
A. M.
Caamaño
,
M. E.
Vázquez
,
J.
Martínez-Costas
,
L.
Castedo
, and
J. L.
Mascareñas
, “
A light-modulated sequence-specific DNA-binding peptide
,”
Angew. Chem. Int. Ed.
39
,
3104
3107
(
2000
).
75.
G. A.
Woolley
,
A. S. I.
Jaikaran
,
M.
Berezovski
,
J. P.
Calarco
,
S. N.
Krylov
,
O. S.
Smart
, and
J. R.
Kumita
, “
Reversible photocontrol of DNA binding by a designed GCN4-bZIP protein
,”
Biochemistry
45
,
6075
6084
(
2006
).
76.
L.
Guerrero
,
O. S.
Smart
,
G. A.
Woolley
, and
R. K.
Allemann
, “
Photocontrol of DNA binding specificity of a miniature engrailed homeodomain
,”
J. Am. Chem. Soc.
127
,
15624
15629
(
2005
).
77.
F.
Zhang
,
K. A.
Timm
,
K. M.
Arndt
, and
G. A.
Woolley
, “
Photocontrol of coiled-coil proteins in living cells
,”
Angew. Chem. Int. Ed.
49
,
3943
3946
(
2010
).
78.
G. M.
Murawska
,
C.
Poloni
,
N. A.
Simeth
,
W.
Szymanski
, and
B. L.
Feringa
, “
Comparative study of photoswitchable zinc-finger domain and AT-hook motif for light-controlled peptide–DNA binding
,”
Chem. Eur. J.
25
,
4965
4973
(
2019
).
79.
K.
Fujimoto
,
M.
Kajino
,
I.
Sakaguchi
, and
M.
Inouye
, “
Photoswitchable, DNA-binding helical peptides assembled with two independently designed sequences for photoregulation and DNA recognition
,”
Chem. Eur. J.
18
,
9834
9840
(
2012
).
80.
D.
Liu
,
J.
Karanicolas
,
C.
Yu
,
Z.
Zhang
, and
G. A.
Woolley
, “
Site-specific incorporation of photoisomerizable azobenzene groups into ribonuclease S
,”
Bioorg. Med. Chem. Lett.
7
,
2677
2680
(
1997
).
81.
I.
Hamachi
,
T.
Hiraoka
,
Y.
Yamada
, and
S.
Shinkai
, “
Photoswitching of the enzymatic activity of semisynthetic ribonuclease S′ bearing phenylazophenylalanine at a specific site
,”
Chem. Lett.
27
,
537
538
(
1998
).
82.
M.
Volgraf
,
P.
Gorostiza
,
R.
Numano
,
R. H.
Kramer
,
E. Y.
Isacoff
, and
D.
Trauner
, “
Allosteric control of an ionotropic glutamate receptor with an optical switch
,”
Nat. Chem. Biol.
2
,
47
52
(
2006
).
83.
M.
Volgraf
,
P.
Gorostiza
,
S.
Szobota
,
M. R.
Helix
,
E. Y.
Isacoff
, and
D.
Trauner
, “
Reversibly caged glutamate:  A photochromic agonist of ionotropic glutamate receptors
,”
J. Am. Chem. Soc.
129
,
260
261
(
2007
).
84.
M.
Banghart
,
K.
Borges
,
E.
Isacoff
,
D.
Trauner
, and
R. H.
Kramer
, “
Light-activated ion channels for remote control of neuronal firing
,”
Nat. Neurosci.
7
,
1381
1386
(
2004
).
85.
N. N.
Mafy
,
K.
Matsuo
,
S.
Hiruma
,
R.
Uehara
, and
N.
Tamaoki
, “
Photoswitchable CENP-E inhibitor enabling the dynamic control of chromosome movement and mitotic progression
,”
J. Am. Chem. Soc.
142
,
1763
1767
(
2020
).
86.
K.
Matsuo
and
N.
Tamaoki
, “
Rational design and development of a lit-active photoswitchable inhibitor targeting CENP-E
,”
Org. Biomol. Chem.
19
,
6979
6984
(
2021
).
87.
D.
Wildemann
,
C.
Schiene-Fischer
,
T.
Aumüller
,
A.
Bachmann
,
T.
Kiefhaber
,
C.
Lücke
, and
G.
Fischer
, “
A nearly isosteric photosensitive amide-backbone substitution allows enzyme activity switching in ribonuclease S
,”
J. Am. Chem. Soc.
129
,
4910
4918
(
2007
).
88.
J.
