Skip to main content

Supramolecular assembly of protein building blocks: from folding to function

Abstract

Several phenomena occurring throughout the life of living things start and end with proteins. Various proteins form one complex structure to control detailed reactions. In contrast, one protein forms various structures and implements other biological phenomena depending on the situation. The basic principle that forms these hierarchical structures is protein self-assembly. A single building block is sufficient to create homogeneous structures with complex shapes, such as rings, filaments, or containers. These assemblies are widely used in biology as they enable multivalent binding, ultra-sensitive regulation, and compartmentalization. Moreover, with advances in the computational design of protein folding and protein–protein interfaces, considerable progress has recently been made in the de novo design of protein assemblies. Our review presents a description of the components of supramolecular protein assembly and their application in understanding biological phenomena to therapeutics.

1 Introduction

Supramolecular assembly is a very common phenomena in nature and these natural supramolecular proteins have various structures, from simple structures to complex structures [1, 2]. As the increasing development of structural analysis, such as Cryo-EM, it is possible to observe the conformation of supramolecular assembled structure [3]. Before the 1990s, protein structures were scarcely elucidated; however, more protein structures have been identified by advanced technology (Fig. 1). More structures emerge from oligomeric proteins than monomeric proteins, providing us more opportunities to study protein assembly [3]. In this paper, we will discuss about the component of supramolecular protein assembly and classify their conformation according to structural dimensionality. Lastly, we speculate the biological phenomena from organism and their application with their own properties.

Fig. 1
figure 1

Statistical analysis of annually defined structure in Protein Data Bank (PDB) (Data obtained from https://www.rcsb.org/stats)

2 The component of protein assembly

We define the components of protein assembly in three aspects: folding structure unit, protein–protein interface, and assembly symmetry (Fig. 2). Each components determine the overall structure of supramolecular protein structure and their functions.

Fig. 2
figure 2

Schematic image for description of three components of supramolecular protein assembly

2.1 Folding structure unit

The folding unit is the basic building block for supramolecular protein assembly [4]. These building blocks are mostly composed of α-helices, β-sheets, or a mixture of them (Fig. 3A). The symmetry of the units and dimensions of the supramolecular protein are determined, depending on how the folding unit is configured. A better understanding of the proteins contributes to designing new supramolecular structures from rational design to de novo design. Rational design is inspired by common features and motifs from existing proteins [5, 6]. The α-helical structure is well characterized by Crick parameterization, and the sequence pattern of heptad repeats has been widely studied [7]. This enables the rational design of the alpha helical coiled-coil motif. Recently, the de novo design of proteins that start from scratch can create a new topology beyond experimentally determined structures. De novo design generates a building block based on the basic physical principle of protein by using computational power. There are many open-source programs that can help us build α-helices or β-sheets and complex forms of building units. To build a coiled-coil motif, we can easily build unit structures from CCCP (Coiled-coil Crick Parameterization) [7], CCBuilder2.0 [8] This allows researchers to build generalized models of coiled coils based on the Crick parameters or to calculate the folding stability of the resulting coiled coils [9]. The design of the β-sheet structure has non-local interactions where more β-sheet fractions give some sequence distance between each β-sheet unit, leading to slower folding rate [10] or misfolding [11, 12]. So, we need to predict and calculate the proper distance for every different scaffold of β-sheet structures. ‘BluePrintBDR’ mover in rosetta makes it easier to create a new topology structure based on the 2D map indicating which residue is pairing with a particular residue. The development of computational power and increased protein structure databases has expanded our knowledge to understand and design folding units.

Fig. 3
figure 3

Representative example of folding unit: α-helix, β-sheet, and α/β mixed structure

2.2 Protein–protein interface

The protein interface is a major determinant of the structure and function of natural proteins [13]. In nature, the signal transduction pathway is determined very precisely according to the difference in the protein interface of each ligand at the binding site of the multiligand receptor [14]. Among them, the supramolecular building block determines the binding stoichiometry between unit structures and controls the overall morphology according to the energetic favorability of each facet. The protein interface is delicately determined according to the size of the exposed solvent-accessible area [15, 16], hydrophobic packing [17, 18], and the existence of hot-spot residues [19, 20]. It can be classified into three categories based on importance and structure: helix interface, beta-sheet elongation, and metal coordinates.

2.2.1 Helical interface

The helical interface is the most common structure of the protein interface [21,22,23,24]. It is used in various structural regulations, from the very small protein assembly structure of the coiled-coil to heterogenous protein assembly, such as bacteriophage virion coating. The helix structure appears to be a simple helical structure; however, it is based on how each helix structure is connected in the helix-turn-helix structure and new interface such as parallel/antiparallel alignment can be created.

Amphiphilic interfaces of the helix are important for designating the directionality of the assembly. Hydrophobic patches at the protein–protein interface allow the monomers to form stable complexes [25]. In contrast, unfavorable buried polar residues destabilize the bonding interface [26]. Therefore, modulating the amphiphilic interface, which includes both hydrophobic and hydrophilic interactions, is a key component of protein assembly.

2.2.2 Beta-strand elongation

The beta-strand interface, which intermolecular network of hydrogen bonds, are widely occur in protein assembly. With the development of protein structure analysis technology, the morphology of the beta-sheet fiber structure of various structures was discovered and classified according to the beta strand elongation direction (parallel/antiparallel) and the symmetry of the steric zippers forming the core of the fiber structure [27]. The interface of beta-strand used to design many artificial amyloid-like structures assembled in the form of fibrils [28, 29] and gel [30, 31] and tandem repeat structures of ring or pore shape. The design of the beta-strand interface is very complex because the sheets bond intermolecularly and do not elongate except when the beta-strand is fully aligned within an axis. Recently, Baker et al. demonstrated the first de novo designed beta-barrel structures by establishing the geometric parameters of antiparallel beta-barrels to build an ideal backbone template and assign residues for structural stabilization [32].

2.2.3 Metal coordinates

Metal coordination plays an essential role in developing unit protein structures for regulatory and enzymatic reactions, including metalloenzymes, chloroplasts, and zinc-finger families [33]. Metal coordinates that exist in nature are necessary for properly folding proteins and creating complex quaternary structures through the bridging of metal ions [34, 35]. As a representative example of metal coordination, zinc finger proteins utilize a coordination zinc ion to maintain the folded structure, which is important for recognizing target DNA for transcriptional regulation [36, 37]. In addition, since metal coordination reacts very sensitively to pH change, the assembly pattern is diversified according to the tissue environment. Salgado et al. suggested the concept of MeTIR, metal-templated interface redesign, which induces self-assembly of the supramolecular structure in the presence or absence of metals [35].

2.3 Assembly symmetry

The overall assembly structure is strongly limited by the number of ways in which protein subunits may associate in three dimensions. Protein assembly can have various structures that can create a more complex and high-order hierarchical structure through a combination of the symmetry of the folding unit. For example, if a possible combination is predicted with two oligomeric components, each with C2 symmetry and D3 symmetry, a total of 11 candidates such as p312, P622, and P4132 are possible (Fig. 4) [38]. This combination modulates the overall size by controlling the number of folding units that determine the overall supramolecular structure according to the difference in the affinity of the interface. Moreover, an internally symmetric folding unit can designate the dimensions of the assembly pattern with open or closed symmetries. When the interface of the folding unit is open, because the unit is assembled in one direction, the entire structure is propagated in the form of a fibril or 2D sheet to form an assembled structure of various sizes. However, in the case of a closed interface, a cage or cyclic structure, in which the size of the entire structure is defined, is formed.

Fig. 4
figure 4

Schematic image of symmetric oligomeric building blocks and example two-component assembled structure. A Illustration of point group symmetries. B Representative example of octahedral symmetry constructed by combining C3 trimer and C4 tetramer (left), p6 symmetry constructed by combining C3 trimer and C6 hexamer (middle), and p422 symmetry assembled by combining D2 tetramer and D4 octamer (right) (Image reprinted with permission from Ref [38])

3 Various structures of supramolecular protein assembly

Supramolecular protein assemblies can be classified according to their orientation along a dimensional axis (Fig. 5). Dimensions are defined in the direction of overall structural expansion as the folding structural units are assembled. Dimensions are classified according to the number of axes that contribute to the addition of units and expansion of the overall structure: Unidimensional assembly, in which units assemble along a single axis to elongate the protein, Bidimensional assembly, in which units join together along a dual axis to expand a protein, and omnidimensional assembly, in which units are assembled along three or more axes to extend a protein.

Fig. 5
figure 5

Various assembled structure and their folding unit of supramolecular protein assembly

3.1 Unidimensional assembly

Proteins with a unidimensional assembly are elongated as the unit structures are added along a single axis. Typical examples of unidimensional assembly are fiber structures, including collagen [39,40,41,42,43], amyloid [28, 29, 44,45,46,47,48], and actin filament [49,50,51,52]. Collagen is a linear triple-helix structured protein [39] that accounts for approximately 70% of the extracellular matrix dry weight; thus, it is the most abundant protein in mammals. The single left-handed polyproline II-type helix composed of the repetition of Xaa-Yaa-Gly assemble into triple helix bundle with right-handed conformation through the hydrogen bonds between neighboring strands, especially N–H(Gly)⋯O = C(Xaa), to bury Gly residues toward the core of the triple helix and expose the other residues to the outer surface. Amyloids are elongated, unbranched β-sheet fibrils that can cause various diseases. There are many different protein units that form amyloid structures, such as β-amyloid peptide(Aβ) [29, 46], α-synuclein [47], and tau protein [48]. Individual β-strands assemble through hydrogen bonds to generate β-sheets, and several β-sheets twist around a central axis, assembling through a steric zipper interface, which is a tight dehydrated packing between neighboring β-sheets. Actin filament (F-actin) is a two-long-pitch helically stranded microfilament composed of an actin monomer called G-actin, which is an important component of the cell cytoskeleton and muscle tissue [50, 51]. Sault bridges, geometric surface complementarity, and hydrophobic interactions on the interstrand and intrastrand interface, particularly the loop compartment on the interface, stabilize F-actin.

Unidimensional assembly proteins, such as fibrils and fibers, can be designed to be extended along a one-dimensional axis to compensate for the short length, low mechanical strength, and stability of self-assembled protein in nature. Nanofibers are designed to form a desired structure and assembly with unique properties only under intended conditions such as temperature and pH to have high utility and specificity; therefore, they can be utilized in various fields such as biomaterials and medicine. In addition, the information obtained in the design process can provide a chance to reveal the mechanisms and properties of self-assembled nanofibers that exist in nature like collagen [39,40,41,42,43], amyloid [28, 29, 44,45,46,47,48], and actin filament [49,50,51,52], previously described, which are difficult to understand only with structural information [53, 54]. Based on this information, self-assembled nanofibers composed of simple alpha helix [55,56,57] or beta-strand [58,59,60,61,62,63,64] can be designed. For example, Ronak et al. designed de novo three-stranded β-sheet nanofibrils (TSS1) composed of 29 amino acids, mostly lysine and valine for self-assembling mechanically rigid hydrogel fabrication [58]. Novel self-assembled nanofibers composed of complex structures that do not exist in nature can be designed over self-assembled nanofibers composed of simple alpha helices or beta sheets. Sabine et al. designed complex βαβ unit structures that self-assemble into fibers with an alpha helical outer surface by linking two β-fibrillizing peptides to the alpha helix for stabilization and easier functionalization of the nanofiber [53].

3.2 Bidimensional assembly

Proteins with a bidimensional assembly are widened as unit structures are added along the two axes. One example of a bidimensional assembly is a sheet structure, including bacterial S-layers [65,66,67,68,69,70], which are two-dimensional arrays that coat the surface of bacteria or archaea [68]. S-layers are generated by the assembly of S-layer protein subunits through van der Waals force, ionic bonds, and hydrogen bonds [67], which develop various types of lattice structures from hexagonal (p3, p6), square (p4), and tilted (p1, p2) space groups [70].

