Next Article in Journal
A Clinical Study of Urine Amino Acids in Children with Autism Spectrum Disorder
Next Article in Special Issue
Targeted Liposomal Drug Delivery: Overview of the Current Applications and Challenges
Previous Article in Journal
What Is New about the Semimembranosus Distal Tendon? Ultrasound, Anatomical, and Histological Study with Clinical and Therapeutic Application
Previous Article in Special Issue
Approaches to Characterize and Quantify Extracellular Vesicle Surface Conjugation Efficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Role of Natural Binding Proteins in Therapy and Diagnostics

by
Marco Eigenfeld
1,
Kilian F. M. Lupp
1 and
Sebastian P. Schwaminger
1,2,*
1
Otto-Loewi Research Center, Division of Medicinal Chemistry, Medical University of Graz, Neue Stiftingtalstraße 6, 8010 Graz, Austria
2
BioTechMed-Graz, Mozartgasse 12/II, 8010 Graz, Austria
*
Author to whom correspondence should be addressed.
Life 2024, 14(5), 630; https://doi.org/10.3390/life14050630
Submission received: 31 March 2024 / Revised: 2 May 2024 / Accepted: 8 May 2024 / Published: 15 May 2024

Abstract

:
This review systematically investigates the critical role of natural binding proteins (NBPs), encompassing DNA-, RNA-, carbohydrate-, fatty acid-, and chitin-binding proteins, in the realms of oncology and diagnostics. In an era where cancer continues to pose significant challenges to healthcare systems worldwide, the innovative exploration of NBPs offers a promising frontier for advancing both the diagnostic accuracy and therapeutic efficacy of cancer management strategies. This manuscript provides an in-depth examination of the unique mechanisms by which NBPs interact with specific molecular targets, highlighting their potential to revolutionize cancer diagnostics and therapy. Furthermore, it discusses the burgeoning research on aptamers, demonstrating their utility as ‘nucleic acid antibodies’ for targeted therapy and precision diagnostics. Despite the promising applications of NBPs and aptamers in enhancing early cancer detection and developing personalized treatment protocols, this review identifies a critical knowledge gap: the need for comprehensive studies to understand the diverse functionalities and therapeutic potentials of NBPs across different cancer types and diagnostic scenarios. By bridging this gap, this manuscript underscores the importance of NBPs and aptamers in paving the way for next-generation diagnostics and targeted cancer treatments.

1. Introduction

The landscape of cancer in the United States in 2023 depicts a daunting scenario, marked by an estimated 1,958,310 new cancer cases and 609,820 cancer-related deaths. For instance, prostate cancer among men has increased with an annual growth of 3% since 2014, accounting for about 99,000 new cases each year [1]. However, breast cancer emerges as the globally most prevalent cancer among women, showing an ongoing upward trend [2,3]. The effective treatment of breast cancer is influenced by several factors, including the stage of the disease, tumor aggressiveness, individual response to treatment, and lifestyle choices like medication, smoking, or alcohol consumption [3]. Despite these challenges, the rate of lung cancer has been declining more slowly in women than in men, with a yearly decrease of 1.1% in women compared to 2.6% in men between 2015 and 2019 [1]. However, the continuous decline in the cancer death rate, including a 1.5% reduction from 2019 to 2020, contributes to a 33% overall reduction since 1991, highlighting the significant impact of advancements in treatment across various cancer types.
By testing 1000 women who are 50 years old every year for a decade, it is possible to prevent one death from breast cancer [4]. This underscores the potential life-saving benefits of early detection in managing breast cancer.
Traditional long-term cancer treatments, such as chemotherapy, are associated with significant side effects, including hair loss, skin issues, hot flashes, and nausea [5,6]. They can, furthermore, include fertility issues and heart and lung problems [5,6]. Breast surgery as a consequence of breast cancer treatment and mastectomy have physical and psychological impacts, such as changes in body image and loss of breast sensitivity [6]. Additionally, some breast cancers may develop drug resistance, necessitating alternative treatment approaches [5].
However, significant progress has been made recently in developing combination therapies for breast cancer. These include monoclonal antibodies (mAbs) like trastuzumab, pertuzumab, and margetuximab, combined with cytostatic drugs and tyrosine kinase inhibitors (TKIs) in regimens such as the Cleopatra regime [7]. The introduction of mAbs has notably improved early-stage breast cancer treatment outcomes, though only a minority of patients respond positively to initial treatment, leading to a poor prognosis in advanced stages [8,9]. Metastasis and resistance development often result in low cure rates and limited survival. The SOPHIA study showed that margetuximab combined with chemotherapeutics offered no significant overall survival benefit compared to trastuzumab but could be an alternative for patients not responding to the Cleopatra regime [10].
In oncology, significant advancements have also been made in the development of prophylactic vaccines, especially against cervical cancer, which is often linked to persistent high-risk human papillomavirus (HPV) infections. Current prophylactic vaccines are effective against 90% of HPV infections but offer limited benefits for preexisting infections, underscoring the need for therapeutic vaccine development [11].
Moreover, in the field of biomedical applications, significant advancements have been achieved through the integration of nanoparticle usage in combination with biological systems. Nanoparticles have gained attention for their potential in medical applications, particularly drug delivery and therefore cancer treatment. Organic nanoparticles are valued for their unique properties but face challenges like potential toxicity, instability, and limited in vivo circulation [12,13,14,15].
To enhance the therapeutic outcomes of nanoparticle-mediated drug delivery, it is crucial to develop carriers that not only release drugs in a controlled manner but also efficiently navigate through and overcome the body’s complex biological barriers [16,17]. Overcoming these barriers significantly contributes to the nanoparticles’ therapeutic effectiveness, which refers to their ability to improve health outcomes by delivering drugs more efficiently to the target site, thereby maximizing the therapeutic impact while minimizing side effects [17,18]. In this context, protein-based nanoparticles stand out as a particularly promising carrier option due to their biocompatibility, biodegradability, and relatively low potential to elicit immune responses [18,19]. These nanoparticles can be engineered to respond to specific physiological stimuli—such as pH changes in tumor cells, temperature shifts, or enzymatic activity—enabling them to cross biological barriers more effectively and release their drug payload precisely where needed [17,18]. The mild and nontoxic conditions required for their preparation further underscore their suitability for therapeutic applications [19,20,21,22]. Thus, by leveraging the unique properties of protein-based nanoparticles, it is possible to achieve a higher level of control over drug release kinetics and targeting, directly linking the design and functionalization of these carriers to improved therapeutic outcomes.