Broichhagen
,
T.
Podewin
,
H.
Meyer-Berg
,
Y.
von Ohlen
,
N. R.
Johnston
,
B. J.
Jones
,
S. R.
Bloom
,
G. A.
Rutter
,
A.
Hoffmann-Röder
,
D. J.
Hodson
, and
D.
Trauner
, “
Optical control of insulin secretion using an incretin switch
,”
Angew. Chem. Int. Ed.
54
,
15565
15569
(
2015
).
89.
L.
Albert
,
J.
Xu
,
R.
Wan
,
V.
Srinivasan
,
Y.
Dou
, and
O.
Vázquez
, “
Controlled inhibition of methyltransferases using photoswitchable peptidomimetics: Towards an epigenetic regulation of leukemia
,”
Chem. Sci.
8
,
4612
4618
(
2017
).
90.
D.
Liu
,
Y.
Xie
,
H.
Shao
, and
X.
Jiang
, “
Using azobenzene-embedded self-assembled monolayers to photochemically control cell adhesion reversibly
,”
Angew. Chem. Int. Ed.
48
,
4406
4408
(
2009
).
91.
S.
Kneissl
,
E. J.
Loveridge
,
C.
Williams
,
M. P.
Crump
, and
R. K.
Allemann
, “
Photocontrollable peptide-based switches target the anti-apoptotic protein Bcl-xL
,”
ChemBioChem
9
,
3046
3054
(
2008
).
92.
L.
Nevola
,
A.
Martín-Quirós
,
K.
Eckelt
,
N.
Camarero
,
S.
Tosi
,
A.
Llobet
,
E.
Giralt
, and
P.
Gorostiza
, “
Light-regulated stapled peptides to inhibit protein–protein interactions involved in clathrin-mediated endocytosis
,”
Angew. Chem. Int. Ed.
52
,
7704
7708
(
2013
).
93.
B.
Jankovic
,
A.
Gulzar
,
C.
Zanobini
,
O.
Bozovic
,
S.
Wolf
,
G.
Stock
, and
P.
Hamm
, “
Photocontrolling protein–peptide interactions: From minimal perturbation to complete unbinding
,”
J. Am. Chem. Soc.
141
,
10702
10710
(
2019
).
94.
L.
Albert
,
J.
Nagpal
,
W.
Steinchen
,
L.
Zhang
,
L.
Werel
,
N.
Djokovic
,
D.
Ruzic
,
M.
Hoffarth
,
J.
Xu
,
J.
Kaspareit
,
F.
Abendroth
,
A.
Royant
,
G.
Bange
,
K.
Nikolic
,
S.
Ryu
,
Y.
Dou
,
L.-O.
Essen
, and
O.
Vázquez
, “
Bistable photoswitch allows in vivo control of hematopoiesis
,”
ACS Cent. Sci.
8
,
57
66
(
2022
).
95.
B.
Schierling
,
A.-J.
Noël
,
W.
Wende
,
L. T.
Hien
,
E.
Volkov
,
E.
Kubareva
,
T.
Oretskaya
,
M.
Kokkinidis
,
A.
Römpp
,
B.
Spengler
, and
A.
Pingoud
, “
Controlling the enzymatic activity of a restriction enzyme by light
,”
Proc. Natl. Acad. Sci. U. S. A.
107
,
1361
1366
(
2010
).
96.
M.
Borowiak
,
W.
Nahaboo
,
M.
Reynders
,
K.
Nekolla
,
P.
Jalinot
,
J.
Hasserodt
,
M.
Rehberg
,
M.
Delattre
,
S.
Zahler
,
A.
Vollmar
,
D.
Trauner
, and
O.
Thorn-Seshold
, “
Photoswitchable inhibitors of microtubule dynamics optically control mitosis and cell death
,”
Cell
162
,
403
411
(
2015
).
97.
J.
Zenker
,
M. D.
White
,
R. M.
Templin
,
R. G.
Parton
,
O.
Thorn-Seshold
,
S.
Bissiere
, and
N.
Plachta
, “
A microtubule-organizing center directing intracellular transport in the early mouse embryo
,”
Science
357
,
925
928
(
2017
).
98.
A. J.
Engdahl
,
E. A.
Torres
,
S. E.
Lock
,
T. B.
Engdahl
,
P. S.
Mertz
, and
C. N.
Streu
, “
Synthesis, characterization, and bioactivity of the photoisomerizable tubulin polymerization inhibitor azo-combretastatin A4
,”
Org. Lett.
17
,
4546
4549
(
2015
).