The ring structure is another example of a bidimensional assembly, and it includes DNA clamps [71,72,73], helicases [74,75,76,77], nucleases [78,79,80], which are ring-shaped protein complexes that contribute to the metabolic process of DNA or RNA. DNA clamps are formed by the tight assembly of subunits that are composed of the α-helix and β-strand complex to generate a closed ring encircling DNA strand [71,72,73]. Subunits assemble into a ring structure with a positively charged inner surface; therefore, the ring can electrostatically interact with the negatively charged DNA strand [71]. In the case of helicase, four out of six superfamilies have a ring-shaped assembly structure [77], and they commonly take the form of a hexameric structure with a central pore. The core of the ring contains nucleotide binding sites between the subunits, which usually have an arginine finger that is involved in nucleoside triphosphate (NTP) binding and hydrolysis [74,75,76,77]. Several types of nucleases, including bacteriophage λ-exonuclease [78], E. coli RecE [79], and Cas4 nuclease SSO0001 [80], have a pore-containing toroid structure, and they exhibit diverse morphologies such as trimer [78], tetramer [79], and pentamer [80]. The assembly of these structures is mediated by subunit interfaces with hydrogen bonding, ion pairs, and hydrophobic interactions [79]. Some nucleases have unique funnel-shaped pores, which have a wide entrance and a narrow exit that can accept double-stranded DNA at the entrance and pass only the single-stranded DNA through exit [78, 79].

The other representative morphology of the bidimensional assembly is a beta-barrel structure. Beta barrel structures include β-barrel transmembrane proteins [81,82,83,84,85] and green fluorescence protein (GFP) [86,87,88]. Transmembrane proteins with β-barrel structure are located in the outer membranes of chloroplasts, gram-negative bacteria, and mitochondria [85], and most β-barrel transmembrane proteins are composed of an even number of strands arranged in anti-parallel [81,82,83,84,85]. β-barrel transmembrane proteins have abundant intrastrand hydrogen bond networks; therefore, they are stable in the membrane environment. GFP is a β-barrel protein that emits green fluorescence, and 11 strands of β-sheet form a cylindrical outer surface connected by short helical loop structures, protecting 3-amino acid fluorophores inside the barrel [86,87,88]. The polar interactions surrounding the central fluorophore mediate proton rearrangements, leading to the activation of GFP [87].

Through the two-dimensional protein array assembly design, proteins are repeated in a specific order. Thus, specific proteins designated by the tailor can be arranged in a constant valency and order. However, 2D protein arrays are very rare in nature, such as the exoskeleton of the surface layer of many bacteria; therefore, it is essential to design the properties of 2D protein arrays [89]. Modulating two docking axes and interfaces can determine the self-assembly mechanism and overall symmetry structure of bidimensional proteins and make them more diverse and controllable structural and functional 2D materials such as 2D arrays [90,91,92,93] that do not exist in nature. For example, Chen et al. designed 2D arrays of homodimer protein building blocks by modulating the 4-binding interface of the helix bundle following the C12 layer symmetry group [90].

The ring structure, which are bidimensionally assembled in a cyclic arrangement, is another representative example of a bidimensional assembly protein design. The protein pore size of the designed self-assembled nanoring can be controlled by the relationship between the oligomer state and pore diameter [94]. However, the de novo design of nanoring with large pores is challenging for thermodynamic stabilization because the large surface-area-to-volume ratio of the nanoring results in a low stabilizing interaction density [95]. The design of the alpha-helix ring structure [94,95,96,97,98] is simple owing to the specific parameter type of the backbone by the crick and versatility of the application using an outer alpha-helical surface. The beta barrel ring is very complicated because it tends to cause misfolding and aggregation easily if not properly controlled [32]. After developing structural understanding and computational design, the design of beta barrels to utilize rigid self-assembly ability has started [32, 99, 100]. In addition, through development of design method containing both alpha helix and beta sheets [101, 102], self-assembling proteins with both secondary structures can be designed for easier stabilization even with compact protein and expansion of design available protein structure pool. For example, Lim et al. increased the binding stability and strength of the unstable short alpha helix bundle by linking beta sheets as a self-assemble inducing segment with an alpha helix, which eventually forms a beta barrel inside the alpha ring [103].

3.3 Omnidimensional assembly

Proteins with an omnidimensional assembly are expanded as the unit structures are added along three or more axes. Tube structures, examples of omnidimensional assemblies, include helical virus capsid [104,105,106,107] and microtubules [108, 109]. Helical viruses, represented by the mosaic virus family [104,105,106,107], are composed of helical nucleic acid coils and capsid proteins, where capsid protein subunits cover the nucleic acid. The assembly of the helical virus is mediated by subunit-RNA interactions and subunit-subunit charge interactions. An abundant salt bridge from intersubunit ion pairs stabilizes the interface between the subunits. In addition, the electrostatic interactions between the RNA backbone and positively charged part of the protein derive the subunit-RNA interaction. Microtubules are cellular structures that form the cytoskeleton of eukaryotic cells [108, 109], and α/β-tubulin dimers polymerize into a cylindrical complex in a specific direction, leading the microtubule to have a subunit-adding site and subunit-dissociation site. Subunits assemble in a way that enables the additional attachment of several microtubule-binding proteins by negatively charged the outer microtubule surface by placing the acidic residues on the subunit C-terminal tails [109].

Another typical example of omnidimensional assembly is a cage structure, including polyhedral virus capsid [110,111,112,113], ferritin [114,115,116,117,118,119,120], lumazine synthase [121, 122], vaults [123,124,125,126], clathrin lattice [127], and heat shock proteins [128,129,130]. Polyhedral viruses are composed of central nucleic acids and surrounding polyhedral capsids, whose subunits are assembled through subunit-subunit interactions and subunit-RNA interactions [110]. Subunit-subunit interactions originate from integrating electrostatic repulsions and hydrophobic attractions, where the subunit-RNA electrostatic interactions originate from the interface between the negative-charge RNA backbone and the positive-charge N-terminal of the subunits [110,111,112,113]. For the other examples of cage structure, there are various protein complexes that control the life activities of diverse organisms, by storing the specific target material [114,115,116,117,118,119,120], performing enzymatic activities [121, 122], mediating cellular processes [123,124,125,126], coating vesicles [127], and protecting cell components from the stressful environment [128,129,130]. Ferritin is a hollow cage structure composed of a four-helix bundle subunit assembly, containing sufficient metal-protein interactions to generate iron binding sites, since this protein structure encapsulates iron [114,115,116,117,118,119,120]. The lumazine synthase cage is generated from the assembly of the pentamer consisting of five subunits, which are built with several β-strands and α-helixes [121, 122] abundant hydrogen bonds and ionic contacts between subunits allow lumazine synthase to function as favorably binding inhibitors [122]. Vaults are created by assembling three subunit proteins, including the major vault protein (MVP), a subunit consisting of several β-strands [123]. Assembly of vaults is mainly derived by helix-helix interactions, with polar residues facing the surface of the whole structure and hydrophobic residues located on the interface between neighboring helixes [123,124,125,126]. The clathrin lattice is an assembly of the clathrin subunits, which can form various cage structures from small cages of 28 and 36 assembly units to hexagonal arrays and soccer ball structures. They are stabilized through fixed contact patterns adjacent to each end of the strand [127]. Heat shock proteins, particularly chaperonin (HSP60), are another representative example of cage structure assembly. Chaperonins consist of two consecutive stacked rings composed of seven, eight, or nine subunits [128,129,130]. The inter-ring interface, which consists of electrostatic interactions between the positively charged residue in one ring and the negatively charged residue of the other ring, provides the structural basis of chaperonins [129].

In addition to the structural limitations of nature-derived self-assembled hierarchical nanostructures that have not suitable valency or size for various applications, researchers have designed highly ordered de novo supramolecular omnidimensional assembly structures such as nanotubes [55, 91, 131,132,133,134] and nanocage [135,136,137,138,139,140,141,142,143] with variable size, shape, and symmetry using computational tools. Designing omidimensional self-assembly structures with regularly repeating subunits in tailor-defined constant valency makes them suitable for use in various fields through interior or exterior surface functionalization. The subunit structure design of hierarchical nanostructures determines the overall size and shape of self-assembled nanostructures, such as the subunits per helical turn, helical pitch, and pore size of nanotubes, or the high-order symmetry of nanocages such as dihedral, tetrahedral, octahedral, and icosahedral. In designing a self-assembling nanocage, combining two types of symmetrical subunits can create nanocage structures with diverse high-order symmetry, size, and valency. [136,137,138,139,140,141, 144]. However, even if the same symmetric building blocks are used, they can be assembled into nanocages of different high-order structures with different valencies, depending on which symmetry axis alignment the blocks are arranged. For example, C3 and C2 building blocks with tetrahedral or icosahedral axes assembled in tetrahedral and icosahedral nanocages, respectively [137, 138]. In a different way, a single type of symmetrical subunit can also form a hierarchical supramolecular nanocage alone by designing the symmetric building block unit interface to self-assemble each other that does not participate in the previous assembly. Using this method, placing trimer proteins as C3 symmetry building blocks to three-fold rotational symmetry axes of tetrahedral or octahedral point group symmetry can form supramolecular nanocages with only one new interface design between the trimeric building blocks for self-assembly [140, 142].

4 Biology of supramolecular protein assembly

Naturally derived protein self-assembly or rationally designed folding units have shown the strength to construct various sophisticated protein nanostructures. It can produce enormous properties that cannot be created within a single monomeric protein conformer. Nature utilizes protein assembly to build complex biological phenomena such as signal transduction, cell growth, and immunology. Furthermore, protein assemblies can be applied as templates for development from functional biomaterials such as biomimetic materials to drug delivery platforms, biomedical diagnostics, and therapeutic platforms including vaccines. Here, we classify representative biology of supramolecular protein assemblies, such as matrix scaffolds of cell growth, encapsulation of functional cargo inside supramolecular structures, and hybridize designed proteins with unprecedented materials.