2. Proteins in Conjunction with Nanoparticles

In the realm of biomedical applications, the fusion of nanotechnology with biological systems has led to remarkable advancements, particularly in the development of sophisticated drug delivery systems. One significant breakthrough in this area is the advent of lipid-based nanoparticles, such as liposomes, which have already made their way into clinical use [16]. For instance, liposomal formulations like Doxil®, the liposome-encapsulated doxorubicin, represent a milestone in chemotherapy, offering targeted delivery that reduces toxicity and improves efficacy in cancer treatment [23]. This innovation exemplifies how nanotechnology can revolutionize therapeutic strategies by enhancing the delivery and bioavailability of drugs. Building on such foundational advancements, researchers are now exploring even more innovative approaches, including the design of stimuli-responsive nanoparticles that can release their payload in response to specific triggers within the body, as mentioned earlier. These cutting-edge developments signify a new era in personalized medicine, where drug delivery systems are not only more efficient but also finely tuned to the patient’s unique physiological conditions.
An innovative method, magnetically controlled drug delivery, leverages drug-loaded magnetic nanoparticles (MNPs) to precisely target affected areas through an external magnetic field. These MNPs possess a large specific surface area, allowing them to transport significant drug doses directly to the target site, thus achieving high local drug concentrations [24,25]. Their growing recognition in the medical field stems from their advantageous characteristics, including small size, cost efficiency, and adaptability in production and modification, which make them valuable for both diagnostic and therapeutic uses [26,27,28].
Iron oxide nanoparticles (IONs) are widely utilized as contrast agents for T1- and T2-weighted magnetic resonance imaging (MRI) in clinical diagnosis [29,30]. IONs are particularly studied for their role as a T2 contrast agent in MRI because they efficiently shorten transverse relaxation times [31]. Notably, IONs have demonstrated several beneficial properties, including long blood half-lives, low toxicity, and flexible surface chemistry [32,33].
Furthermore, the combination of proteins with nanoparticles is a key example of this technological integration. Proteins or antibodies, with their surface functional groups, allow for easy modification of nanoparticle surfaces as indicated in Figure 1. This characteristic is highly beneficial for targeted drug delivery, diagnostics, and tissue engineering applications [34,35]. Additionally, the hollow structure of certain proteins facilitates the encapsulation of small-molecule drugs or metal nanoparticles, enhancing the potential for drug delivery and combination therapy [36,37].
Nanoparticles, particularly MNPs, are versatile in tumor treatment, with two distinct approaches: (i) conjugating specific antibodies to MNPs for selective binding to receptors and inhibition of tumor growth through drugs, loaded on the particles, resulting in targeted therapy; (ii) employing targeted MNPs for hyperthermia in tumor therapy [38]. These approaches exemplify how nanotechnology enhances the precision and efficacy of biomedical treatments. Beyond liposomal drugs, several nanoparticle-based formulations have gained approval for clinical use, such as Eligard® for prostate cancer in the USA [39] and Nanoxel® for various cancers in India [40]. Additionally, the European Medicines Agency (EMA) has approved Apealea for ovarian cancers [41] and NanoTherm for glioblastoma and other cancers [42], highlighting nanotechnology’s expanding role in approved cancer therapies and approaches. Tests on superparamagnetic iron oxide nanoparticles have also been conducted to track their migration as a magnetic tracer, showing an increase in monitoring counts on the skin’s surface [43].
There exists a variety of proteins suitable for the creation of protein-based nanoparticles, with many being producible through recombinant protein synthesis. Commonly used proteins include fibroin, human serum albumin, gliadin, lipoproteins, and elastin-like polypeptides, as detailed in reviews by Hong et al. [44], Jain et al. [19], and Yao et al. [45]. This chapter focuses on biomimetic materials and natural binding proteins, areas yet to be extensively reviewed and of particular interest in this discussion.

2.1. Various Nanoparticles

In the landscape of advancing biomedical technologies, it is crucial to distinguish between biomimetic nanoparticles and nanoparticles with immobilized proteins, while both are integral to the realm of drug delivery, they exhibit distinct characteristics and functionalities.

2.1.1. Biomimetic Materials

Biomimetic nanoparticles are engineered to mimic biological surfaces, like cells or viruses, in both form and function. This mimicry extends beyond mere structural imitation; it encompasses the replication of biological behaviors and interactions. Therefore, one should differentiate between biomimetic materials/nanoparticles and protein-based biomaterials. Biomimetic materials are an integral part of generating natural mimicry. These nanoparticles are made from a variety of materials, including metals, polymers, lipids, and even composite materials. By emulating not only the chemical composition and structure but also the biological characteristics and functions of natural materials, this approach is instrumental in creating efficient drug delivery systems capable of navigating biological barriers and utilizing cellular recognition and uptake mechanisms [46]. Due to their programmable chemistry and biocompatibility, biomimetic materials have found applications in innovative medical technologies, such as tendon-driven myoelectric soft hand exoskeletons [47,48], biomimetic scaffolds for tendon regeneration [49], cartilage-lubricating polymers [50], and in dentistry [51].
Specifically, apatite nanocrystals [52] and biomimetic F e 3 O 4 nanoparticles coated with red blood cell (RBC) membranes demonstrate the capacity for targeted and controlled drug delivery, with the latter showing prolonged circulation time [53], immune response evasion [54], and immunomodulatory effects [53] resembling artificial antigen-presenting cells (APCs) [54]. These characteristics highlight the potential of biomimetic nanoparticles in regulating immune responses, ensuring long-term circulation and achieving high target specificity.
The ongoing debate around the definition of biocompatibility, as discussed by Naahidi et al. [55], underscores the absence of standardized criteria for evaluating nanoparticle safety in drug delivery systems. This lack of clear guidelines highlights the need for safety assessments of nanoparticles’ impacts on human health, considering their dynamic physicochemical properties and the importance of understanding their biological interactions [55]. The capability of nanoparticles to provide targeted therapy, minimizing side effects while maximizing efficacy, points to their significant potential in revolutionizing drug delivery and diagnostic methods, further emphasizing the importance of research on nanoparticle biocompatibility and mechanisms of action [55].
The second class are protein-based biomaterials, as outlined in the review by Zhang et al. (2023) [56], which exemplify this integration in medical applications. These biomaterials are noted for their “encoded and programmable mechanical properties such as superelasticity, plasticity, shape adaptability, and excellent interfacial behavior, derived from sequence-guided backbone structures” [56]. These materials are primarily made from proteins, which can be sourced from animals or plants. Despite several advancements, the traditional method of regenerating protein materials from natural sources faces issues like low yield and structural damage due to extraction process steps. As highlighted by Lavickova et al. [57], the concentration of DNA templates used for the regeneration of specific proteins plays a significant role in achieving optimal regeneration efficiency. Therefore, developing alternative strategies for fabricating protein materials, like membrane proteins, is crucial [58]. A promising approach is the heterologous expression of natural proteins with a modular assembly approach, involving the creation of standardized, easy-to-assemble protein modules with specific structures and functions [59].
A notable example within this area are “virus-like particles” (VLPs), which are protein-based nanoparticles formed by the self-assembly of viral coat proteins [60,61,62]. These nanoparticles mimic natural viruses in structure but are safe for human use as they lack viral nucleic acids, thus preventing replication and viral infection [63]. However, their resemblance to viruses can potentially activate the immune system [60,64], a challenge that various research projects are addressing through different production hosts like plant [63], yeast [65], or insect cells [62].
In summary, the main difference between both parts of biomimetic materials is their function and utilization. Biomimetic nanoparticles focus on mimicking specific biological functions at the nanoscale for targeted therapies, whereas protein-based biomaterials focus on exploiting the inherent properties of proteins for the applications. In the next section, it is explained how nanoparticles with immobilized natural binding proteins leverage the specificity of protein functions to achieve targeting and interaction within the body, reflecting a more focused approach in biomedical applications.