99.
J. E.
Sheldon
,
M. M.
Dcona
,
C. E.
Lyons
,
J. C.
Hackett
, and
M. C. T.
Hartman
, “
Photoswitchable anticancer activity via trans–cis isomerization of a combretastatin A-4 analog
,”
Org. Biomol. Chem.
14
,
40
49
(
2015
).
100.
S. K.
Rastogi
,
Z.
Zhao
,
S. L.
Barrett
,
S. D.
Shelton
,
M.
Zafferani
,
H. E.
Anderson
,
M. O.
Blumenthal
,
L. R.
Jones
,
L.
Wang
,
X.
Li
,
C. N.
Streu
,
L.
Du
, and
W. J.
Brittain
, “
Photoresponsive azo-combretastatin A-4 analogues
,”
Eur. J. Med. Chem.
143
,
1
7
(
2018
).
101.
A.
Müller-Deku
,
J. C. M.
Meiring
,
K.
Loy
,
Y.
Kraus
,
C.
Heise
,
R.
Bingham
,
K. I.
Jansen
,
X.
Qu
,
F.
Bartolini
,
L. C.
Kapitein
,
A.
Akhmanova
,
J.
Ahlfeld
,
D.
Trauner
, and
O.
Thorn-Seshold
, “
Photoswitchable paclitaxel-based microtubule stabilisers allow optical control over the microtubule cytoskeleton
,”
Nat. Commun.
11
,
4640
(
2020
).
102.
L.
Gao
,
J. C. M.
Meiring
,
Y.
Kraus
,
M.
Wranik
,
T.
Weinert
,
S. D.
Pritzl
,
R.
Bingham
,
E.
Ntouliou
,
K. I.
Jansen
,
N.
Olieric
,
J.
Standfuss
,
L. C.
Kapitein
,
T.
Lohmüller
,
J.
Ahlfeld
,
A.
Akhmanova
,
M. O.
Steinmetz
, and
O.
Thorn-Seshold
, “
A robust, GFP-orthogonal photoswitchable inhibitor scaffold extends optical control over the microtubule cytoskeleton
,”
Cell Chem. Biol.
28
,
228
241.e6
(
2021
).
103.
A.
Sailer
,
J. C. M.
Meiring
,
C.
Heise
,
L. N.
Pettersson
,
A.
Akhmanova
,
J.
Thorn-Seshold
, and
O.
Thorn-Seshold
, “
Pyrrole hemithioindigo antimitotics with near-quantitative bidirectional photoswitching that photocontrol cellular microtubule dynamics with single-cell precision
,”
Angew. Chem. Int. Ed.
60
,
23695
23704
(
2021
).
104.
L.
Gao
,
J. C. M.
Meiring
,
C.
Heise
,
A.
Rai
,
A.
Müller-Deku
,
A.
Akhmanova
,
J.
Thorn-Seshold
, and
O.
Thorn-Seshold
, “
Photoswitchable epothilone-based microtubule stabilisers allow GFP-imaging-compatible, optical control over the microtubule cytoskeleton
,”
Angew. Chem. Int. Ed.
61
,
e202114614
(
2022
).
105.
L.
Gao
,
J. C. M.
Meiring
,
A.
Varady
,
I. E.
Ruider
,
C.
Heise
,
M.
Wranik
,
C. D.
Velasco
,
J. A.
Taylor
,
B.
Terni
,
T.
Weinert
,
J.
Standfuss
,
C. C.
Cabernard
,
A.
Llobet
,
M. O.
Steinmetz
,
A. R.
Bausch
,
M.
Distel
,
J.
Thorn-Seshold
,
A.
Akhmanova
, and
O.
Thorn-Seshold
, “
In vivo photocontrol of microtubule dynamics and integrity, migration and mitosis, by the potent GFP-imaging-compatible photoswitchable reagents SBTubA4P and SBTub2M
,”
J. Am. Chem. Soc.
144
,
5614
5628
(
2022
).
106.
M.
Wranik
,
T.
Weinert
,
C.
Slavov
,
T.
Masini
,
A.
Furrer
,
N.
Gaillard
,
D.
Gioia
,
M.
Ferrarotti
,
D.
James
,
H.
Glover
,
M.
Carrillo
,
D.
Kekilli
,
R.
Stipp
,
P.
Skopintsev
,
S.
Brünle
,
T.
Mühlethaler
,
J.
Beale
,
D.
Gashi
,
K.
Nass
,
D.
Ozerov
,
P. J. M.
Johnson
,
C.