4.1 Matrix scaffold in cell growth

The nature-derived self-assembly of various proteins and peptides, such as collagens, proteoglycans, laminins, and fibronectin form the extracellular matrix (ECM) of each tissue, and their association determines the structure, function, and organization of the tissue. Hence, elucidating the mimic of ECM-derived biomolecules is an important aspect of cell biology and growth [145,146,147,148]. The utilization of naturally derived ECM as a cell culture substrate has become a viable option for cell growth. To this end, the self-assembly properties of various peptides, such as RAD16 [149,150,151], amyloids [152,153,154], have been utilized as substrates to control cell growth. Furthermore, ECM based on various adhesion proteins, such as collagen [155], fibronectin [156], and laminin [157] has been utilized as a scaffold for cell growth [158]. These proteins utilize cell-binding epitopes for binding to integrins, which consists of an adhesion protein derived small peptide sequence. For instance, the peptide sequence derived from collagen, fibronectin and laminin is RGD, RGDS and IKVAV/YIGSR, respectively. [159]. In addition, the self-assembly property of silk like protein (FN-silk from fibronectin) to form networks of microfibers has been utilized as an ECM. This association provides a three-dimensional (3D) microfiber network with specific sites for cell anchorage, which makes the cells remain viable for more than 90 days [160]. In addition, the spontaneous self-association of peptides can be utilized for cell growth facilities by employing matrix-like platforms. Particularly, the peptides with elongated hydrophobic hydrocarbon chains (HHC) or having hydrophilic groups on ends, -self-assembled with—external hydrophilic surfaces. These peptides, known as bolaamphiphile peptides, have a higher propensity to form flat layers upon self-assembly, making them a good candidate for working as substrates [161,162,163]. Recently, da Silva et al. utilized the self-assembly property of a bolaamphiphile peptide which form well defined nanosheets in water via association of the peptide backbone. The assembly is driven by the nucleation phase and serves as a substrate for cell growth matrix, which—proved its worth with human corneal stromal fibroblast (hCSF) cell growth [164]. On the other hand, the self-assembly property of laminin-derived peptides, triggered by slight hydrophobic modification, has been utilized for the formation of supramolecular structures suitable for cell growth applications [165,166,167]. To this field, most recently, Jain et al. controlled neuronal cell growth by a peptide-based supramolecular self-assembled matrix. The developed supramolecular gel was composed of self-assembly of two peptides (IKVAV and YIGSR), which facilitated the interaction between cells and the matrix. Their developed gel not only proved its worth by controlling neuronal cell growth and SHSY5Y neuroblastoma cells but also the proliferation in C6-glial cells [168]. In the field of gel-based matrices, another group of scientists made hydrogels from self-assembled peptides [167, 169, 170]. Recently, Aye et al. proposed the use of peptide-based hydrogels as biocompatible scaffolds for regenerative medicine. They developed a hydrogel composed of biomimetic inclusion of three different peptide sequences. During the in-vitro testing with human mammary fibroblasts cell culture, the hydrogel proved its worth by improved cell adhesion, growth, and proliferation [171]. Meanwhile, a group of scientists have explored the self-assembly of a set of dipeptides (X-ΔPhe) containing ΔPhe (α,β-dehydrophenylalanine) at their C-terminus and their possible use in drug delivery [172,173,174,175]. Inspired by the self-assembly property of dipeptides, Yadav et al. utilized a 3D platform, composed of dipeptide‑based hydrogel [174] as a scaffold for three dimensional cell growth. They utilized a leucine-α,β-dihydrophenylalanine (Leu-ΔPhe)-based hydrogel capable of forming hydrogels in the MPa range required for bone-like engineering stiffness at relatively low concentrations. Their results indicated the healthy condition of cells with good functionality, making Leu-Δphe hydrogel a suitable dipeptide candidate for three dimensional scaffold for cell culture [176].

4.2 Membrane associated protein assembly

Membrane proteins are proteins that span the cell membrane and play important roles in controlling cellular behavior, including selective molecular transport through the bilayer, nutrient uptake, and signal transduction [177, 178]. In particular, the transmembrane domain regulates cellular activity by organizing oligomers with specific symmetry. Most transmembrane domains have a α-helix structure and are reverse amphiphilic, with hydrophilic surfaces facing inward and hydrophobic surfaces facing outward. There were many attempts to mimic and engineer the nature-derived transmembrane protein, such as exporting key motifs of E.coli Wza to conduct ion and bind blocker [97], deleting specific domain of FhuA to conduct ion [179], attaching ligands on Salmonella typhi ClyA to selectively shuttle the large analyte proteins[180] or DNA molecules [181], and adding subunits of ClyA to vary the size of pore [182].

The de novo design can artificially build up the transmembrane protein with desired function or structure that is hardly found in nature. 12-helical potassium conducting protein that performs superior conductance of K+ over CH3NH3+, Cs+, Na+, Ba2+ is computationally designed and inserted into membrane through repositioning hydrophobic residues [95], and four-helical bundle containing two di-metal binding sites that selectively passes Zn2+ and Co2+ is reported [183]. The structure of concentric ring [95] and polar ionizable ligand containing complex [183] can be further utilized as a platform to selectively conduct various target ions, in ways that are not possible with native channels. The study of transmembrane pore that is large enough to pass the small molecule [95] can be further applicated in small molecule delivery system, which is attractive in medical and pharmaceutical fields. Furthermore, by computationally designing combinations of non-natural iron diphenylporphyrins and D2-symmetric four-helix bundles, it is possible to mimic the transmembrane electron transportation, a key component of ATP production and photosynthesis. This unique metaloprotein complex successfully transported electrons, which opens up the possibility of further designing artificial photosystems [96].

4.3 Nano-cage structure for surface displaying epitopes, encapsulation and cell signaling

Nanocage structures can be used for displaying epitopes, encapsulation, and cell signaling using multivalency, highly ordered repetitive arrays, and hollow structures. Multivalency and highly ordered repetitive arrays of clustered epitopes displayed on the surface of self-assembled nanostructures can increase avidity compared to single epitopes alone [184,185,186,187,188,189]. In addition, multiple types of surface functionalization units attached to the surface allow supramolecular structures to become multifunctional nanocarriers [190]. In addition, the hollow structure of the hierarchical structure reassembly and disassembly under specific conditions by subunit interface interactions can be used to encapsulate drugs or genes that make nanocages an effective delivery platform [191, 192].

Naturally derived self-assembly proteins such as ferritin and lumazine synthase have been used for cell signaling [186] and display of epitopes (antigens) from influenza [193], HIV-1 [194], Epstein-Barr virus [195], SARS-CoV-2 spike protein [188, 189], and Enterovirus 71 (EV71) [196] for immune response activation due to their biocompatibility, stability, and easy modification [197] (Fig. 6). The hollow structure of nature-derived self-assembly supramolecular proteins are also suitable for internal modification for delivery of chemicals or other protein delivery, such as GFP [198], HIV protease [199], daunomycin [192] carboplatin [191] and even show possibility of gene delivery [200] using a hollow supramolecular structure. For example. Uchida et al. synthesized human H-chain ferritin with iron oxide in the hollow and cancer cell-targeting peptide RGD-4C attached to the surface. This supramolecular showed increased cancer targeting ability without perturbation of the self-assembly of ferritin structure. [201] In addition, nanocage can be an effective drug delivery vehicle with a targeting function using the characteristic of self-assembled supramolecular at once [202].

Fig. 6
figure 6

Schematic image of application of a nanocage. A The ferritin nanocages displaying EV71 antigens on the surface as epitopes for the use of vaccine. B The BG505 SOSIP displaying de novo designed icosahedral nanocage. C The octahedral (up) and icosahedral (down) antibody nanocages with fusion of angiopoietin-1 F-domain (A1F) and Fc (Image reprinted with permission from Ref [135, 196, 203])

Although the application of self-assembled hierarchical nanocages derived from nature also shows a high utility value, it is unsuitable for various applications because of the limited number of structural elements in nature. However, de novo designed nanocages have various controllable shapes, sizes, and high-order symmetries, providing different surfaces and structures for antigen valency and spacing. Different valency and spacing of the functionalization unit can enhance the effect of drugs by allowing the desired nanocage to be selected according to the characteristics of the target ligand distance or uptake optimal size [184]. The application of various symmetric multivalent antigens presenting designed nanocages as vaccines was developed by genetic conjugation between the N-terminus of the trimer subunit and the C-terminal of trimeric viral glycoproteins and improved immunogenicity. (Fig. 6B) [203, 204] In a different way, Divine et al. designed self-assembled nanocage with various symmetry using antibody-Fc binders as a building block. They demonstrated enhanced target avidity experimentally by activating Tie-2 pathway (Fig. 6C) and enhancing SARS-CoV-2 neutralization with self-assembled supramolecular structures with each targeting antibody [135]. Through this, the de novo supramolecular protein cage could serve a significant role in improving effectiveness of immune activity and drug delivery since they have optimal size and superior avidity, which can effectively treat cancer and virus-derived diseases suffering from low rates of drug delivery rates.

4.4 Epitope recognition for immunology and diagnosis

Proteins are macromolecules that perform complex but essential tasks in living cells through the formation of protein clusters via self-assembly. For instance, naturally existing hemoproteins that can specifically recognize and associate with heme groups to capture oxygen molecules by their heme-prosthetic groups [205]. Moreover, the self-association of proteins or peptides, naturally existing or de novo designed, can be utilized to improve the diagnostic potential, especially in the detection of bacterial infection at the initial stage. Like, Liu et al. utilized the self-assembly of a rhodamine-modified peptide derivative for detecting bacterial infections in gram-positive bacteria [206]. For the same purpose, Qianet developed a method for sensitive and specific detection of infected phagocytic cells (Staphylococcus aureus), owing to molecular self-assembly [207]. Recently, Yang et al. developed a self-assembling peptide based probe for vancomycin in which a luminogen with aggregation-induced emission (AIEgen) was used as a responsive fluorescence turn-on motif [208]. Furthermore, the nanotubes composed of various self-assembled peptide have also been utilized for the detection of various bacterial pathogens, like E. coli [209] and other nitro-functionalized neurotoxins [210]. The self-assembly of another class of peptides, ion-complementary, has also been utilized for immobilization of various biomolecules, including enzymes and analytes. After its self-assembly, the peptides have outwardly oriented charged residues (K and E), which can be utilized for binding moieties such as glucose [211]. However, many researchers have utilized the self-assembly of proteins or peptides for the recognition of target species of concern. For instance, Bianchi et al. detected the oxidation of ammonia (sensitivity 2.83 L−1 cm−2) and urea (sensitivity 81.3 L−1 cm−2) through aligned deposition of peptide (diphenylalanine) microstructures onto thiolated gold electrodes [212]. Furthermore, Zhang et al. utilized the self-assembly of streptavidin protein to create arrays, facilitated by a DNA lattice, and utilized it for efficient protein detection. The developed system detects protein arrays (biotin) aided by rhodamine or other fluorescent probes [213]. Recently, Arai et al. developed a protein-sensing device from a self-organized glycopeptide bundle with glucose or galactose at the C-terminals. Because of this insertion, the glycopeptides rearranged to form a bundle that acted as an ion channel due to the interaction between the target protein (peanut lectin and concanavalin A) and the terminal sugar groups of the glycopeptides [214]. Furthermore, self-assembly has been utilized to diagnose the structural details of small proteins. Like, Liu et al. utilized protein engineering to create a self-assembled scaffolding system, making possible to see structural details of a small sized protein via cryo-EM. Assisted by a rigid alpha helical linker, they utilized the self-assembly of 17-KDa DARPin protein to create a cage with cubic symmetry. The resulting construct was analyzed via cryo-EM to explore the structural details of DARPin at near-atomic level (3.5-to 5-Ã… resolution) [215].

In addition, the self-assembled structures of peptides have been used to flexibly functionalize the devices for various biomolecular detections, avoiding complex lithography, and improving selectivity and performance. The functionalizations obtained from such self-assemblies are stable, both thermally and electrochemically. Like, Ryu et.al utilized the self-assembly of peptide building blocks into vertically aligned nanowires (thermally stable up to 200 Â°C) to enhance their capability for electrochemical detection of biomolecules. For this purpose, they utilized aromatic dipeptides such as diphenylalanine in aqueous solutions under ambient conditions [216]. The same class of peptides has been utilized for developing organic field effect transistors (O-FETs) for electrical measurements in which a self-assembled peptide layer was used as a dielectric in FET to sustain electric fields [217]. A similar approach was utilized by Gupta et al. as they developed the self-assembly of di-peptide based building blocks into ordered nanotubes with expanding stability over a wide range of pH [172].