2.1.2. Natural Binding Proteins

NBPs, including those that attach to DNA and RNA, play a crucial role in both cancer development and treatment, as noted in recent studies [66,67]. Future research will also look into fat- and sugar-binding proteins, which have unique sections known as fatty acid- or carbohydrate-binding domains [68,69]. These proteins are found in various organisms, such as the bacteria Bacillus circulans and fungi like Trichoderma species. They are remarkable for their ability to specifically and strongly attach to certain molecules, including fats, chitin, chitosan, and cellulose [70,71,72]. DNA- and RNA-binding proteins are essential in all forms of life, helping control gene activity by acting as switches that turn genes on or off [73]. This regulation is crucial for making proteins correctly and responding to changes in the environment [73]. Having introduced the pivotal role of natural binding proteins (NBPs) in cancer development and treatment, we now delve deeper into their specific functionalities and mechanisms, which underscore their dual utility in both suppressing tumors and enhancing drug delivery.
NBPs play a pivotal role in cancer therapy, exhibiting dual functionality by both suppressing tumors and enhancing the targeted delivery of drugs with their highly specific binding abilities. These proteins can indirectly influence tumor growth and progression through various mechanisms. They might block interactions between tumors and their surrounding microenvironment, inhibit angiogenesis as noted by Smith [74], or direct the immune system to target and destroy tumor cells. This multifaceted approach not only underscores the importance of NBPs in devising precise treatment strategies but also highlights their utility in diagnostics and therapeutic applications, where their selective binding properties are leveraged for targeted treatments and diagnostic procedures. Beyond their direct impact on tumor growth and interaction, NBPs’ unique capabilities extend to innovative applications in targeted drug delivery and diagnostics. This is exemplified through their precision in attaching to specific molecular targets, a principle that is foundational to advancing cancer therapy.
Furthermore, a specific application of NBPs in targeted drug delivery involves the use of recombinant proteins that possess, for example, either a C- or N-terminal chitin-binding domain. These binding proteins can specifically attach to inert chitin particles [71,75,76], allowing for oriented immobilization. This method is particularly compatible with the human bloodstream, which naturally lacks chitin-like compounds. While targeted drug delivery showcases the therapeutic potential of NBPs, their role is not confined to treatment alone. The following discussion explores how NBPs contribute to the immune system’s response to cancer and serve as powerful tools in diagnostics and prognosis, highlighting their versatility in oncology.
In contrast, substances like chitin and chitosan are known to stimulate cytokine production by activation of transcription factors like NF-κB and AP-1, draw leukocytes, and differently activate macrophages [77,78,79], showcasing a different mechanism by which NBPs can contribute to the immune system’s response to cancer.
NBPs also serve as tumor markers, aiding in diagnostics and prognosis, like prostate-specific antigen (PSA) used for prostate cancer screening [80]. Their specificity makes them ideal candidates for targeted drug development, aligning with personalized medicine trends in oncology [45]. Additionally, NBPs’ influence on the tumor microenvironment provides insights into cancer progression and new treatment strategies.
A promising technique for detecting liver cancer uses nanoparticles paired with a special protein linker, illustrated in Figure 2. This linker has two main functions: it connects to the nanoparticles through a chitin-binding domain, and it targets liver cancer cells that express a specific protein, glypican-3 (GPC3) [81]. GPC3 is often found in high amounts on liver cancer cells but is rare in healthy liver tissue.
In this method, nanoparticles carrying this dual-function protein linker act as enhanced contrast agents in medical imaging. When injected into a patient, the GIP1 part of the linker specifically binds to GPC3 on the liver cancer cells. This targeted binding leads the nanoparticles to accumulate precisely at the tumor site. To illustrate the practical impact of NBPs in oncology, let us examine a case study focusing on liver cancer detection. This example demonstrates how NBPs, when integrated with advanced nanoparticle technology, can revolutionize cancer diagnostics by improving accuracy and specificity.
The key to this approach is the dual ability of the protein linker. It can both adhere to the chitin on the nanoparticles and latch onto GPC3 on the liver cancer cells. This dual action improves the performance of the nanoparticles as imaging agents and increases the accuracy of tumor detection in scans.
Using this targeted approach for contrast agents results in more precise and detailed imaging. This is vital for the early detection of liver cancer, accurately determining the tumor’s size and location. The deployment of GIP1, efficiently produced in Escherichia coli cells as indicated by Janski et al. [82], marks a significant step forward in liver cancer diagnostic imaging.
In the treatment phase, the nanoparticles are loaded with chemotherapy drugs or other effective medications. Postinjection, these complexes bind specifically to liver cancer cells, releasing the drug right at the target site. This targeted approach allows for a higher concentration of the drug at the tumor site, sparing healthy tissue. It enhances treatment efficacy and simultaneously reduces side effects.
This method represents a novel approach to targeted drug delivery, illustrating how NBPs can be utilized. Due to their specific attachment to nanoparticles, they enable more effective and less invasive cancer therapies.
In summary, the integration of natural binding domains represents a promising direction in drug delivery technology. These novel platforms aim to overcome existing limitations and revolutionize drug delivery. This chapter underscores the importance of bio-inspired design and advanced material engineering in developing effective drug delivery systems that navigate the complexities of biological systems and optimize therapeutic outcomes.

3. Exploring the Role of Binding Domains in Cancer Treatment: Applications, Innovations, and Impact in Oncology

The current landscape of oncology is experiencing a significant paradigm shift, largely propelled by the advancements in antibody–drug conjugates (ADCs) [83], which epitomize the exploration and application of novel approaches. This evolution is characterized by a strategic transition towards highly targeted cancer therapies that promise enhanced efficacy and minimized toxicity, a leap forward from the constraints of traditional chemotherapy [83]. This shift is quantitatively evident in the deployment of medications, notably through a significant reduction in chemotherapy treatments, such as a 20% decrease in its use for breast cancer due to genomic testing [84].
Exploring further becomes possible through the use of binding domains found in NBPs. NBPs are distinguished by their remarkable capability to bind specifically to certain molecules. The integration of ADCs with NBP motifs is heralding a new era in the realm of oncology, offering a promising pathway toward the realization of highly specific and efficacious cancer therapies and diagnostics. This innovative approach not only leverages the precision targeting capabilities of monoclonal antibodies inherent in ADCs but also enhances therapeutic outcomes through the synergistic combination with diverse NBP motifs.
The following section is dedicated to examining the binding domains in NBPs shown in Figure 3, their natural occurrence, and their significant roles. It highlights the various applications they have in cancer treatment, the innovative approaches being developed around them, and the extensive impact they are making in the field of oncology. Furthermore, the utilization of these binding domains across different facets of cancer therapy are explored, including the domains’ use in targeted drug delivery systems, the development of novel diagnostic tools, and their potential to fundamentally transform cancer treatment methodologies.