Cirelli
,
C.
Bacellar
,
M.
Braun
,
M.
Wang
,
F.
Dworkowski
,
C.
Milne
,
A.
Cavalli
,
J.
Wachtveitl
,
M. O.
Steinmetz
, and
J.
Standfuss
, “
Watching the release of a photopharmacological drug from tubulin using time-resolved serial crystallography
,”
Nat. Commun.
14
,
903
(
2023
).
107.
M.
Borowiak
,
F.
Küllmer
,
F.
Gegenfurtner
,
S.
Peil
,
V.
Nasufovic
,
S.
Zahler
,
O.
Thorn-Seshold
,
D.
Trauner
, and
H.-D.
Arndt
, “
Optical manipulation of F-actin with photoswitchable small molecules
,”
J. Am. Chem. Soc.
142
,
9240
9249
(
2020
).
108.
S.
Pospich
,
F.
Küllmer
,
V.
Nasufović
,
J.
Funk
,
A.
Belyy
,
P.
Bieling
,
H.-D.
Arndt
, and
S.
Raunser
, “
Cryo-EM resolves molecular recognition of an optojasp photoswitch bound to actin filaments in both switch states
,”
Angew. Chem. Int. Ed.
60
,
8678
8682
(
2021
).
109.
F.
Küllmer
,
N. A.
Vepřek
,
M.
Borowiak
,
V.
Nasufović
,
S.
Barutzki
,
O.
Thorn-Seshold
,
H.-D.
Arndt
, and
D.
Trauner
, “
Next generation Opto-Jasplakinolides enable local remodeling of actin networks
,”
Angew. Chem. Int. Ed.
61
,
e202210220
(
2022
).
110.
H.
Zhang
,
C.
Wang
,
T.
Jiang
,
H.
Guo
,
G.
Wang
,
X.
Cai
,
L.
Yang
,
Y.
Zhang
,
H.
Yu
,
H.
Wang
, and
K.
Jiang
, “
Microtubule-targetable fluorescent probe: Site-specific detection and super-resolution imaging of ultratrace tubulin in microtubules of living cancer cells
,”
Anal. Chem.
87
,
5216
5222
(
2015
).
111.
J. J.
Keya
,
R.
Suzuki
,
A. M. R.
Kabir
,
D.
Inoue
,
H.
Asanuma
,
K.
Sada
,
H.
Hess
,
A.
Kuzuya
, and
A.
Kakugo
, “
DNA-assisted swarm control in a biomolecular motor system
,”
Nat. Commun.
9
,
453
(
2018
).
112.
M.
Akter
,
J. J.
Keya
,
A. M. R.
Kabir
,
H.
Asanuma
,
K.
Murayama
,
K.
Sada
, and
A.
Kakugo
, “
Photo-regulated trajectories of gliding microtubules conjugated with DNA
,”
Chem. Commun.
56
,
7953
7956
(
2020
).
113.
M.
Akter
,
J. J.
Keya
,
K.
Kayano
,
A. M. R.
Kabir
,
D.
Inoue
,
H.
Hess
,
K.
Sada
,
A.
Kuzuya
,
H.
Asanuma
, and
A.
Kakugo
, “
Cooperative cargo transportation by a swarm of molecular machines
,”
Sci. Robot.
7
,
eabm0677
(
2022
).
114.
S.
Kirchner
,
A.-L.
Leistner
,
P.
Gödtel
,
A.
Seliwjorstow
,
S.
Weber
,
J.
Karcher
,
M.
Nieger
, and
Z.
Pianowski
, “
Hemipiperazines as peptide-derived molecular photoswitches with low-nanomolar cytotoxicity
,”
Nat. Commun.
13
,
6066
(
2022
).
115.
H.
Inaba
,
M.
Sakaguchi
,
S.
Watari
,
S.
Ogawa
,
A. M. R.
Kabir
,
A.
Kakugo
,
K.
Sada
, and
K.
Matsuura
, “
Reversible photocontrol of microtubule stability by spiropyran-conjugated Tau-derived peptides
,”
ChemBioChem
24
,
e202200782
(
2023
).
116.
F. B.
Peters
,
A.
Brock
,
J.
Wang
, and
P. G.
Schultz
, “
Photocleavage of the polypeptide backbone by 2-nitrophenylalanine
,”
Chem. Biol.
16
,
148
152
(
2009
).
117.
A.
Taniguchi
,
Y.
Sohma
,
M.
Kimura
,
T.
Okada
,
K.
Ikeda
,
Y.