4.5 Supramolecular assembly in enzymatic function

In fact, ribosomal proteins (r-proteins) self-assemble to produce large protein content in cells by catalyzing ribosome production, which drives this process. This nature-derived catalytic process helps ribosomes to reduce the time required to produce a new set of r-proteins, facilitating cell growth [218]. Inspired by this nature-derived self-assembly process for catalyzing cell growth, researchers have utilized the self-assembly of designer proteins or peptides to create catalytic sites [35, 219], especially metallo enzyme [220,221,222], which can efficiently tune the properties of a metal ion to catalyze difficult chemical transformations (Fig. 7). Like, Woon et al. utilized the self-assembly of monomeric redox protein already possessing catalytic zinc sites at its interfaces, to design an artificial metallo-β-lactamase enzyme. The designed catalyst was not only functional in the periplasm of gram negative bacteria (Escherichia coli), enabling them to survive in the presence of antibiotic like ampicillin, but also displayed catalytic proficiency for ampicillin hydrolysis [223]. The designed metalloenzyme has also been utilized to catalyze oxidation reactions. As Olga et al. designed the supramolecular self-assembly of a peptide assisted by copper, for efficient catalysis of oxidation of dimethoxyphenol in the presence of dioxygen [224]. In the recent Lee et al., have also contributed to the work by improving the efficiency of the oxygen reduction reaction (ORR) by designing a self-assembly (Hexcoil-Ala) peptide. The designed peptide readily assembles on single-walled carbon nanotubes and helps in dispersion in aqueous solutions. Moreover, through mutation of the cysteine residue, a size-controlled and well-dispersed arrangement of AuNPs around the designed peptide has made possible which in turns gives improved electronic properties for enhanced oxygen reduction reaction performance in fuel cells [225]. Meanwhile, another group of scientists utilized the self-assembly property of designed collagen-mimetic peptides (CMPs) to create a supramolecular structure. They improved the self-assembly property of CMPs via metal-histidine coordination method and utilized the resultant superstructure for to catalyze ester hydrolysis in the presence of Zn(II) ions [226]. Recently, Hyun et al. designed a metallo-catalyst for efficient and industrially applicable biological conversion of methane to methanol. For the purpose, they reassemble the native catalytic domains of an already existing enzyme (methane monooxygenase). Through their construct, they not only successfully synthesize a stable and soluble protein construct in Escherichia coli, but also improved the yield of methanol while retaining enzymatic activity [227]. In addition, another group of scientists utilized the self-assembly of naturally existing proteins to create catalytically active sites for their application in catalysis. Like, Rubinov et al. utilized the self-assembly of amphiphilic peptides to form various well-defined structures like β-sheets, β-plates, fibrils and nanotubes, etc. Through these structures, they synthesized monomeric peptides, starting from basic building blocks [228]. To this end, a group of scientists explored another pathway of utilizing the self-assembly of two structurally different proteins (for instance Bovine Serum Albumin and bacterial microcompartment domain protein, PduBB), which do not possess any catalytic activity individually. However, upon self-assembly, a floral nanohybrid is formed which can catalyze the oxidation of pyrogallol to purpurogallin [229]. Despite of the way, many researchers have utilized the self-assembly phenomena of designer proteins to enhance the catalytic activity of various enzymes. Like, Zhang et al. designed an artificial hydrolase via self-assembled peptide nanofibers as biological enzymes for catalyzing ester hydrolysis [219]. Also, Wang et al. utilized the self-assembly of de novo designed helical hepta-peptides to create a phosphate mimic. The hydrolysis efficiency of these designer catalysts is comparable to already available enzymes like adenosine triphosphatase (ATPase) and alkaline phosphatase (ALPase), and hence can serve as a substitute [230]. Meanwhile, another group of scientists utilized naturally existing proteins to enhance their native catalytic activity via self-assembly. Like, Soares et al. explored the self-assembly of lipopeptides in water and utilized them for catalysis (yield was raised to 499% with selectivity up to 85%) of aldol reaction, using cyclohexanone and p-nitrobenzaldehyde as reactants [231]. In addition, Shan et al. utilized self-assembled films made from three distinct proteins (chitosan, laponite, and hemoglobin) for their application in electrochemistry and catalysis for hydrogen peroxide production. The film exhibited long-term stability, while the cyclic voltammetry peak potentials remained unchanged, and the cathodic peak currents remained undeclined even after 60 days [232].

Fig. 7
figure 7

Example of enzymatic supramolecular structure via metal coordinate interaction (Image reprinted with permission from Ref [34])

4.6 Protein design in the unprecedented material

Inspired by the binding of amphipathic peptides with natural carbon materials such as graphite [233, 234], de novo designed peptides have been developed to induce self-assembled 2D structures on the surface of various carbon materials, such as graphene [235,236,237] or graphite (Fig. 8) [238, 239]. Remarkably, Grigoryan et al. designed peptides having capability of creating well textured surface by wrapping single-walled carbon nanotubes (SWNTs) in a structurally specific manner, creating a richly textured molecular surface. For binding with SWNTs, despite selecting peptides by phage-display or synthesizing peptides favoring SWNT binding, they developed a new design rule by utilizing a computational approach with an intrinsic recognition motif. They achieved higher-order assembly and dense packing by placing Gly and Ala in a repeating manner on a helix as the elementary structural unit [240]. Ko et al. thus, synthesized peptides that can self-assemble onto carbon nanotubes, the gold–platinum (AuPt) bimetallic nanostructures of which can induce a catalytic response to oxygen reduction [241].Meanwhile, a group of scientists developed the designer proteins or peptides that can direct self-assembly to produce ordered structures. Garima et al. developed a self-assembled monolayer induced by human serum albumin protein, that can convert 2D polyethylene glycol into discrete ring structures [242]. Similarly, Kim et al. demonstrated that the formation of ordered superstructures of buckminsterfullerene (C60) can also be directed by proteins [243]. In addition, researchers have utilized the self-assembly properties of proteins or peptides to hybridize other 2-dimensional carbon-based materials, such as graphene [244,245,246]. In this regard, Mustata et al. designed peptides forming two-dimensional monolayer crystals via self-assembly, eventually forming long, parallel, in-register b-sheets [247]. Despite designing peptides to self-assemble on graphene or graphite surfaces, some researchers have targeted pristine graphene owing to its excellent electrical properties and absence of any kind of functional moieties. For instance, No et al. developed nature-inspired designer peptides through the optimization of peptide-peptide and peptide-graphene interactions. Further, through the simulation, followed by experimentation, they proved that the designer peptides can self-assemble on to pristine graphene [248].

Fig. 8
figure 8

Example of supramolecular protein assembly with carbon material A single-wall carbon nanotube, B C60 fullerene, and C pristine graphene (Image reprinted with permission from Ref [240, 243, 248])

5 Conclusions

To date, utilizing protein self-assembly from nature to designed supramolecular interfaces has been widely developed and has been proven to be a powerful tool for various application. Recently, as protein folding prediction using artificial intelligence, such as Alphafold [249] and RoseTTAFold [250], and computational tools have become very sophisticated, it is possible to make folding units with more diverse structures. This enables the assembly of proteins with more complex structures and functions. Folding units with structural flexibility or allosteric properties are among the most fascinating research directions. With these noble folding units, many researchers have designed protein assemblies that have more clinical and therapeutic applications beyond academic curiosity. For example, while maintaining stability in the in vivo environment, the assembly pattern is changed, the drug contained in the structure is delivered into the body, or the ligand related to signal transduction and receptor signal transduction in the body is labeled on the surface of the supramolecule.

Availability of data and materials

The datasets used and analyzed during the current study are available from the corresponding references listed.

References

  1. B.J. Pieters et al., Natural supramolecular protein assemblies. Chem. Soc. Rev 45(1), 24–39 (2016)

    CAS  Google Scholar 

  2. H. Garcia-Seisdedos et al., Proteins evolve on the edge of supramolecular self-assembly. Nature 548(7666), 244–247 (2017)

    CAS  Google Scholar 

  3. J.A. Marsh, S.A. Teichmann, Structure, Dynamics, Assembly, and Evolution of Protein Complexes. Annu. Rev. Biochem. 84(84), 551–575 (2015)

    CAS  Google Scholar 

  4. I.V. Korendovych, W.F. DeGrado, De novo protein design, a retrospective. Q. Rev. Biophys. 53, e3 (2020)

    CAS  Google Scholar 

  5. W.F. DeGrado, L. Regan, S.P. Ho, The design of a four-helix bundle protein. Cold Spring Harb. Symp. Quant. Biol 52, 521–526 (1987)

    CAS  Google Scholar 

  6. L. Regan, W.F. DeGrado, Characterization of a helical protein designed from first principles. Science 241(4868), 976–978 (1988)

    CAS  Google Scholar 

  7. G. Grigoryan, W.F. Degrado, Probing designability via a generalized model of helical bundle geometry. J. Mol. Biol. 405(4), 1079–1100 (2011)

    CAS  Google Scholar 

  8. C.W. Wood, D.N. Woolfson, CCBuilder 2.0: Powerful and accessible coiled-coil modeling. Protein Sci. 27(1), 103–111 (2018)

    CAS  Google Scholar 

  9. R.D. Fraser, T.P. Macrae, A. Miller, The Fourier Transform of the coiled-coil model for alpha-keratin. A correction. Acta. Crystallogr. 18, 1087 (1965)

    CAS  Google Scholar 

  10. K.W. Plaxco, K.T. Simons, D. Baker, Contact order, transition state placement and the refolding rates of single domain proteins. J Mol Biol 277(4), 985–994 (1998)

    CAS  Google Scholar 

  11. E. Marcos et al., De novo design of a non-local beta-sheet protein with high stability and accuracy. Nat. Struct. Mol. Biol. 25(11), 1028–1034 (2018)

    CAS  Google Scholar 

  12. M.H. Hecht, De novo design of beta-sheet proteins. Proc. Natl. Acad. Sci. USA 91(19), 8729–8730 (1994)

    CAS  Google Scholar 

  13. S. Jones, J.M. Thornton, Principles of protein–protein interactions. Proc. Natl. Acad. Sci. USA 93(1), 13–20 (1996)

    CAS  Google Scholar 

  14. B. Lee, F.M. Richards, The interpretation of protein structures: estimation of static accessibility. J. Mol. Biol. 55(3), 379–400 (1971)

    CAS  Google Scholar 

  15. M. Gerstein, J. Tsai, M. Levitt, The volume of atoms on the protein surface: calculated from simulation, using Voronoi polyhedra. J. Mol. Biol. 249(5), 955–966 (1995)

    CAS  Google Scholar 

  16. T.N. Bhat et al., Bound water molecules and conformational stabilization help mediate an antigen-antibody association. Proc Natl Acad Sci U S A 91(3), 1089–1093 (1994)

    CAS  Google Scholar 

  17. F.M. Richards, The interpretation of protein structures: total volume, group volume distributions and packing density. J. Mol. Biol. 82(1), 1–14 (1974)

    CAS  Google Scholar 

  18. Y. Harpaz, M. Gerstein, C. Chothia, Volume changes on protein folding. Structure 2(7), 641–649 (1994)

    CAS  Google Scholar 

  19. M.J. Castro, S. Anderson, Alanine point-mutations in the reactive region of bovine pancreatic trypsin inhibitor: effects on the kinetics and thermodynamics of binding to beta-trypsin and alpha-chymotrypsin. Biochemistry 35(35), 11435–11446 (1996)

    CAS  Google Scholar 

  20. T. Clackson, J.A. Wells, A hot spot of binding energy in a hormone-receptor interface. Science 267(5196), 383–386 (1995)

    CAS  Google Scholar 

  21. J. de Wit et al., Role of leucine-rich repeat proteins in the development and function of neural circuits. Annu. Rev. Cell. Dev. Biol. 27, 697–729 (2011)

    Google Scholar 

  22. G.L. Blatch, M. Lassle, The tetratricopeptide repeat: a structural motif mediating protein-protein interactions. BioEssays 21(11), 932–939 (1999)

    CAS  Google Scholar 

  23. P.V. Lishko et al., The ankyrin repeats of TRPV1 bind multiple ligands and modulate channel sensitivity. Neuron 54(6), 905–918 (2007)

    CAS  Google Scholar 

  24. R. Latorre, C. Zaelzer, S. Brauchi, Structure-functional intimacies of transient receptor potential channels. Q. Rev. Biophys. 42(3), 201–246 (2009)

    CAS  Google Scholar 

  25. C. Choma et al., Asparagine-mediated self-association of a model transmembrane helix. Nat. Struct. Biol. 7(2), 161–166 (2000)

    CAS  Google Scholar 

  26. P.B. Harbury et al., A switch between 2-stranded, 3-stranded and 4-stranded coiled coils in gcn4 leucine-zipper mutants. Science 262(5138), 1401–1407 (1993)