3.1. DNA-Binding Domains

DNA–protein interactions are crucial for regulatory proteins, which recognize specific DNA sequences of 8–20 base pairs amidst millions, guiding the protein to its functional area [85]. DNA-binding domains (DBDs) on the other hand, are crucial molecular components that enable proteins, particularly transcription factors, to interact specifically with DNA. These domains have evolved to recognize and bind to specific DNA sequences, displaying diverse structural features, including α -helices, β -sheets, and disordered regions [86]. These structures, such as helices and loops, interact with DNA’s grooves and turns to identify specific sequences [86]. What distinguishes these proteins from others is primarily their ability to specifically identify DNA sequences among the vast expanse of the genome, enabling precise regulation of gene expression. This specificity is achieved through the combination of structural motifs within the protein that match the unique shape and chemical properties of target DNA sequences. For instance, the helix–turn–helix motif, commonly found in these proteins, allows for snug fitting into the DNA major groove, where it can make specific contacts with the bases [87]. Similarly, zinc finger motifs use a combination of alpha helices and beta sheets stabilized by zinc ions to recognize specific DNA sequences [88]. In contrast, other proteins might interact with DNA in a more generalized manner, lacking the fine-tuned specificity of these binding domains. These might include histone chaperones, which effectively prevent nonspecific contacts between the negatively charged DNA and the positively charged histones, ensuring an orderly assembly of the nucleosome structure [89], or enzymes like DNA polymerase, which reads the DNA template but does not have the sequence-specific binding properties of transcription factors or other DNA-binding proteins discussed here. The target search of DNA polymerase is dominated by transient nonspecific DNA binding [90]. The mobility of these proteins during their target search is dictated by DNA interactions rather than their molecular weights [90]. Specifically, DNA-binding proteins, regardless of their size, concentration, or function, spend the majority (58–99%) of their search time bound to DNA, indicating that transient DNA-binding events dominate the target search process [90].

3.1.1. Classification, Characteristics, and Function

DBDs are typically categorized based on their structural characteristics. As a consequence, transcription factors (TFs) are grouped into families according to the type of DBD they contain. In general, large domain databases classify protein domains hierarchically; while the class reflects the three-dimensional structure, the architecture describes the arrangement of secondary structures (Table 1). A superfamily is a protein group with common evolutionary origin, and the family has clear evolutionary relationships [91]. However, the literature does not appear to have adopted this structure. Regarding DBDs, a distinction is made between five different superclasses of domains: basic domains, zinc-coordinating DNA-binding domains, helix–turn–helix, beta scaffold factors with minor groove contacts, and other transcription factors (indicated in Table 1). Each superclass consists of several classes; for example, leucine zipper factors, helix–loop–helix factors, and their combinations, are classes of basic domain superclass [86]. Generally, transcription factors within the same family show similar DNA-binding specificities, although variations can occur due to changes in specific amino acids within the DBD [92].
One of the defining characteristics of DBDs is their modular nature, allowing them to be isolated from the rest of the transcription factor without loss of function and therefore allowing the study of multiple effects [94]. This modularity is advantageous for structural studies, facilitating techniques like crystallization or nuclear magnetic resonance (NMR) spectroscopy. Hence, the structures of DNA-binding domains alone or combined with DNA can be easily observed [86].
A prominent example of a protein with distinct DBDs is the Epstein–Barr virus nuclear antigen 1 (EBNA1). EBNA1’s DNA-binding region consists of two different domains: the C-terminal (core domain, residues 504–607 [95]) and the N-terminal (flanking domain, residues XY-YZ) [96]. The flanking domain is unique to EBNA1 and crucial for sequence-specific binding. This domain attaches to the outer portion of the EBNA1-binding site, while the core domain connects to the inner portion [96]. Interestingly, the core domain structurally resembles the DNA-binding and dimerization domain of the E2 protein from bovine papilloma virus, indicating also a role in sequence-specific DNA binding. This structural resemblance is notable given the lack of sequence similarity and known evolutionary links between the EBNA1 (herpes) and E2 (papovavirus) virus families [96].
Certain types of DNA-binding domains stand out due to their ubiquity and functional relevance. For example, the superclass of zinc-coordinating DNA-binding domains utilizes a zinc atom, often coordinated by cysteine and histidine residues, to recognize three to four bases of DNA [97,98]. This domain is frequently found in transcription factors like Sp1 [99]. Studying the superclass of helix–turn–helix (HTH) transcription factors can provide further insight into DBD functionality. X-ray crystallography has revealed their surface structure, including a short α -helix known as the recognition helix, predicted to fit partially within DNA’s major groove [87]. This structural feature enables specific interactions between residues and DNA bases, crucial for sequence-specific DNA binding, as observed in proteins like the cyclic AMP receptor protein (CRP) of E. coli, the bacteriophage λ regulatory protein Cro, and the NH2-terminal domain of λ repressor [97]. The typical dissociation constants of DNA-binding proteins are in the mid- to lower molar range [100,101] (130–1000 nM), indicating a very high affinity.

3.1.2. The Role of DBDs in Oncology and Applications

Transcription factors, crucial in oncology due to their DNA-binding roles, become potential drug targets when mutated or dysregulated, leading to cancer by disrupting gene expression, including pathways for cell differentiation and death [102,103,104]. Targeting transcription factor activity has shown promise both preclinically and clinically through strategies like inhibiting protein interactions, DNA binding, and modulating degradation processes [105]. Innovations including modulation of auto-inhibition, use of proteolysis targeting chimeras (PROTACs), and combination therapies aim to refine cancer treatment by targeting these transcription factors’ unique properties [106].
Enhancers are regulatory parts of DNA that are involved in controlling which genes are turned on in different body tissues. New research indicates that point mutations in these enhancers, or in elements that help enhancers communicate with other parts of DNA, can lead to cancers that specifically affect certain tissues [106].
One of the key approaches in developing cancer therapeutics involves targeting the specific interactions between DBDs and DNA. This targeted approach is pivotal in enhancing the efficacy of cancer therapy because it directly interferes with the functioning of potent oncogenic transcription factors [107]. One example is FOXM1, a transcription factor crucial for cancer initiation, progression, and drug resistance, and its regulatory network, which is therefore a major predictor of adverse outcomes in various human cancers [107]. Furthermore, high-throughput screening methods have been instrumental in identifying and selectively inhibiting DNA-binding proteins [108]. Additionally, the study of proteins like Smad4, a TGF- β -inducible DNA-binding protein, underscores the importance of these proteins in understanding cancer biology and devising treatment strategies [109]. Smad4’s involvement in TGF- β signaling pathways highlights the intricate relationship between growth factors and gene regulation in the development of cancer [109]. The advancement of such recent therapeutic strategies represents a significant development [108].
Beyond the scope of oncology, DBDs play a crucial role in molecular biology and biotechnology. Customized DBDs can be used to manipulate DNA in a sequence-specific manner, a principle integral to technologies such as CRISPR/Cas9 [110]. By coupling DBDs with transcriptional modulators, researchers can regulate gene expression, providing valuable insights into cellular pathways [110].
In the context of DNA-binding drugs, examples such as amsacrine demonstrate their effectiveness in treating acute lymphoblastic leukemia by targeting DNA topoisomerase II [111]. The development of DACA (N-[2-(dimethylamino)ethyl]acridine-4-carboxamide) for lung adenocarcinoma [112], along with its derivatives like SN 28049, illustrates the evolving landscape of DNA-binding drugs in cancer therapy [113]. While these drugs act by disrupting topoisomerase II, an enzyme that helps manage the structure of DNA during cell division, there remain challenges in pharmacokinetics and toxicity because amsacrine can also affect normal, healthy cells that divide rapidly [114]. Future research is geared towards understanding the interplay between DNA-binding drugs, topoisomerase, and the immune system, with the aim of improving cancer treatment strategies by preventing harm to healthy cells.