Hayashi
,
T.
Kimura
,
S.
Hirota
,
K.
Matsuzaki
, and
Y.
Kiso
, “
Click peptide based on the O-acyl isopeptide method: Control of Aβ1 − 42 production from a photo-triggered Aβ1 − 42 analogue
,”
J. Am. Chem. Soc.
128
,
696
697
(
2006
).
118.
K.
Katayama
,
S.
Tsukiji
,
T.
Furuta
, and
T.
Nagamune
, “
A bromocoumarin-based linker for synthesis of photocleavable peptidoconjugates with high photosensitivity
,”
Chem. Commun.
42
,
5399
5401
(
2008
).
119.
A.
Shigenaga
,
D.
Tsuji
,
N.
Nishioka
,
S.
Tsuda
,
K.
Itoh
, and
A.
Otaka
, “
Synthesis of a stimulus-responsive processing device and its application to a nucleocytoplasmic shuttle peptide
,”
ChemBioChem
8
,
1929
1931
(
2007
).
120.
A.
Shigenaga
,
J.
Yamamoto
,
Y.
Sumikawa
,
T.
Furuta
, and
A.
Otaka
, “
Development and photo-responsive peptide bond cleavage reaction of two-photon near-infrared excitation-responsive peptide
,”
Tetrahedron Lett.
51
,
2868
2871
(
2010
).
121.
A.
Shigenaga
,
J.
Yamamoto
,
N.
Nishioka
, and
A.
Otaka
, “
Enantioselective synthesis of stimulus-responsive amino acid via asymmetric α-amination of aldehyde
,”
Tetrahedron
66
,
7367
7372
(
2010
).
122.
A. E.
Mangubat-Medina
,
S. C.
Martin
,
K.
Hanaya
, and
Z. T.
Ball
, “
A vinylogous photocleavage strategy allows direct photocaging of backbone amide structure
,”
J. Am. Chem. Soc.
140
,
8401
8404
(
2018
).
123.
C.
Bosques
and
B.
Imperiali
, “
Photolytic control of peptide self-assembly
,”
J. Am. Chem. Soc.
125
,
7530
7531
(
2003
).
124.
Y.
Xiang
,
H.
Mao
,
S.
Tong
,
C.
Liu
,
R.
Yan
,
L.
Zhao
,
L.
Zhu
, and
C.
Bao
, “
A facile and versatile approach to construct photoactivated peptide hydrogels by regulating electrostatic repulsion
,”
ACS Nano
17
,
5536
5547
(
2023
).
125.
M.
Furutani
,
A.
Uemura
,
A.
Shigenaga
,
C.
Komiya
,
A.
Otaka
, and
K.
Matsuura
, “
A photoinduced growth system of peptide nanofibres addressed by DNA hybridization
,”
Chem. Commun.
51
,
8020
8022
(
2015
).
126.
H.
Inaba
,
A.
Uemura
,
K.
Morishita
,
T.
Kohiki
,
A.
Shigenaga
,
A.
Otaka
, and
K.
Matsuura
, “
Light-induced propulsion of a giant liposome driven by peptide nanofibre growth
,”
Sci. Rep.
8
,
6243
(
2018
).
127.
H.
Inaba
,
K.
Hatta
, and
K.
Matsuura
, “
Directional propulsion of DNA microspheres based on light-induced asymmetric growth of peptide nanofiber
,”
ACS Appl. Bio Mater.
4
,
5425
5434
(
2021
).
128.
A. S.
Carlini
,
M. A.
Touve
,
H.
Fernández-Caro
,
M. P.
Thompson
,
M. F.
Cassidy
,
W.
Cao
, and
N. C.
Gianneschi
, “
UV-responsive cyclic peptide progelator bioinks
,”
Faraday Discuss.
219
,
44
57
(
2019
).
129.
L. A.
Haines
,
K.
Rajagopal
,
B.
Ozbas
,
D. A.
Salick
,
D. J.
Pochan
, and
J. P.
Schneider
, “
Light-activated hydrogel formation via the triggered folding and self-assembly of a designed peptide
,”
J. Am. Chem. Soc.
127
,
17025
17029
(
2005
).
130.
T.
Muraoka
,
C.-Y.
Koh
,
H.
Cui
, and
S. I.
Stupp
, “
Light-triggered bioactivity in three dimensions
,”
Angew. Chem. Int. Ed.
48
,
5946
5949
(
2009
).
131.
L.
Awad
,
N.
Jejelava
,
R.