    CAS  Google Scholar 

  27. R. Riek, D.S. Eisenberg, The activities of amyloids from a structural perspective. Nature 539(7628), 227–235 (2016)

    Google Scholar 

  28. A.W. Fitzpatrick et al., Atomic structure and hierarchical assembly of a cross-beta amyloid fibril. Proc. Natl. Acad. Sci. USA 110(14), 5468–5473 (2013)

    CAS  Google Scholar 

  29. Y. Xiao et al., Abeta(1–42) fibril structure illuminates self-recognition and replication of amyloid in Alzheimer’s disease. Nat. Struct. Mol. Biol. 22(6), 499–505 (2015)

    CAS  Google Scholar 

  30. A. Aggeli et al., Responsive gels formed by the spontaneous self-assembly of peptides into polymeric beta-sheet tapes. Nature 386(6622), 259–262 (1997)

    CAS  Google Scholar 

  31. J.P. Schneider et al., Responsive hydrogels from the intramolecular folding and self-assembly of a designed peptide. J. Am. Chem. Soc. 124(50), 15030–15037 (2002)

    CAS  Google Scholar 

  32. A.A. Vorobieva et al., De novo design of transmembrane beta barrels. Science 371(6531), eabc8182 (2021)

    Google Scholar 

  33. H. Sun et al., Hierarchical self-assembly of proteins through rationally designed supramolecular interfaces. Front. Bioeng. Biotechnol. 8, 295 (2020)

    Google Scholar 

  34. W.J. Song, F.A. Tezcan., A designed supramolecular protein assembly with in vivo enzymatic activity. Science 346(6216), 1525–1528 (2014). https://doi.org/10.1126/science.1259680

    Article  CAS  Google Scholar 

  35. E.N. Salgado et al., Metal templated design of protein interfaces. Proc. Natl. Acad. Sci. USA 107(5), 1827–1832 (2010)

    CAS  Google Scholar 

  36. S.J. Lee, S.L. Michel, Structural metal sites in nonclassical zinc finger proteins involved in transcriptional and translational regulation. Acc. Chem. Res 47(8), 2643–2650 (2014)

    CAS  Google Scholar 

  37. J. Miller, A.D. McLachlan, A. Klug, Repetitive zinc-binding domains in the protein transcription factor IIIA from Xenopus oocytes. EMBO J 4(6), 1609–1614 (1985)

    CAS  Google Scholar 

  38. J. Laniado, T.O. Yeates, A complete rule set for designing symmetry combination materials from protein molecules. Proc. Natl. Acad. Sci. USA 117(50), 31817–31823 (2020)

    CAS  Google Scholar 

  39. M.D. Shoulders, R.T. Raines, Collagen structure and stability. Annu. Rev. Biochem. 78, 929–958 (2009)

    CAS  Google Scholar 

  40. B. Brodsky, J.A. Ramshaw, The collagen triple-helix structure. Matrix Biol 15(8–9), 545–554 (1997)

    CAS  Google Scholar 

  41. B. An, Y.S. Lin, B. Brodsky, Collagen interactions: drug design and delivery. Adv. Drug. Deliv. Rev. 97, 69–84 (2016)

    CAS  Google Scholar 

  42. G.N. Ramachandran, G. Kartha, Structure of collagen. Nature 174(4423), 269–270 (1954)

    CAS  Google Scholar 

  43. K. Okuyama et al., Crystal structure of (Gly-Pro-Hyp)(9): implications for the collagen molecular model. Biopolymers 97(8), 607–616 (2012)

    CAS  Google Scholar 

  44. D.S. Eisenberg, M.R. Sawaya, Structural studies of amyloid proteins at the molecular level. Annu. Rev. Biochem. 86, 69–95 (2017)

    CAS  Google Scholar 

  45. B.H. Toyama, J.S. Weissman, Amyloid structure: conformational diversity and consequences. Annu. Rev. Biochem 80, 557–585 (2011)

    CAS  Google Scholar 

  46. L. Gremer et al., Fibril structure of amyloid-beta(1–42) by cryo-electron microscopy. Science 358(6359), 116–119 (2017)

    CAS  Google Scholar 

  47. R. Guerrero-Ferreira et al., Cryo-EM structure of alpha-synuclein fibrils. Elife 7, e36402 (2018)

    Google Scholar 

  48. A.W.P. Fitzpatrick et al., Cryo-EM structures of tau filaments from Alzheimer’s disease. Nature 547(7662), 185–190 (2017)

    CAS  Google Scholar 

  49. R. Dominguez, K.C. Holmes, Actin structure and function. Annu. Rev. Biophys. 40, 169–186 (2011)

    CAS  Google Scholar 

  50. K.C. Holmes et al., Atomic model of the actin filament. Nature 347(6288), 44–49 (1990)

    CAS  Google Scholar 

  51. J. von der Ecken et al., Structure of the F-actin-tropomyosin complex. Nature 519(7541), 114–117 (2015)

    Google Scholar 

  52. P.S. Gural et al., Cryo-EM structures reveal specialization at the myosin VI-actin interface and a mechanism of force sensitivity. Elife 6, e31125 (2017)

    Google Scholar 

  53. S. Kaltofen et al., Computational de novo design of a self-assembling peptide with predefined structure. J. Mol. Biol 427(2), 550–562 (2015)

    CAS  Google Scholar 

  54. Yokoi, H., Kinoshita, T. Strategy for designing self-assembling peptides to prepare transparent nanofiber hydrogel at neutral pH. J. Nanomater. 2012. 2012.

  55. S.A. Potekhin et al., De novo design of fibrils made of short alpha-helical coiled coil peptides. Chem. Biol. 8(11), 1025–1032 (2001)

    CAS  Google Scholar 

  56. D.E. Wagner et al., Toward the development of peptide nanofilaments and nanoropes as smart materials. Proc. Natl. Acad. Sci. U.S.A. 102(36), 12656–12661 (2005)

    CAS  Google Scholar 

  57. T.H. Sharp et al., Cryo-transmission electron microscopy structure of a gigadalton peptide fiber of de novo design. Proc. Natl. Acad. Sci. U.S.A. 109(33), 13266–13271 (2012)

    Google Scholar 

  58. R.V. Rughani et al., Folding, self-assembly, and bulk material properties of a de novo designed three-stranded beta-sheet hydrogel. Biomacromol 10(5), 1295–1304 (2009)

    CAS  Google Scholar 

  59. C. Wang et al., One-dimensional self-assembly of a rational designed beta-structure peptide. Biopolymers 86(1), 23–31 (2007)

    CAS  Google Scholar 

  60. A. Rani et al., Nanoribbon self-assembly and hydrogel formation from an NOctanoyl octapeptide derived from the antiparallel beta-Interface of a protein homotetramer. Acta Biomater 114, 233–243 (2020)

    CAS  Google Scholar 

  61. C.M. Rufo et al., Short peptides self-assemble to produce catalytic amyloids. Nat. Chem. 6(4), 303–309 (2014)

    CAS  Google Scholar 

  62. N. Haspel et al., De novo tubular nanostructure design based on self-assembly of beta-helical protein motifs. Structure 14(7), 1137–1148 (2006)

    CAS  Google Scholar 

  63. K. Nagy-Smith et al., Molecular structure of monomorphic peptide fibrils within a kinetically trapped hydrogel network. Proc. Natl. Acad. Sci. U.S.A. 112(32), 9816–9821 (2015)

    CAS  Google Scholar 

  64. A.R. Cormier et al., Molecular structure of RADA16-I designer self-assembling peptide nanofibers. ACS Nano 7(9), 7562–7572 (2013)

    CAS  Google Scholar 

  65. R.P. Fagan, N.F. Fairweather, Biogenesis and functions of bacterial S-layers. Nat. Rev. Microbiol. 12(3), 211–222 (2014)

    CAS  Google Scholar 

  66. L. Gambelli et al., Architecture and modular assembly of Sulfolobus S-layers revealed by electron cryotomography. Proc. Natl. Acad. Sci. USA. 116(50), 25278–25286 (2019)

    CAS  Google Scholar 

  67. M.A. Arbing et al., Structure of the surface layer of the methanogenic archaean Methanosarcina acetivorans. Proc. Natl. Acad. Sci. U.S.A. 109(29), 11812–11817 (2012)

    CAS  Google Scholar 

  68. E. Baranova et al., SbsB structure and lattice reconstruction unveil Ca2+ triggered S-layer assembly. Nature 487(7405), 119 (2012)

    CAS  Google Scholar 

  69. T.A.M. Bharat et al., Structure of the hexagonal surface layer on Caulobacter crescentus cells. Nat. Microbiol. 2(7), 17059 (2017)

    CAS  Google Scholar 

  70. U.B. Sleytr et al., S-layers: principles and applications. FEMS Microbiol. Rev. 38(5), 823–864 (2014)

    CAS  Google Scholar 

  71. G.D. Bowman, M. O’Donnell, J. Kuriyan, Structural analysis of a eukaryotic sliding DNA clamp-clamp loader complex. Nature 429(6993), 724–730 (2004)

    CAS  Google Scholar 

  72. X.P. Kong et al., Three-dimensional structure of the beta subunit of E. coli DNA polymerase III holoenzyme: a sliding DNA clamp. Cell 69(3), 425–37 (1992)

    CAS  Google Scholar 

  73. J.M. Gulbis et al., Structure of the C-terminal region of p21(WAF1/CIP1) complexed with human PCNA. Cell 87(2), 297–306 (1996)

    CAS  Google Scholar 

  74. M.A. Trakselis, Structural mechanisms of hexameric helicase loading, assembly, and unwinding. F1000 Res 5, F1000 (2016)

    Google Scholar 

  75. S. Bailey, W.K. Eliason, T.A. Steitz, Structure of hexameric DnaB helicase and its complex with a domain of DnaG primase. Science 318(5849), 459–463 (2007)

    CAS  Google Scholar 

  76. M.R. Singleton et al., Crystal structure of T7 gene 4 ring helicase indicates a mechanism for sequential hydrolysis of nucleotides. Cell 101(6), 589–600 (2000)

    CAS  Google Scholar 

  77. M.R. Singleton, M.S. Dillingham, D.B. Wigley, Structure and mechanism of helicases and nucleic acid translocases. Annu. Rev. Biochem. 76, 23–50 (2007)

    CAS  Google Scholar 

  78. R. Kovall, B.W. Matthews, Toroidal structure of lambda-exonuclease. Science 277(5333), 1824–1827 (1997)

    CAS  Google Scholar 

  79. J. Zhang et al., Crystal structure of E. coli RecE protein reveals a toroidal tetramer for processing double-stranded DNA breaks. Structure 17(5), 690–702 (2009)

    Google Scholar 

  80. S. Lemak et al., Toroidal structure and DNA cleavage by the CRISPR-associated [4Fe-4S] cluster containing Cas4 nuclease SSO0001 from Sulfolobus solfataricus. J. Am. Chem. Soc. 135(46), 17476–17487 (2013)

    CAS  Google Scholar 

  81. L.K. Tamm, H. Hong, B.Y. Liang, Folding and assembly of beta-barrel membrane proteins. BBA-Biomembranes 1666(1–2), 250–263 (2004)

    CAS  Google Scholar 

  82. L.Z. Song et al., Structure of staphylococcal alpha-hemolysin, a heptameric transmembrane pore. Science 274(5294), 1859–1866 (1996)

    CAS  Google Scholar 

  83. J.S. Jiang et al., Atomic structure of anthrax protective antigen pore elucidates toxin translocation. Nature 521(7553), 545-U323 (2015)

    CAS  Google Scholar 

  84. S.W. Cowan et al., Crystal-structures explain functional-properties of 2 Escherichia coli porins. nature 358(6389), 727–733 (1992)