3.2. Protein-Binding Domains

In the complex landscape of cellular biology, proteins rarely operate independently. Instead, they engage in intricate networks of interactions, which are crucial in a multitude of cellular functions. This section delves into protein–protein interaction (PPI) domains, specialized regions that enable such interactions with high specificity [115].

3.2.1. Characteristics and Functions of PPI Domains

PPI domains facilitate the precise and selective interaction between proteins. They act as specialized docking stations, allowing proteins to recognize and bind to each other. PPI domains are fundamental to mechanisms such as signal transduction pathways, cellular trafficking, DNA replication, and cell-cycle control [1]. Many of these processes involve a protein domain binding to a short sequence (3–10 amino acids) of another protein characterized by a specific pattern [116]. For instance, the POZ (pox virus and zinc finger) domain [117], or the BTB/POZ domain found in genes of DNA viruses [118], exemplifies this binding specificity. While zinc finger domains are predominantly recognized for their DNA-binding abilities and role in transcription factors [97], they also possess the capacity to bind to protein sequences [119]. In general, PPI domains are integral in ensuring that cellular processes are conducted with precision, specificity, and coordination [115]. A more detailed holistic classification of these domains is not available from the literature. The determined affinities of PPIs vary in ranges between 100 and 3000 nM (Table 4). Due to the sheer diversity and size of protein–protein bonds, mapping and classifying the domains is a challenge [120]. While rule-based algorithms allow for some classification of PPIs, the specific classification process may vary depending on the algorithm used [121]. Therefore, many different approaches for classification, based on machine learning, have been published. For example, Urquiza et al. found the eight important features for the prediction of PPIs, which were validated by a ROC analysis [122]. A web server called Protein Complex Prediction by Interface Properties (PCPIP) is provided by Subhrangshu and Saikat [120], which can predict whether the interface of a given protein–protein dimer complex resembles known protein interfaces. The server is freely available at http://www.hpppi.iicb.res.in/pcpip/ (accessed on 29 March 2024).

3.2.2. The Role of PPI Domains in Cellular Processes and Infections

The significance of PPI domains is underscored by their governance over a vast array of cellular processes.
Through this exploration, we aim to highlight the indispensable role of PPI domains in the orchestration of cellular activities, in particular bacteriophage infection, and their potential implications in understanding and targeting various biological processes.
Receptor-binding proteins (RBPs), a subclass of PPI, play a crucial role in the specificity of bacteriophages, primarily determining their host range through interactions with various bacterial surface structures [123]. RBPs can be divided into two main classes based on their morphology: tail fibers and tailspike proteins (TSPs) [124]. Tail fibers are characterized by their long, slender, fibrous structure without enzymatic activity [125]. In contrast, TSPs are shorter, stockier, and typically possess enzymatic activity, often targeting specific surface structures like sugar moieties [125].
The interaction between bacteriophages and bacterial hosts is mediated by RBPs, which are the first point of contact. They bind to a range of structures displayed on the bacterial surface, including outer membrane proteins, lipopolysaccharides, capsular polysaccharides, and even organelles such as flagella or pili [126,127]. This interaction is a two-stage capture model, beginning with initial reversible binding, followed by more specific and irreversible binding to the receptors [128,129]. This process is essential for the phage’s infection process.
Furthermore, RBPs serve as the primary and most important checkpoint in the infection process [125]. These domains show significant sequence diversity, reflecting their specificity to host receptors and varying depending on the type of host receptor recognized and the infection process [123]. This diversity underscores the critical role of RBPs in mediating the specificity of bacteriophages to their bacterial hosts. In Gram-positive bacteria like B. anthracis, the cell wall is distinct in composition and structure. It lacks an outer membrane and features a thick peptidoglycan layer, transmembrane peptidoglycan-recognition proteins, and the nucleotide-binding oligomerization domain [130]. The transmembrane peptidoglycan-recognition proteins are potential phage receptors. For instance, the B. anthracis receptor for Wγ phage has been identified as the LPXTG protein (a motif, known to be anchored by sortases to the bacterial peptidoglycan) GamR (gamma phage receptor) [131,132]. This protein’s role in virion binding, and the necessity of a potential secondary receptor for DNA delivery, highlights the complexity of phage–host interactions [132].

3.3. Fatty Acid-Binding Domains

In the following section, we turn our attention to fatty acid-binding domains, a minor but ubiquitous class of binding domains present across all organisms. After examining the two primary classes of binding domains that play pivotal roles in cellular processes, this segment aims to explore the significance and applications of fatty acid-binding domains. Notably, their potential in diagnostics, such as identifying structures composed of fatty acids like hydrophobic layers, highlights their importance despite being a less prominent class.
In cellular biology, fatty acid-binding domains (FABDs), such as the intestinal fatty acid binding domains (IFABPs) [68], are a key part of intracellular lipid-binding proteins. These domains are essential for identifying and attaching to fatty acids. The structure of fatty acid-binding proteins includes a β -barrel made up of 10 antiparallel β -sheets, which is topped by two short α -helical segments [68]. Proteins with these specific areas are known as intracellular lipid-binding proteins [133,134]. According to the literature, there is no existing classification for fatty acid-binding domains. However, classifications for fatty acid-binding proteins in human cells have been published based on the gene that expresses them (Table 2).
Table 2. List of fatty acid-binding proteins, based on the gene expression data from Smathers and Petersen [134].
Table 2. List of fatty acid-binding proteins, based on the gene expression data from Smathers and Petersen [134].
GeneCommon NameAliases for ProteinsLocalization
FABP 1Liver FABPL-FABP, hepatic FABP, Z-protein, heme-binding proteinLiver, intestine, pancreas, kidney, lung, stomach
FABP 2Intestinal FABPI-FABP, gut FABP (gFABP)Intestine, liver
FABP 3Heart FABPH-FABP, O-FABP, mammary-derived growth inhibitor (MDGI)Cardiac and skeletal muscle, brain, kidney, lung, stomach, testis, adrenal gland, mammary gland, placenta, ovary, brown adipose tissue
FABP 4Adipocyte FABPA-FABP, aP2Adipocytes, macrophages, dendritic cells, skeletal muscle fibers
FABP 5Epidermal FABPE-FABP, keratinocyte-type FABP (KFABP), psoriasis-associated-FABP (PA-FABP)Skin, tongue, adipocyte, macrophage, dendritic cells, mammary gland, brain, stomach, intestine, kidney, liver, lung, heart, skeletal, muscle, testis, retina, lens, spleen, placenta
FABP 6Ileal FABPIl-FABP, ileal lipid-binding protein (ILLBP), intestinal bile acid-binding protein (I-BABP), gastrophinIleum, ovary, adrenal gland, stomach
FABP 7Brain FABPB-FABP, brain lipid-binding protein (BLBP), MRGBrain, central nervous system (CNS), glial cell, retina, mammary gland
FABP 8Myelin FABPM-FABP, peripheral myelin protein 2 (PMP2)Peripheral nervous system, Schwann cells
FABP 9Testis FABPT-FABP, testis lipid-binding protein (TLBP), PERF, PERF 15Testis, salivary gland, mammary gland
FABP 12//Retinoblastoma cell 1, retina (ganglion and inner nuclear layer cells) 2, testicular germ cells 2, cerebral cortex 2, kidney 2, epididymis 2
1 Expression found in humans, 2 expression found in rodents.