Burai
, and
H. A.
Lashuel
, “
A new caged-glutamine derivative as a tool to control the assembly of glutamine-containing amyloidogenic peptides
,”
ChemBioChem
17
,
2353
2360
(
2016
).
132.
P. J.
Salveson
,
S.
Haerianardakani
,
A.
Thuy-Boun
,
A. G.
Kreutzer
, and
J. S.
Nowick
, “
Controlling the oligomerization state of Aβ-derived peptides with light
,”
J. Am. Chem. Soc.
140
,
5842
5852
(
2018
).
133.
W.
Ji
,
S.
Zhang
,
S.
Yukawa
,
S.
Onomura
,
T.
Sasaki
,
K.
Miyazawa
, and
Y.
Zhang
, “
Regulating higher-order organization through the synergy of two self-sorted assemblies
,”
Angew. Chem. Int. Ed.
57
,
3636
3640
(
2018
).
134.
M. E.
Vázquez
,
M.
Nitz
,
J.
Stehn
,
M. B.
Yaffe
, and
B.
Imperiali
, “
Fluorescent caged phosphoserine peptides as probes to investigate phosphorylation-dependent protein associations
,”
J. Am. Chem. Soc.
125
,
10150
10151
(
2003
).
135.
T.
Hiraoka
and
I.
Hamachi
, “
Caged RNase: Photoactivation of the enzyme from perfect off-state by site-specific incorporation of 2-nitrobenzyl moiety
,”
Bioorg. Med. Chem. Lett.
13
,
13
15
(
2003
).
136.
H.
Li
,
J.-M.
Hah
, and
D. S.
Lawrence
, “
Light-mediated liberation of enzymatic activity: ‘Small molecule’ caged protein equivalents
,”
J. Am. Chem. Soc.
130
,
10474
10475
(
2008
).
137.
S. F. A.
Wouters
,
E.
Wijker
, and
M.
Merkx
, “
Optical control of antibody activity by using photocleavable bivalent peptide–DNA locks
,”
ChemBioChem
20
,
2463
2466
(
2019
).
138.
G. M.
Grotenbreg
,
M. J.
Nicholson
,
K. D.
Fowler
,
K.
Wilbuer
,
L.
Octavio
,
M.
Yang
,
A. K.
Chakraborty
,
H. L.
Ploegh
, and
K. W.
Wucherpfennig
, “
Empty class II major histocompatibility complex created by peptide photolysis establishes the role of DM in peptide association
,”
J. Biol. Chem.
282
,
21425
21436
(
2007
).
139.
B.
Jedlitzke
,
Z.
Yilmaz
,
W.
Dörner
, and
H. D.
Mootz
, “
Photobodies: Light-activatable single-domain antibody fragments
,”
Angew. Chem. Int. Ed.
59
,
1506
1510
(
2020
).
140.
E.
Uhl
,
F.
Wolff
,
S.
Mangal
,
H.
Dube
, and
E.
Zanin
, “
Light-controlled cell-cycle arrest and apoptosis
,”
Angew. Chem. Int. Ed.
60
,
1187
1196
(
2021
).
141.
T.
Kawakami
,
H.
Cheng
,
S.
Hashiro
,
Y.
Nomura
,
S.
Tsukiji
,
T.
Furuta
, and
T.
Nagamune
, “
A caged phosphopeptide-based approach for photochemical activation of kinases in living cells
,”
ChemBioChem
9
,
1583
1586
(
2008
).
142.
K. A.
Mosiewicz
,
L.
Kolb
,
A. J.
van der Vlies
,
M. M.
Martino
,
P. S.
Lienemann
,
J. A.
Hubbell
,
M.
Ehrbar
, and
M. P.
Lutolf
, “
In situ cell manipulation through enzymatic hydrogel photopatterning
,”
Nat. Mater.
12
,
1072
1078
(
2013
).
143.
P. H. N.
Celie
,
M.
Toebes
,
B.
Rodenko
,
H.
Ovaa
,
A.
Perrakis
, and
T. N. M.
Schumacher
, “
UV-induced ligand exchange in MHC class I protein crystals
,”
J. Am. Chem. Soc.
131
,
12298
12304
(
2009
).
144.
M. P.
Negroni
and
L. J.
Stern
, “
The N-terminal region of photocleavable peptides that bind HLA-DR1 determines the kinetics of fragment release
,”
PLoS One
13
,
e0199704
(
2018
).
145.
A.
Jiménez-Balsa
,
E.
Pazos
,
B.