    CAS  Google Scholar 

  85. K. Lundquist et al., The assembly of beta-barrel membrane proteins by BAM and SAM. Mol. Microbiol. 115(3), 425–435 (2021)

    CAS  Google Scholar 

  86. R.Y. Tsien, The green fluorescent protein. Annu. Rev. Biochem. 67, 509–544 (1998)

    CAS  Google Scholar 

  87. F. Yang, L.G. Moss, G.N. Phillips, The molecular structure of green fluorescent protein. Nat. Biotechnol. 14(10), 1246–1251 (1996)

    CAS  Google Scholar 

  88. M. Ormo et al., Crystal structure of the Aequorea victoria green fluorescent protein. Science 273(5280), 1392–1395 (1996)

    CAS  Google Scholar 

  89. F. Baneyx, J.F. Matthaei, Self-assembled two-dimensional protein arrays in bionanotechnology: from S-layers to designed lattices. Curr. Opin. Biotechnol. 28, 39–45 (2014)

    CAS  Google Scholar 

  90. Z.B. Chen et al., Self-assembling 2D arrays with de novo protein building blocks. J. Am. Chem. Soc. 141(22), 8891–8895 (2019)

    CAS  Google Scholar 

  91. J.D. Brodin et al., Metal-directed, chemically tunable assembly of one-, two- and three-dimensional crystalline protein arrays. Nat. Chem. 4(5), 375–382 (2012)

    CAS  Google Scholar 

  92. Y. Suzuki et al., Self-assembly of coherently dynamic, auxetic, two-dimensional protein crystals. Nature 533(7603), 369 (2016)

    CAS  Google Scholar 

  93. S. Gonen et al., Design of ordered two-dimensional arrays mediated by noncovalent protein-protein interfaces. Science 348(6241), 1365–1368 (2015)

    CAS  Google Scholar 

  94. A.R. Thomson et al., Computational design of water-soluble alpha-helical barrels. Science 346(6208), 485–488 (2014)

    CAS  Google Scholar 

  95. C. Xu et al., Computational design of transmembrane pores. Nature 585(7823), 129–134 (2020)

    CAS  Google Scholar 

  96. I.V. Korendovych et al., De novo design and molecular assembly of a transmembrane diporphyrin-binding protein complex. J. Am. Chem. Soc. 132(44), 15516–15518 (2010)

    CAS  Google Scholar 

  97. K.R. Mahendran et al., A monodisperse transmembrane alpha-helical peptide barrel. Nat. Chem. 9(5), 411–419 (2017)

    CAS  Google Scholar 

  98. N.R. Zaccai et al., A de novo peptide hexamer with a mutable channel. Nat. Chem. Biol 7(12), 935–941 (2011)

    CAS  Google Scholar 

  99. M. Yamagami, T. Sawada, M. Fujita, Synthetic beta-barrel by metal-induced folding and assembly. J. Am. Chem. Soc. 140(28), 8644–8647 (2018)

    CAS  Google Scholar 

  100. I.S. Park et al., Designer nanorings with functional cavities from self-assembling beta-sheet peptides. Chem. Asian J. 6(2), 452–458 (2011)

    CAS  Google Scholar 

  101. G. Bhardwaj et al., Accurate de novo design of hyperstable constrained peptides. Nature 538(7625), 329–335 (2016)

    CAS  Google Scholar 

  102. E.G. Baker et al., Miniprotein design: past, present, and prospects. Acc. Chem. Res. 50(9), 2085–2092 (2017)

    CAS  Google Scholar 

  103. Y.B. Lim, K.S. Moon, M. Lee, Stabilization of an alpha helix by beta-sheet-mediated self-assembly of a macrocyclic peptide. Angew. Chem. Int. Ed. Engl. 48(9), 1601–1605 (2009)

    CAS  Google Scholar 

  104. K. Namba, G. Stubbs, Structure of tobacco mosaic virus at 3.6 A resolution: implications for assembly. Science 231(4744), 1401–6 (1986)

    CAS  Google Scholar 

  105. G.P. Lomonossoff, C. Wege, TMV particles: the journey from fundamental studies to bionanotechnology applications. Adv. Virus Res. 102, 149–176 (2018)

    CAS  Google Scholar 

  106. H. Wang, J.N. Culver, G. Stubbs, Structure of ribgrass mosaic virus at 2.9 A resolution: evolution and taxonomy of tobamoviruses. J. Mol. Biol. 269(5), 769–79 (1997)

    CAS  Google Scholar 

  107. D.K. Clare et al., Novel inter-subunit contacts in barley stripe mosaic virus revealed by cryo-electron microscopy. Structure 23(10), 1815–1826 (2015)

    CAS  Google Scholar 

  108. H.V. Goodson, E.M. Jonasson, Microtubules and microtubule-associated proteins. Cold Spring Harb. Perspect. Biol. 10(6), a022608 (2018)

    Google Scholar 

  109. E. Mandelkow, E.M. Mandelkow, Microtubules and microtubule-associated proteins. Curr. Opin. Cell. Biol. 7(1), 72–81 (1995)

    CAS  Google Scholar 

  110. R.F. Garmann et al., Physical principles in the self-assembly of a simple spherical virus. Acc. Chem. Res. 49(1), 48–55 (2016)

    CAS  Google Scholar 

  111. J.A. Speir et al., Structures of the native and swollen forms of cowpea chlorotic mottle virus determined by X-ray crystallography and cryo-electron microscopy. Structure 3(1), 63–78 (1995)

    CAS  Google Scholar 

  112. M.R. Perkett, D.T. Mirijanian, M.F. Hagan, The allosteric switching mechanism in bacteriophage MS2. J. Chem. Phys. 145(3), 035101 (2016)

    Google Scholar 

  113. S.C. Harrison et al., Tomato bushy stunt virus at 2.9 A resolution. Nature 276(5686), 368–73 (1978)

    CAS  Google Scholar 

  114. Y. Zhang, B.P. Orner, Self-assembly in the ferritin nano-cage protein superfamily. Int. J. Mol. Sci. 12(8), 5406–5421 (2011)

    CAS  Google Scholar 

  115. L.E. Bevers, E.C. Theil, Maxi- and mini-ferritins: minerals and protein nanocages. Prog Mol Subcell Biol 52, 29–47 (2011)

    CAS  Google Scholar 

  116. X.F. Liu, E.C. Theil, Ferritins: Dynamic management of biological iron and oxygen chemistry. Acc. Chem. Res. 38(3), 167–175 (2005)

    CAS  Google Scholar 

  117. R.A. Grant et al., The crystal structure of Dps, a ferritin homolog that binds and protects DNA. Nat. Struct. Biol. 5(4), 294–303 (1998)

    CAS  Google Scholar 

  118. Y. Ha et al., Crystal structure of bullfrog M ferritin at 2.8 angstrom resolution: analysis of subunit interactions and the binuclear metal center. J. Biol. Inorg. Chem. 4(3), 243–256 (1999)

    CAS  Google Scholar 

  119. A. Dautant et al., Structure of a monoclinic crystal form of cytochrome b1 (bacterioferritin) from E. coli. Acta Crystallogr. Sect. D Struct. Biol. 54, 16–24 (1998)

    CAS  Google Scholar 

  120. H.S. Huan et al., Structure determination of ferritin from Dendrorhynchus zhejiangensis. Biochem. Biophys. Res. Commun. 531(2), 195–202 (2020)

    CAS  Google Scholar 

  121. R. Ladenstein, M. Fischer, A. Bacher, The lumazine synthase/riboflavin synthase complex: shapes and functions of a highly variable enzyme system. FEBS J. 280(11), 2537–2563 (2013)

    CAS  Google Scholar 

  122. E. Morgunova et al., Crystal structure of lumazine synthase from Mycobacterium tuberculosis as a target for rational drug design: Binding mode of a new class of purinetrione inhibitors. Biochemistry 44(8), 2746–2758 (2005)

    CAS  Google Scholar 

  123. A. Casanas et al., Vault particles: a new generation of delivery nanodevices. Curr. Opin. Biotechnol. 23(6), 972–977 (2012)

    CAS  Google Scholar 

  124. L.B. Kong et al., Structure of the vault, a ubiquitous celular component. Structure 7(4), 371–379 (1999)

    CAS  Google Scholar 

  125. G. Frascotti et al., The vault nanoparticle: a gigantic ribonucleoprotein assembly involved in diverse physiological and pathological phenomena and an ideal nanovector for drug delivery and therapy. Cancers 13(4), 707 (2021)

    CAS  Google Scholar 

  126. Y. Mikyas et al., Cryoelectron microscopy imaging of recombinant and tissue derived vaults: localization of the MVP N termini and VPARP. J. Mol. Biol. 344(1), 91–105 (2004)

    CAS  Google Scholar 

  127. A. Fotin et al., Molecular model for a complete clathrin lattice from electron cryomicroscopy. Nature 432(7017), 573–579 (2004)

    CAS  Google Scholar 

  128. C.M. Kumar, S.C. Mande, G. Mahajan, Multiple chaperonins in bacteria–novel functions and non-canonical behaviors. Cell Stress Chaperones 20(4), 555–574 (2015)

    CAS  Google Scholar 

  129. X. Fei et al., Formation and structures of GroEL:GroES2 chaperonin footballs, the protein-folding functional form. Proc. Natl. Acad. Sci. USA. 111(35), 12775–12780 (2014)

    CAS  Google Scholar 

  130. A. Leitner et al., The molecular architecture of the eukaryotic chaperonin TRiC/CCT. Structure 20(5), 814–825 (2012)

    CAS  Google Scholar 

  131. H. Shen et al., De novo design of self-assembling helical protein filaments. Science 362(6415), 705 (2018)

    CAS  Google Scholar 

  132. S.A. Hughes et al., Ambidextrous helical nanotubes from self-assembly of designed helical hairpin motifs. Proc. Natl. Acad. Sci. U.S.A. 116(29), 14456–14464 (2019)

    CAS  Google Scholar 

  133. N.C. Burgess et al., Modular design of self-assembling peptide-based nanotubes. J. Am. Chem. Soc. 137(33), 10554–10562 (2015)

    CAS  Google Scholar 

  134. K.H. Chen et al., X-ray Crystallographic Structure of a Giant Double-Walled Peptide Nanotube Formed by a Macrocyclic beta-Sheet Containing A beta(16–22). J. Am. Chem. Soc. 139(24), 8102–8105 (2017)

    CAS  Google Scholar 

  135. R. Divine et al., Designed proteins assemble antibodies into modular nanocages. Science 372(6537), 47 (2021)

    Google Scholar 

  136. Y. Hsia et al., Design of multi-scale protein complexes by hierarchical building block fusion. Nat. Commun. 12(1), 2294 (2021)

    CAS  Google Scholar 

  137. J.B. Bale et al., Accurate design of megadalton-scale two-component icosahedral protein complexes. Science 353(6297), 389–394 (2016)

    CAS  Google Scholar 

  138. N.P. King et al., Accurate design of co-assembling multi-component protein nanomaterials. Nature 510(7503), 103 (2014)

    CAS  Google Scholar 

  139. Y.T. Lai et al., Structure of a designed protein cage that self-assembles into a highly porous cube. Nat. Chem. 6(12), 1065–1071 (2014)

    CAS  Google Scholar 

  140. Y. Hsia et al., Design of a hyperstable 60-subunit protein icosahedron. Nature 535(7610), 136 (2016)

    CAS  Google Scholar 

  141. A.J. Wargacki et al., Complete and cooperative in vitro assembly of computationally designed self-assembling protein nanomaterials. Nat. Commun. 12(1), 883 (2021)

    CAS  Google Scholar 

  142. N.P. King et al., Computational design of self-assembling protein nanomaterials with atomic level accuracy. Science 336(6085), 1171–1174 (2012)