3.3.1. Function and Specificity of FABDs

FABDs are key to the transport, storage, and metabolism of fatty acids within cells [135]. Their primary function is to bind long-chain fatty acids, enhancing their solubility in the aqueous environment of the cell and aiding their transportation to specific cellular sites. The affinity and specificity of these domains for particular fatty acids are influenced by the fatty acid’s saturation level. Generally, these domains show increased affinities for more hydrophobic molecules and decreased affinities for molecules with shorter chain lengths and higher unsaturation levels [134]. The binding affinity of FABDs is usually in the nanomolar range and varies depending on the chain length of the fatty acid [136]. While there is a high affinity for long-chain fatty acids, the affinity significantly drops (often >500 nM) for other hydrophobic ligands [136].

3.3.2. Structural Characteristics of FABDs

The structure of fatty acid-binding domains (FABDs) is characterized by hydrophobic pockets that create an ideal environment for accommodating the fatty acid tail [137,138]. In tandem, specific amino acid residues within these domains engage the carboxyl head of the fatty acid, ensuring efficient and precise binding. This dual interaction plays a crucial role in the stability and functionality of FABDs.
Moreover, the stability of these interactions is enhanced by hydrogen bonds [139,140] and van der Waals forces [141]. These molecular forces not only stabilize the binding but also increase the affinity and specificity of the process. The amino acids are arranged in such a way that they often form a binding groove or cavity, tailored to fit specific fatty acids.
For example, in FAB-5—a subtype of FABDs—this tailored binding cavity is essential for its function, demonstrating the critical role of structural specificity. This specificity is vital for the biological functionality of FABDs, as it governs the selectivity for different fatty acids, influencing various cellular processes [141].

3.4. Carbohydrate-Binding Domains: Chitin-, Chitosan-, and Cellulose-Binding Domains

Carbohydrate-binding domains, also known as carbohydrate-binding modules (CBMs), are critical for specific binding to insoluble polysaccharides such as chitin, chitosan, and cellulose [142]. CBMs, naturally found in various organisms including Bacillus species and soil organisms, are integral to enzymes like chitinase. They function by reducing the distance between the substrate and the catalytic domain, thereby enhancing enzyme efficiency [142]. Usually, CBMs are classified based on amino acid similarities [69,143]. In the last 20 years, the number of families has increased from 39 to over 100. A further grouping into superfamilies has not been imposed yet [69]. However, Boraston et al. [69] further organized the CBM families into the following seven different groups based on structural similarities: β -sandwich, β -trefoil, cysteine knot, unique, OB fold, hevein fold, and unique (contains hevein-like fold). For the overview of our structure, we grouped the individual families according to the ligands chitosan, chitin, cellulose, and others; see Figure 4. For a more comprehensive breakdown of the classification based on the bound substrate, refer to the Appendix A section, specifically Table A1.

3.4.1. Chitosan-Binding Domain

Chitosan has been widely studied for biomedical applications due to its biocompatibility and biodegradability. It is a derivate of the linear polysaccharide chitin. However, while chitin is composed of GlcNAc, chitosan is composed of GlcNAc and GlcN.
The chitosan-binding domain is a specific region within proteins or peptides that binds to chitosan, such as chitosanases [72]. Classified as carbohydrate-binding modules (CBMs), these domains are part of the carbohydrate-active enzymes, and their binding often depends on chitosan’s physical and chemical properties [72,145]. Chitosan-binding domains interact with chitosan through electrostatic interactions, hydrogen bonding, and hydrophobic effects. For example, since chitosan is amorphous, it is readily hydrolyzed by chitosanases. Chitosanases, however, do not act on chitin. The binding of chitosan-binding domains to chitosan, but not chitin due to acetylation, is facilitated by Van der Waals interactions and hydrophobic residues [145].
Furthermore, discoidin domains (DDs) in proteins, particularly those in CBM32 from Dictyostelium discoideum, demonstrate affinity for carbohydrates, including chitosan [146,147].
When combined with probes or markers, chitosan matrices can detect specific cancer cells or tumor microenvironments. This specificity can pave the way for developing diagnostic tools with higher accuracy and sensitivity [148].

3.4.2. Chitin-Binding Domain

Chitin-binding domains (ChBDs) are a crucial component in enzymes that interact with carbohydrates. They bind catalytically active parts of the enzyme to a specific carbohydrate and concentrate them near the substrate [69,144].
Intein-mediated protein splicing is an application of ChBDs in recombinant protein purification [149,150]. The target protein is present as an N-extein to which an intein is bound. The ChBD is in turn fixed to this intein. The protein is isolated using a chitin affinity column by binding the ChBD to chitin in the column material. The thioester bond between the target protein and intein can be cleaved by adding higher concentrations of a free thiol via thiolysis. Higher temperatures also result in the release of the target protein [150]. The binding of the ChBD to chitin is mainly based on hydrophobic interactions between the aromatic side chains and the aliphatic regions in the pyranose ring of chitin [151]. The ChBD selectively binds to chitin and not to soluble derivatives of chitin or cellulose, as an antibody selectively binds to an antigen [75,152,153]. In general, different methods are known to characterize bindings. The binding affinity is defined as the tendency of two molecules to form a bond. The dissociation constant K D , also known as the binding constant, is often used to describe this affinity. It reflects the balance between the dissociated and undissociated form and thus the average amount of binding; while high K D values (> 10 3 mol/L) indicate weak, unspecific binding, low K D values (< 10 10 mol/L) are a sign of very strong binding. Antigen–antibody bonds have binding constants in the nano- to micromolar ranges. The affinity also depends on the conditions in which the binding partners are present. The lower the affinity, the less specific the reaction of the antibody with the antigen.
The affinity of in E. coli recombinant synthesized ChBD from B. circulans indicates a dissociation constant of 149.72 ± 30.44 nM toward chitin of yeast cell bud scars [75]. Most K D values determined in the literature for proteins with bacterial ChBDs are 1–10 μM [76]. When determining the dissociation constant, a distinction is often made between α -chitin and β -chitin. For ChB proteins from B. thuringiensis, 3.460 ± 1.300 μM ( β -chitin) and 5.250 ± 1.400 μM ( α -chitin) were determined [154]. The same research group also determined ChBP values for ChBP derived from B. licheniformis with 4.120 ± 1.600 μM ( β -chitin) and 5.980 ± 2.100 μM ( α -chitin) [154]. A K D value of 1.400 ± 0.400 μM (β-chitin) was determined for the ChBP CBP21p from Serratia marcescens [155]. CBP21 is part of the chitinase B of S. marcescens [155].