Martínez-Albardonedo
,
J. L.
Mascareñas
, and
M. E.
Vázquez
, “
Temporary electrostatic impairment of DNA recognition: Light-driven DNA binding of peptide dimers
,”
Angew. Chem. Int. Ed.
51
,
8825
8829
(
2012
).
146.
M.
Wirkner
,
J. M.
Alonso
,
V.
Maus
,
M.
Salierno
,
T. T.
Lee
,
A. J.
García
, and
A.
del Campo
, “
Triggered cell release from materials using bioadhesive photocleavable linkers
,”
Adv. Mater.
23
,
3907
3910
(
2011
).
147.
S.
Sur
,
J. B.
Matson
,
M. J.
Webber
,
C. J.
Newcomb
, and
S. I.
Stupp
, “
Photodynamic control of bioactivity in a nanofiber matrix
,”
ACS Nano
6
,
10776
10785
(
2012
).
148.
S.
Nishimura
,
N.
Hokazono
,
Y.
Taki
,
H.
Motoda
,
Y.
Morita
,
K.
Yamamoto
,
N.
Higashi
, and
T.
Koga
, “
Photocleavable peptide–poly(2-hydroxyethyl methacrylate) hybrid graft copolymer via postpolymerization modification by click chemistry to modulate the cell affinities of 2D and 3D materials
,”
ACS Appl. Mater. Interfaces
11
,
24577
24587
(
2019
).
149.
S.
Petersen
,
J. M.
Alonso
,
A.
Specht
,
P.
Duodu
,
M.
Goeldner
, and
A.
del Campo
, “
Phototriggering of cell adhesion by caged cyclic RGD peptides
,”
Angew. Chem. Int. Ed.
47
,
3192
3195
(
2008
).
150.
Y.
Ohmuro-Matsuyama
and
Y.
Tatsu
, “
Photocontrolled cell adhesion on a surface functionalized with a caged arginine-glycine-aspartate peptide
,”
Angew. Chem. Int. Ed.
47
,
7527
7529
(
2008
).
151.
T. T.
Lee
,
J. R.
García
,
J. I.
Paez
,
A.
Singh
,
E. A.
Phelps
,
S.
Weis
,
Z.
Shafiq
,
A.
Shekaran
,
A.
del Campo
, and
A. J.
García
, “
Light-triggered in vivo activation of adhesive peptides regulates cell adhesion, inflammation and vascularization of biomaterials
,”
Nat. Mater.
14
,
352
360
(
2015
).
152.
T.
Kakiyama
,
K.
Usui
,
K.
Tomizaki
,
M.
Mie
,
E.
Kobatake
, and
H.
Mihara
, “
A peptide release system using a photo-cleavable linker in a cell array format for cell-toxicity analysis
,”
Polym. J.
45
,
535
539
(
2013
).
153.
Y.
Watai
,
I.
Sase
,
H.
Shiono
, and
Y.
Nakano
, “
Regulation of nuclear import by light-induced activation of caged nuclear localization signal in living cells
,”
FEBS Lett.
488
,
39
44
(
2001
).
154.
M. E.
Hahn
and
T. W.
Muir
, “
Photocontrol of Smad2: A multiphosphorylated cell-signaling protein, through caging of activating phosphoserines
,”
Angew. Chem. Int. Ed.
43
,
5800
5803
(
2004
).
155.
J.-P.
Pellois
and
T. W.
Muir
, “
A ligation and photorelease strategy for the temporal and spatial control of protein function in living cells
,”
Angew. Chem. Int. Ed.
44
,
5713
5717
(
2005
).
156.
R. V.
Rughani
,
M. C.
Branco
,
D. J.
Pochan
, and
J. P.
Schneider
, “
De novo design of a shear-thin recoverable peptide-based hydrogel capable of intrafibrillar photopolymerization
,”
Macromolecules
43
,
7924
7930
(
2010
).
157.
Y.
Ding
,
Y.
Li
,
M.
Qin
,
Y.
Cao
, and
W.
Wang
, “
Photo-cross-linking approach to engineering small tyrosine-containing peptide hydrogels with enhanced mechanical stability
,”
Langmuir
29
,
13299
13306
(
2013
).
158.
R.
Pugliese
,
M.
Montuori
, and
F.
Gelain
, “
Bioinspired photo-crosslinkable self-assembling peptides with pH-switchable ‘on–off’ luminescence
,”
Nanoscale Adv.
4
,
447
456
(
2022
).
159.
D. P.
Smith
,
J.
Anderson
,
J.