    CAS  Google Scholar 

  143. G. Marsh, Symmetry-directed self-assembly of a tetrahedral protein cage mediated by de novo-designed coiled coils. ChemBioChem 18(19), 1871–1871 (2017)

    Google Scholar 

  144. Y. Hsia et al., Design of a hyperstable 60-subunit protein icosahedron (vol 535, pg 136, 2016). Nature 540(7631), 150 (2016)

    CAS  Google Scholar 

  145. E. Ostuni et al., Using self-assembled monolayers to pattern ECM proteins and cells on substrates. Methods Mol. Biol. 522, 183–194 (2009)

    CAS  Google Scholar 

  146. G.A. Hudalla, W.L. Murphy, Chemically well-defined self-assembled monolayers for cell culture: toward mimicking the natural ECM. Soft Matter 7(20), 9561–9571 (2011)

    CAS  Google Scholar 

  147. J.K. Mouw, G. Ou, V.M. Weaver, Extracellular matrix assembly: a multiscale deconstruction. Nat. Rev. Mol. Cell. Biol 15(12), 771–785 (2014)

    CAS  Google Scholar 

  148. K.S. Hellmund, B. Koksch, Self-assembling peptides as extracellular matrix mimics to influence stem cell’s fate. Front. Chem. 7, 172 (2019)

    CAS  Google Scholar 

  149. E.C. Wu, S.G. Zhang, C.A.E. Hauser, Self-assembling peptides as cell-interactive scaffolds. Adv. Func. Mater. 22(3), 456–468 (2012)

    CAS  Google Scholar 

  150. Y.L. Yang et al., Designer self-assembling peptide nanomaterials. Nano Today 4(2), 193–210 (2009)

    CAS  Google Scholar 

  151. X. Liu et al., Functionalized self-assembling peptide nanofiber hydrogels mimic stem cell niche to control human adipose stem cell behavior in vitro. Acta Biomater. 9(6), 6798–6805 (2013)

    CAS  Google Scholar 

  152. R.S. Jacob et al., Self healing hydrogels composed of amyloid nano fibrils for cell culture and stem cell differentiation. Biomaterials 54, 97–105 (2015)

    CAS  Google Scholar 

  153. K.F. Bruggeman et al., Temporally controlled release of multiple growth factors from a self-assembling peptide hydrogel. Nanotechnology 27(38), 385102 (2016)

    Google Scholar 

  154. A.L. Rodriguez et al., Using minimalist self-assembling peptides as hierarchical scaffolds to stabilise growth factors and promote stem cell integration in the injured brain. J. Tissue Eng. Regen. Med. 12(3), E1571–E1579 (2018)

    CAS  Google Scholar 

  155. K.M. Hennessy et al., The effect of collagen I mimetic peptides on mesenchymal stem cell adhesion and differentiation, and on bone formation at hydroxyapatite surfaces. Biomaterials 30(10), 1898–1909 (2009)

    CAS  Google Scholar 

  156. X.H. Zhang et al., Peptide-conjugated hyaluronic acid surface for the culture of human induced pluripotent stem cells under defined conditions. Carbohyd. Polym. 136, 1061–1064 (2016)

    CAS  Google Scholar 

  157. M. Yamada et al., Ile-Lys-Val-Ala-Val (IKVAV)-containing laminin alpha 1 chain peptides form amyloid-like fibrils. FEBS Lett. 530(1–3), 48–52 (2002)

    CAS  Google Scholar 

  158. F.M. Watt, W.T.S. Huck, Role of the extracellular matrix in regulating stem cell fate. Nat. Rev. Mol. Cell Biol. 14(8), 467–473 (2013)

    CAS  Google Scholar 

  159. K.J. Lampe, S.C. Heilshorn, Building stem cell niches from the molecule up through engineered peptide materials. Neurosci. Lett. 519(2), 138–146 (2012)

    CAS  Google Scholar 

  160. U. Johansson et al., Assembly of functionalized silk together with cells to obtain proliferative 3D cultures integrated in a network of ECM-like microfibers. Sci. Rep. 9, 6291 (2019)

    Google Scholar 

  161. S. Zhang, Lipid-like self-assembling peptides. Acc Chem Res 45(12), 2142–2150 (2012)

    CAS  Google Scholar 

  162. N. Nuraje, H.Y. Bai, K. Su, Bolaamphiphilic molecules: assembly and applications. Prog. Polym. Sci. 38(2), 302–343 (2013)

    CAS  Google Scholar 

  163. N. Schmidt et al., Arginine-rich cell-penetrating peptides. FEBS Lett. 584(9), 1806–1813 (2010)

    CAS  Google Scholar 

  164. E.R. da Silva et al., Self-assembled arginine-capped peptide bolaamphiphile nanosheets for cell culture and controlled wettability surfaces. Biomacromol 16(10), 3180–3190 (2015)

    Google Scholar 

  165. S.P. Massia, S.S. Rao, J.A. Hubbell, Covalently immobilized laminin peptide Tyr-Ile-Gly-Ser-Arg (Yigsr) supports cell spreading and colocalization of the 67-kiloDalton laminin receptor with alpha-actinin and vinculin. J. Biol. Chem. 268(11), 8053–8059 (1993)

    CAS  Google Scholar 

  166. K.L. Niece et al., Self-assembly combining two bioactive peptide-amphiphile molecules into nanofibers by electrostatic attraction. J. Am. Chem. Soc. 125(24), 7146–7147 (2003)

    CAS  Google Scholar 

  167. R. Jain, S. Roy, Tuning the gelation behavior of short laminin derived peptides via solvent mediated self-assembly. Mater. Sci. Eng. C Mater. Biol. Appl. 108, 110483 (2020)

    CAS  Google Scholar 

  168. R. Jain, S. Roy, Controlling neuronal cell growth through composite laminin supramolecular hydrogels. ACS Biomater. Sci. Eng. 6(5), 2832–2846 (2020)

    CAS  Google Scholar 

  169. M. Zhou et al., Self-assembled peptide-based hydrogels as scaffolds for anchorage-dependent cells. Biomaterials 30(13), 2523–2530 (2009)

    CAS  Google Scholar 

  170. C.C. Horgan et al., Characterisation of minimalist co-assembled fluorenylmethyloxycarbonyl self-assembling peptide systems for presentation of multiple bioactive peptides. Acta Biomater. 38, 11–22 (2016)

    CAS  Google Scholar 

  171. S.S.S. Aye et al., Scaffolds formed via the non-equilibrium supramolecular assembly of the synergistic ecm peptides RGD and PHSRN demonstrate improved cell attachment in 3D. Polymers 10(7), 690 (2018)

    Google Scholar 

  172. M. Gupta et al., Self-assembly of a dipeptide-containing conformationally restricted dehydrophenylalanine residue to form ordered nanotubes. Adv. Mater. 19(6), 858 (2007)

    CAS  Google Scholar 

  173. J.J. Panda, V.S. Chauhan, Short peptide based self-assembled nanostructures: implications in drug delivery and tissue engineering. Polym. Chem. 5(15), 4418–4436 (2014)

    Google Scholar 

  174. C.K. Thota, N. Yadav, V.S. Chauhan, A novel highly stable and injectable hydrogel based on a conformationally restricted ultrashort peptide. Sci. Rep. 6, 31167 (2016)

    CAS  Google Scholar 

  175. N. Yadav, M.K. Chauhan, V.S. Chauhan, Short to ultrashort peptide-based hydrogels as a platform for biomedical applications. Biomater. Sci. 8(1), 84–100 (2020)

    CAS  Google Scholar 

  176. N. Yadav, M.K. Chauhan, V.S. Chauhan, Conformationally constrained dipeptide-based hydrogel as a platform for 3D cell growth and tissue engineering applications. Appl. Nanosci. 11(7), 2019–2031 (2021)

    CAS  Google Scholar 

  177. J.M. Perez-Aguilar, J.G. Saven, Computational design of membrane proteins. Structure 20(1), 5–14 (2012)

    CAS  Google Scholar 

  178. J.D. Lear, Z.R. Wasserman, W.F. DeGrado, Synthetic amphiphilic peptide models for protein ion channels. Science 240(4856), 1177–1181 (1988)

    CAS  Google Scholar 

  179. M.M. Mohammad, K.R. Howard, L. Movileanu, Redesign of a plugged beta-barrel membrane protein. J. Biol. Chem. 286(10), 8000–8013 (2011)

    CAS  Google Scholar 

  180. M. Soskine et al., An engineered ClyA nanopore detects folded target proteins by selective external association and pore entry. Nano Lett. 12(9), 4895–4900 (2012)

    CAS  Google Scholar 

  181. L. Franceschini et al., A nanopore machine promotes the vectorial transport of DNA across membranes. Nat. Commun. 4, 2415 (2013)

    Google Scholar 

  182. M. Soskine et al., Tuning the size and properties of ClyA nanopores assisted by directed evolution. J. Am. Chem. Soc. 135(36), 13456–13463 (2013)

    CAS  Google Scholar 

  183. N.H. Joh et al., De novo design of a transmembrane Zn(2)(+)-transporting four-helix bundle. Science 346(6216), 1520–1524 (2014)

    CAS  Google Scholar 

  184. B. Nguyen, N.H. Tolia, Protein-based antigen presentation platforms for nanoparticle vaccines. NPJ Vaccines 6(1), 70 (2021)

    CAS  Google Scholar 

  185. J. Lopez-Sagaseta et al., Self-assembling protein nanoparticles in the design of vaccines. Comput. Struct. Biotechnol. J. 14, 58–68 (2016)

    CAS  Google Scholar 

  186. J. Min et al., Lumazine synthase protein cage nanoparticles as modular delivery platforms for targeted drug delivery. RSC Adv. 4(89), 48596–48600 (2014)

    CAS  Google Scholar 

  187. B.S. Zhang et al., A platform incorporating trimeric antigens into self-assembling nanoparticles reveals SARS-CoV-2-spike nanoparticles to elicit substantially higher neutralizing responses than spike alone. Sci. Rep. 10(1), 18149 (2020)

    CAS  Google Scholar 

  188. A.E. Powell et al., A single immunization with spike-functionalized ferritin vaccines elicits neutralizing antibody responses against SARS-CoV-2 in mice. bioRxiv (2020). https://doi.org/10.1101/2020.08.28.272518

    Article  Google Scholar 

  189. K. Sliepen et al., Presenting native-like HIV-1 envelope trimers on ferritin nanoparticles improves their immunogenicity. Retrovirology 12, 82 (2015)

    Google Scholar 

  190. R. Divine et al., Designed proteins assemble antibodies into modular nanocages. Science 372(6537), eabd9994 (2021)

    CAS  Google Scholar 

  191. Z. Yang et al., Encapsulation of platinum anticancer drugs by apoferritin. Chem. Commun. 33, 3453–3455 (2007)

    Google Scholar 

  192. A.H. Ma-Ham et al., Apoferritin-based nanomedicine platform for drug delivery: equilibrium binding study of daunomycin with DNA. J. Mater. Chem. 21(24), 8700–8708 (2011)

    CAS  Google Scholar 

  193. M. Kanekiyo et al., Self-assembling influenza nanoparticle vaccines elicit broadly neutralizing H1N1 antibodies. Nature 499(7456), 102–106 (2013)

    CAS  Google Scholar 

  194. J. Jardine et al., Rational HIV immunogen design to target specific germline B cell receptors. Science 340(6133), 711–716 (2013)

    CAS  Google Scholar 

  195. M. Kanekiyo et al., Rational design of an epstein-barr virus vaccine targeting the receptor-binding site. Cell 162(5), 1090–1100 (2015)

    CAS  Google Scholar 

  196. A. Antanasijevic et al., Structural and functional evaluation of de novo-designed, two-component nanoparticle carriers for HIV Env trimer immunogens. PLoS Pathog 16(8), e1008665 (2020)