3.4.3. Cellulose-Binding Domain

Cellulose-binding domains (CBDs) are polypeptide bonds that belong to the subcategory of carbohydrate-binding modules. There are more and more modules being found in carbohydrate-active enzymes [156]. For this reason, these are also often investigated.
Cellulose-binding domains are generally found in cellulose-degrading enzymes such as cellulase [156]. Cellulase has a modular structure and is equipped with two domains. Most cellulases consist of a catalytic domain and a cellulose-binding domain, which are connected by a linker [157,158]. CBDs can occur both singly and repeatedly in these enzymes, with amino- or carboxy-terminal localization with respect to the catalytically active domain [157]. The catalytic domain contains the active center with the amino acid residues, which is responsible for the hydrolysis mechanism [158]. CBDs have highly conserved sequences with three aromatic residues. The binding of CBD to cellulose substrates is based on the interaction between the glucose rings of cellulose and aromatic amino acids, which are structurally located on the flat side of the domain [157,159].
CBDs mediate the adsorption of the enzyme to the substrate. This adsorption increases the concentration of the enzyme on the insoluble cellulose surface [157], which leads, for example, to an acceleration of enzyme-catalyzed hydrolysis [159].
To date, more than 180 different CBDs have been identified and categorized into more than 13 different protein families based on their amino acid sequence similarities. These can vary in size from 4 to 20 kDa and occur at different positions within the polypeptides: N-terminal, C-terminal, or internal [160]. Most CBDs belong to families I, II, and III [159]. Family I CBDs are compact polypeptides binding cellulose by three aromatic residues [161]. The CBDs of families II and III are much larger (and contain 90–100 and 130–172 residues), respectively [160]. They are specific for bacterial enzymes [160].
In addition to different structures, CBDs also have various properties. Some CBDs bind strongly to cellulose and can be used to immobilize active enzymes tightly [162]. Others bind reversibly and are better suited for separation and purification. Family I CBDs bind reversibly to crystalline cellulose and are a useful tag for affinity chromatography [163]. Interaction occurs through hydrogen bonding and van der Waals interaction [163]. They bind to cellulose in a pH range of 3.5 to 9.5, and the affinity of the tag is so strong that an immobilized fusion protein can only be released with buffers containing urea or guanidine hydrochloride. Thus, these denaturing elution conditions require refolding of the recombinant target protein [164]. In contrast, proteins with CBDs of families II and III can be eluted with ethylene glycol [164]. This is due to the low polarity of the solvent, which presumably interferes with the hydrophobic interaction at the binding site. Ethylene glycol can be easily removed by dialysis. In contrast to family I CBDs, family II CBDs can enhance the physical destruction of cellulosic fibers and release small particles from cotton fibers [165].

3.5. RNA-Binding Domains

RNA-binding domains (RBDs) are crucial regions within proteins, enabling specific recognition and binding to RNA molecules [166]. RNA-binding proteins (RBPs), a vast class of over 2000 proteins, ubiquitously interact with and regulate transcripts across various RNA-driven processes [167]. The central role of RNA in numerous cellular functions, from protein synthesis to gene regulation, underscores the importance of understanding RBDs and their interactions with RNA. This group of binding domains is categorized using various approaches, with the two most prevalent ones detailed in Table 3. These classifications are founded on distinctions among various domains.
Another method classifies RNA-binding proteins by the type of RNA that binds within their catalytic domains, according to Jahandide et al. [169]. The second approach for classification focuses on categorizing RNA-binding proteins based on the type of RNA they interact with. This methodology delineates specific groups depending on whether the proteins bind to 7S RNA, double-stranded (DS) RNA, messenger RNA (mRNA), or ribosomal RNA (rRNA). This system allows for a nuanced understanding of the functional diversity among RNA-binding proteins, emphasizing the significance of the RNA type engaged in the catalytic domain of these proteins.

3.5.1. Structure and Function of RNA-Binding Proteins

RBDs engage with RNA through various interaction mechanisms, including hydrogen bonds, Van der Waals interactions, hydrophobic interactions, and π stacking interactions [170,171]. Statistical analysis reveals that approximately 23% of these contacts are potential hydrogen bonds, 72% are van der Waals interactions, and 5% are short contacts [170]. Specific binding typically arises from the combination of multiple RNA-binding regions along with additional weaker interactions with all parts of the RNA nucleotide.
The diverse RNA-binding protein family includes several notable subfamilies. The CUGBP Elav-like family (CELF) and muscleblind-like (MBNL) RBPs are instrumental in regulating alternative splicing and mRNA stability [172,173]. CELF proteins, comprising six members, have complex functions in both the nucleus and cytoplasm, influencing mRNA processing and stability [174]. Notably, CELF1 and CELF2 can function as tumor suppressors or oncogenes, depending on the cancer type [174]. Pharmacological targeting of CELF proteins, especially through organelle-specific drug delivery, presents new possibilities in cancer treatment [174].

3.5.2. Important RNA-Binding Proteins in Therapeutic Applications

MBNL proteins, including MBNL1, MBNL2, and MBNL3, exert multifaceted control over gene expression. A study highlighting MBNL2’s role in tumorigenesis revealed its influence on cyclin-dependent kinase inhibitor 1A (p21CDKN1A) expression and DNA damage responses [175]. Manipulating MBNL2 levels impacts checkpoint kinase 1 (CHK1) phosphorylation, DNA repair, and cellular senescence, suggesting potential therapeutic avenues [170].
Generally, RNA-binding proteins can be classified by their binding mechanisms or their structural organization [176]. Around two-thirds of all studied mRNA-binding proteins are identified as having RNA recognition motif (RRM) domains. Within the MBNL family, zinc finger domains are recognized as superior [172]. Other important domains include DEAD-box helicase, KH domains, and cold shock domains, which are discussed in references [168,176,177].
AU-rich element RBPs (AU-RBPs) are another group of RNA-binding proteins with canonical and noncanonical functions. They are crucial in post-transcriptional gene regulation, particularly regarding DNA damage response and genomic stability. AU-RBPs like ZFP36 and AUF1 have implications in breast cancer [178]. Musashi proteins (MSI-1 and MSI-2), post-transcriptional regulators, are associated with cancer stem cell characteristics in ovarian cancer [179]. Strategies involving the dual knockdown of MSI1 and MSI2 show promise in ovarian cancer therapy [179].
Stress granules (SGs), cytosolic compartments formed under cellular stress, are emerging as important factors in liver diseases, including hepatocellular carcinoma (HCC) [180]. The RBP components of SGs are linked to HCC, highlighting their therapeutic potential.
Moreover, Kang et al. present various therapeutic strategies involving RNA-binding proteins, suggesting that detailed analyses of tumor molecular signatures could identify specific RBPs as targets in personalized cancer treatment [181].
R-loops, RNA/DNA hybrids, play dual roles in cells, affecting genomic stability and DNA damage responses [182]. Understanding the regulation of R-loops is vital for future therapeutic strategies, especially in cancer. For instance, Rad51, a factor in homologous recombination, is involved in R-loop formation, connecting RBPs to genomic stability [182,183].
In summary, RBPs operate in both the nucleus and cytoplasm, regulating RNA transcription and metabolism. Mutations in RBPs are associated with tumorigenesis, emphasizing their role in genomic stability. Future research may uncover the complex mechanisms by which RBPs control RNA/DNA hybrids, offering insights for treating cancer and other disorders [184].

3.6. Aptamers: The Nucleic Acid Antibodies

The last binding elements to be considered are aptamers. Aptamers are oligonucleotides, encompassing ribonucleic acid (RNA), single-strand deoxyribonucleic acid (ssDNA), or peptide molecules, characterized by their ability to bind to targets with high specificity and affinity. This binding capability arises from their unique three-dimensional structures [185]. Aptamers vary in length, typically ranging from 20 to 100 nucleotides. RNA and ssDNA aptamers, despite binding to the same targets, may differ in sequence and structural patterning [185,186]. As versatile biomaterials, aptamers have garnered attention in various fields, including biosensing, drug discovery, therapeutics, diagnostics, and drug delivery systems [187,188].