Plante
,
A. E.
Ashcroft
,
S. E.
Radford
,
A. J.
Wilson
, and
M. J.
Parker
, “
Trifluoromethyldiazirine: An effective photo-induced cross-linking probe for exploring amyloid formation
,”
Chem. Commun.
2008
,
5728
5730
.
160.
G. W.
Preston
,
S. E.
Radford
,
A. E.
Ashcroft
, and
A. J.
Wilson
, “
Analysis of amyloid nanostructures using photo-cross-linking: In situ comparison of three widely used photo-cross-linkers
,”
ACS Chem. Biol.
9
,
761
768
(
2014
).
161.
F.
Kotzyba-Hibert
,
I.
Kapfer
, and
M.
Goeldner
, “
Recent trends in photoaffinity labeling
,”
Angew. Chem. Int. Ed.
34
,
1296
1312
(
1995
).
162.
J.
Das
, “
Aliphatic diazirines as photoaffinity probes for proteins: Recent developments
,”
Chem. Rev.
111
,
4405
4417
(
2011
).
163.
X.
Li
and
T. M.
Kapoor
, “
Approach to profile proteins that recognize post-translationally modified histone “tails
,””
J. Am. Chem. Soc.
132
,
2504
2505
(
2010
).
164.
J. Z.
Hui
,
S.
Tamsen
,
Y.
Song
, and
A.
Tsourkas
, “
LASIC: Light activated site-specific conjugation of native IgGs
,”
Bioconjugate Chem.
26
,
1456
1460
(
2015
).
165.
J.
Park
,
Y.
Lee
,
B. J.
Ko
, and
T. H.
Yoo
, “
Peptide-directed photo-cross-linking for site-specific conjugation of IgG
,”
Bioconjugate Chem.
29
,
3240
3244
(
2018
).
166.
S.
Watari
,
H.
Inaba
,
T.
Tamura
,
A. M. R.
Kabir
,
A.
Kakugo
,
K.
Sada
,
I.
Hamachi
, and
K.
Matsuura
, “
Light-induced stabilization of microtubules by photo-crosslinking of Tau-derived peptide
,”
Chem. Commun.
58
,
9190
9193
(
2022
).
167.
S.
Yusa
,
M.
Sugahara
,
T.
Endo
, and
Y.
Morishima
, “
Preparation and characterization of a ph-responsive nanogel based on a photo-cross-linked micelle formed from block copolymers with controlled structure
,”
Langmuir
25
,
5258
5265
(
2009
).
168.
J.
Ding
,
X.
Zhuang
,
C.
Xiao
,
Y.
Cheng
,
L.
Zhao
,
C.
He
,
Z.
Tang
, and
X.
Chen
, “
Preparation of photo-cross-linked pH-responsive polypeptide nanogels as potential carriers for controlled drug delivery
,”
J. Mater. Chem.
21
,
11383
11391
(
2011
).
169.
G. A.
Bullen
,
J. H. R.
Tucker
, and
A. F. A.
Peacock
, “
Exploiting anthracene photodimerization within peptides: Light induced sequence-selective DNA binding
,”
Chem. Commun.
51
,
8130
8133
(
2015
).
170.
J. P.
Joseph
,
A.
Singh
,
D.
Gupta
,
C.
Miglani
, and
A.
Pal
, “
Tandem interplay of the host–guest interaction and photoresponsive supramolecular polymerization to 1D and 2D functional peptide materials
,”
ACS Appl. Mater. Interfaces
11
,
28213
28220
(
2019
).
171.
D.
Taura
,
A.
Urushima
,
Y.
Sugioka
,
N.
Ousaka
, and
E.
Yashima
, “
Remote-controlled regio- and diastereodifferentiating photodimerization of a dynamic helical peptide-bound 2-substituted anthracene
,”
Chem. Commun.
56
,
13433
13436
(
2020
).
172.
V.
Emiliani
,
E.
Entcheva
,
R.
Hedrich
,
P.
Hegemann
,
K. R.
Konrad
,
C.
Lüscher
,
M.
Mahn
,
Z.
Pan
,
R. R.
Sims
,
J.
Vierock
, and
O.
Yizhar
, “
Optogenetics for light control of biological systems
,”
Nat. Rev. Methods Primers
2
,
55
(
2022
).
173.
S.
Murata
,
T.
Toyota
,
S. M.
Nomura
,
T.
Nakakuki
, and
A.
Kuzuya
, “
Molecular cybernetics: challenges toward cellular chemical artificial intelligence
,”
Adv. Funct. Mater.
32
,
2201866
(
2022
).