    CAS  Google Scholar 

  197. B. Tesarova et al., Taking advantage of cellular uptake of ferritin nanocages for targeted drug delivery. J. Control. Release 325, 176–190 (2020)

    CAS  Google Scholar 

  198. B. Worsdorfer, Z. Pianowski, D. Hilvert, Efficient in vitro encapsulation of protein cargo by an engineered protein container. J. Am. Chem. Soc. 134(2), 909–911 (2012)

    CAS  Google Scholar 

  199. B. Worsdorfer, K.J. Woycechowsky, D. Hilvert, Directed evolution of a protein container. Science 331(6017), 589–592 (2011)

    Google Scholar 

  200. T. Yamada et al., Nanoparticles for the delivery of genes and drugs to human hepatocytes. Nat Biotechnol 21(8), 885–890 (2003)

    CAS  Google Scholar 

  201. M. Uchida et al., Targeting of cancer cells with ferrimagnetic ferritin cage nanoparticles. J. Am. Chem. Soc. 128(51), 16626–16633 (2006)

    CAS  Google Scholar 

  202. Y. Lei et al., Targeted tumor delivery and controlled release of neuronal drugs with ferritin nanoparticles to regulate pancreatic cancer progression. J. Control Release 232, 131–142 (2016)

    CAS  Google Scholar 

  203. A. Antanasijevic et al., Structural and functional evaluation ofde novo-designed, two-component nanoparticle carriers for HIV Env trimer immunogens. Plos Pathogens 16(8), e1008665 (2020)

    CAS  Google Scholar 

  204. G. Ueda et al., Tailored design of protein nanoparticle scaffolds for multivalent presentation of viral glycoprotein antigens. Elife 9, e57659 (2020)

    CAS  Google Scholar 

  205. H.C. Sun et al., Hierarchical self-assembly of proteins through rationally designed supramolecular interfaces. Front. Bioeng. Biotechnol. 8, 295 (2020)

    Google Scholar 

  206. C.H. Yang et al., Dual fluorescent- and isotopic-labelled self-assembling vancomycin for in vivo imaging of bacterial infections. Angew. Chem. Int. Ed. 56(9), 2356–2360 (2017)

    CAS  Google Scholar 

  207. Q. Cai et al., Chemotaxis-instructed intracellular Staphylococcus aureus infection detection by a targeting and self-assembly signal-enhanced photoacoustic probe. Nano Lett. 18(10), 6229–6236 (2018)

    CAS  Google Scholar 

  208. C.H. Yang et al., Combating bacterial infection by in situ self-assembly of AIEgen-peptide conjugate. Biomaterials 244, 119972 (2020)

    CAS  Google Scholar 

  209. R. de la Rica et al., Peptide-nanotube biochips for label-free detection of multiple pathogens. Small 6(10), 1092–1095 (2010)

    Google Scholar 

  210. J.H. Kim, J. Ryu, C.B. Park, selective detection of neurotoxin by photoluminescent peptide nanotubes. Small 7(6), 718–722 (2011)

    CAS  Google Scholar 

  211. Z.Y. Qian et al., Improved enzyme immobilization on an ionic-complementary peptide-modified electrode for biomolecular sensing. Langmuir 26(3), 2176–2180 (2010)

    CAS  Google Scholar 

  212. R.C. Bianchi et al., A nonenzymatic biosensor based on gold electrodes modified with peptide self-assemblies for detecting ammonia and urea oxidation. Langmuir 30(38), 11464–11473 (2014)

    CAS  Google Scholar 

  213. J.P. Zhang, R.J. Narayan, DNA-directed self-assembly of fluorescent dye-labeled streptavidin arrays for protein detection. J. Nanosci. Nanotechnol. 8(11), 6048–6051 (2008)

    CAS  Google Scholar 

  214. M. Arai et al., Protein sensing device with multi-recognition ability composed of self-organized glycopeptide bundle. Int. J. Mol. Sci. 22(1), 366 (2021)

    CAS  Google Scholar 

  215. Y.X. Liu et al., Near-atomic cryo-EM imaging of a small protein displayed on a designed scaffolding system. Proc. Natl. Acad. Sci. USA. 115(13), 3362–3367 (2018)

    CAS  Google Scholar 

  216. J. Ryu, C.B. Park, High-temperature self-assembly of peptides into vertically well-aligned nanowires by aniline vapor. Adv. Mater. 20(19), 3754 (2008)

    CAS  Google Scholar 

  217. T. Cipriano et al., Bioinspired peptide nanostructures for organic field-effect transistors. ACS Appl. Mater. Interfaces. 6(23), 21408–21415 (2014)

    CAS  Google Scholar 

  218. S. Reuveni, M. Ehrenberg, J. Paulsson, Ribosomes are optimized for autocatalytic production. Nature 547(7663), 293 (2017)

    CAS  Google Scholar 

  219. C.Q. Zhang et al., Self-assembled peptide nanofibers designed as biological enzymes for catalyzing ester hydrolysis. ACS Nano 8(11), 11715–11723 (2014)

    CAS  Google Scholar 

  220. A. Medina-Morales et al., In vitro and cellular self-assembly of a Zn-binding protein cryptand via templated disulfide bonds. J. Am. Chem. Soc. 135(32), 12013–12022 (2013)

    CAS  Google Scholar 

  221. J.B. Bailey et al., Metal-directed design of supramolecular protein assemblies. Pept. Protein Enzyme Des. 580, 223–250 (2016)

    CAS  Google Scholar 

  222. O. Zozulia, M.A. Dolan, I.V. Korendovych, Catalytic peptide assemblies. Chem. Soc. Rev. 47(10), 3621–3639 (2018)

    CAS  Google Scholar 

  223. W.J. Song, F.A. Tezcan, A designed supramolecular protein assembly with in vivo enzymatic activity. Science 346(6216), 1525–1528 (2014)

    CAS  Google Scholar 

  224. O.V. Makhlynets, P.M. Gosavi, I.V. Korendovych, Short self-assembling peptides are able to bind to copper and activate oxygen. Angewandte Chemie-International Edition 55(31), 9017–9020 (2016)

    CAS  Google Scholar 

  225. K.R. Lee et al., Enhancement of catalytic activity of a programmed gold nanoparticle superstructure modulated by supramolecular protein assembly. Catal. Today 295, 95–101 (2017)

    CAS  Google Scholar 

  226. Y.H. Ting et al., Zinc(II)-histidine induced collagen peptide assemblies: morphology modulation and hydrolytic catalysis evaluation. Biomacromol 19(7), 2629–2637 (2018)

    CAS  Google Scholar 

  227. H.J. Kim et al., Biological conversion of methane to methanol through genetic reassembly of native catalytic domains. Nat. Catal. 2(4), 342–353 (2019)

    CAS  Google Scholar 

  228. B. Rubinov et al., Transient fibril structures facilitating nonenzymatic self-replication. ACS Nano 6(9), 7893–7901 (2012)

    CAS  Google Scholar 

  229. H. Kaur et al., Protein morphology drives the structure and catalytic activity of bio-inorganic hybrids. Int. J. Biol. Macromol 176, 106–116 (2021)

    CAS  Google Scholar 

  230. Y.T. Wang et al., Bioinspired phosphatase-like mimic built from the self-assembly of de novo designed helical short peptides. ACS Catal. 11(9), 5839–5849 (2021)

    CAS  Google Scholar 

  231. B.M. Soares et al., Chiral organocatalysts based on lipopeptide micelles for aldol reactions in water. Phys. Chem. Chem. Phys. 19(2), 1181–1189 (2017)

    CAS  Google Scholar 

  232. D. Shan et al., Self-assembled films of hemoglobin/laponite/chitosan: application for the direct electrochemistry and catalysis to hydrogen peroxide. Biomacromol 8(10), 3041–3046 (2007)

    CAS  Google Scholar 

  233. H. Yang et al., Modification of hydrophilic and hydrophobic surfaces using an ionic-complementary peptide. Plos ONE 2(12), e1325 (2007)

    Google Scholar 

  234. C.L. Brown et al., Template-directed assembly of a de novo designed protein. J. Am. Chem. Soc. 124(24), 6846–6848 (2002)

    CAS  Google Scholar 

  235. X.B. Mao et al., Sequence effects on peptide assembly characteristics observed by using scanning tunneling microscopy. J. Am. Chem. Soc. 135(6), 2181–2187 (2013)

    CAS  Google Scholar 

  236. W.P. Lv et al., Interlayer water regulates the bio-nano interface of a beta-sheet protein stacking on graphene. Sci. Rep. 5, 7572 (2015)

    CAS  Google Scholar 

  237. J.J. Guo et al., The adsorption mechanism and induced conformational changes of three typical proteins with different secondary structural features on graphene. RSC Adv. 4(20), 9953–9962 (2014)

    CAS  Google Scholar 

  238. C.R. So et al., Controlling self-assembly of engineered peptides on graphite by rational mutation. ACS Nano 6(2), 1648–1656 (2012)

    CAS  Google Scholar 

  239. D. Khatayevich et al., Controlling the surface chemistry of graphite by engineered self-assembled peptides. Langmuir 28(23), 8589–8593 (2012)

    CAS  Google Scholar 

  240. G. Grigoryan et al., Computational design of virus-like protein assemblies on carbon nanotube surfaces. Science 332(6033), 1071–1076 (2011)

    CAS  Google Scholar 

  241. Y.S. Ko et al., Peptide-based bimetallic nanostructures with tailored surface compositions and their oxygen electroreduction activities. CrystEngComm 18(32), 6024–6028 (2016)

    CAS  Google Scholar 

  242. G. Thakur, K. Prashanthi, T. Thundat, Directed self-assembly of proteins into discrete radial patterns. Sci. Rep. 3, 1923 (2013)

    Google Scholar 

  243. K.H. Kim et al., Protein-directed self-assembly of a fullerene crystal. Nat. Commun. 7, 11429 (2016)

    CAS  Google Scholar 

  244. J.D. Hartgerink et al., Self-assembling peptide nanotubes. J. Am. Chem. Soc. 118(1), 43–50 (1996)

    CAS  Google Scholar 

  245. J.Y. Rho et al., Dual self-assembly of supramolecular peptide nanotubes to provide stabilisation in water. Nat. Commun. 10, 4708 (2019)

    Google Scholar 

  246. B. Dinesh et al., Self-assembly of diphenylalanine backbone homologues and their combination with functionalized carbon nanotubes. Nanoscale 7(38), 15873–15879 (2015)

    CAS  Google Scholar 

  247. G.M. Mustata et al., Graphene symmetry amplified by designed peptide self-assembly. Biophys. J . 110(11), 2507–2516 (2016)

    CAS  Google Scholar 

  248. Y.H. No et al., Nature-inspired construction of two-dimensionally self-assembled peptide on pristine graphene. J. Phys. Chem. Lett. 8(16), 3734–3739 (2017)

    CAS  Google Scholar 

  249. J. Jumper et al., Highly accurate protein structure prediction with AlphaFold. Nature 596(7873), 583–589 (2021)

    CAS  Google Scholar 

  250. M. Baek et al., Accurate prediction of protein structures and interactions using a three-track neural network. Science 373(6557), 871–876 (2021)

    CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by NRF-2019R1A2C1085712, KIST-2E30641-20-150, and the Nano Convergence Industrial Strategic Technology Development program (No. 200000105).

Author information

Authors and Affiliations

Authors

Contributions

NHK, HC, and ZMS contributed to this manuscript equally in writing the manuscript. HK, JL, and Prof. HC have participated in searching references and writing the manuscript. Prof. YHK supervised this work. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Yong Ho Kim.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, N.H., Choi, H., Shahzad, Z.M. et al. Supramolecular assembly of protein building blocks: from folding to function. Nano Convergence 9, 4 (2022). https://doi.org/10.1186/s40580-021-00294-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40580-021-00294-3

Keywords