3.6.1. Stability and Viability of Aptamers

Aptamers are composed of oligonucleotides, which exhibit greater thermal resistance compared to proteins, maintaining their structures through repeated cycles of denaturation and renaturation [189,190]. In contrast, proteins tend to denature and lose their tertiary structure at elevated temperatures [190]. This robustness at high temperatures provides a significant advantage for aptamers over protein-based antibodies, as aptamers can reanneal to regain their original shape and binding capability [191,192].

3.6.2. Binding Mechanism

The binding mechanism of aptamers involves various forces, including van der Waals forces, hydrogen bonding, and electrostatic interactions [193,194,195]. Aptamers often exhibit a preference for positively charged sites in target proteins, as seen in complexes with NF-κB, bacteriophage MS2 capsid, and lysin and arginin side chains [185,195]. However, exceptions exist, like the RNA aptamer targeting the human IgG1 Fc domain (hFc1), which binds despite the absence of positive charges on hFc1’s surface. It relies on weaker forces such as hydrogen bonds and hydrophobic contacts [196].

3.6.3. Applications and Regulatory Milestones

Aptamers, particularly in oncology, offer potential in targeting cancer cells, tumor microenvironments, and molecules associated with tumor progression. They serve as both therapeutic agents and diagnostic tools because of the specific binding [197,198]. Optimizing aptamer sequences to improve binding affinity and specificity is crucial. This process can be achieved using the systematic evolution of ligands by exponential enrichment (SELEX) approach [185,199], which is reviewed by Kohlberger and Gadermeier [199].
Furthermore, aptamers have been investigated for targeting molecules associated with diseases like cancer or viral infection, such as adenovirus or SARS-CoV-2 [200]. A notable regulatory milestone was the FDA’s approval of pegaptanib, an aptamer targeting vascular endothelial growth factor (VEGF), for treating neovascular (wet) age-related macular degeneration (AMD) in 2004 [187].

4. Conclusions

The exploration of NBPs and aptamers offers a promising horizon for revolutionizing cancer therapy and diagnostics. Through their unparalleled specificity and affinity for target molecules, indicated in (Table 4), NBPs hold the potential to redefine precision medicine, enabling the development of highly effective, minimally invasive diagnostic tools and treatments. Despite significant advancements, the intricate mechanisms governing NBP interactions within the vast biological milieu remain partially understood, presenting a formidable barrier to their clinical adoption. Addressing this knowledge gap necessitates a multidisciplinary approach, integrating advanced bioinformatics, structural biology, and nanotechnology. As we delve deeper into the molecular intricacies of NBPs, the future of oncology and diagnostic medicine stands on the brink of a new era, promising more personalized, accurate, and effective healthcare solutions.
Ongoing research in this field is key to driving forward the evolution of cancer therapy. By delving deeper into the roles and functionalities of binding domains, there is a significant potential to transform cancer treatment paradigms and ultimately improve patient survival rates. This pursuit of knowledge in the realm of NBPs and their related domains is a crucial step towards a future where cancer treatment is more efficient, precise, and tailored to individual patient needs.

Author Contributions

Conceptualization, M.E. and S.P.S.; methodology, M.E.; writing—original draft preparation, M.E. and K.F.M.L.; writing—review and editing, M.E., K.F.M.L. and S.P.S.; visualization, M.E. and K.F.M.L.; supervision, S.P.S.; project administration, M.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new experimental data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We express our gratitude to the OeAD for their support through the scholarship (MPC-2023-00335) awarded to Marco Eigenfeld. Furthermore we express our gratitude to ERASMUS+ for their support of Kilian F.M. Lupp (2023-1-DE01-KA131-HED-000120711).

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. Classification of carbohydrate-binding domains according to structural and functional properties [144]; however, to date, there remain over 6000 domains that are still unclassified [143].
Table A1. Classification of carbohydrate-binding domains according to structural and functional properties [144]; however, to date, there remain over 6000 domains that are still unclassified [143].
FamilyProtein FoldDemonstrated Binding Specificities
CBM1Cysteine knotCellulose (chitin one case)
CBM2 β -sandwichCellulose, chitin, xylan
CBM3 β -sandwichCellulose and chitin
CBM4 β -sandwichXylan, β -1,3-glucan, β -1,3-1,4-glucan, β -1,6-glucan, and amorphous cellulose
CBM5UniqueChitin
CBM6 β -sandwichAmorphous cellulose, β -1,4-xylan, β -1,3-glucan, β -1,3-1,4-glucan, and β -1,4-glucan
CBM7Deleted
CBM8UnknownCellulose
CBM9 β -sandwichCellulose
CBM10OB foldCellulose
CBM11 β -sandwich β -1,4-glucan and β -1,3-1,4-mixed-linked glucans
CBM12UniqueChitin
CBM13 β -trefoilMannose, xylan, N-acetylgalactosamine
CBM14UniqueChitin
CBM15 β -sandwichXylan and xylooligosaccharides
CBM16 β -sandwichCellulose and glucomannan
CBM17 β -sandwichAmorphous cellulose, cellooligosaccharides, and derivatized cellulose
CBM18Hevein foldChitin
CBM19UnknownChitin
CBM20 β -sandwichGranular starch, cyclodextrines
CBM21 β -sandwichStarch
CBM22 β -sandwichXylan, β -1,3/ β -1,4-glucans
CBM23UnknownMannan
CBM24Unknown α -1,3-glucan
CBM25 β -sandwichStarch
CBM26 β -sandwichStarch
CBM27 β -sandwichMannan
CBM28 β -sandwichNoncrystalline cellulose, cello-oligosaccharides, and β -(1,3)(1,4)-glucans
CBM29 β -sandwichMannan and glucomannan
CBM30 β -sandwichCellulose
CBM31 β -sandwich β -1,3-xylan
CBM32 β -sandwichGalactose, lactose, polygalacturonic acid, β -D-galactosyl-1,4- β -D-N-acetylglucosamine
CBM33 β -sandwichChitin and chitosan
CBM34 β -sandwichGranular starch
CBM35 β -sandwich4,5-deoxygalaturonic acid, glucuronic acid, xylan, β -galactan
CBM36 β -sandwichXylan and xylooligosaccharides
CBM37UnknownXylan, chitin, microcrystalline and phosphoric acid-swollen cellulose, alfalfa cell walls, banana stem, and wheat straw
CBM38UnknownInulin
CBM39 β -sandwich β -1,3-glucan, lipopolysaccharide, and lipoteichoic acid
CBM40 β -sandwichSialic acid
CBM41 β -sandwichAmylose, amylopectin, pullulan, and α -glucan oligosaccharide fragments
CBM42 β -trefoilArabinofuranose
CBM43CtD-Ole e 9 β -1,3-glucan
CBM44 β -sandwichCellulose and xyloglucan
CBM45UnknownStarch
CBM46UnknownCellulose
CBM47 β -sandwichFucose
CBM48 β -sandwichGlycogen
CBM49UnknownCellulose
CBM50LysM-domainChitopentaose
CBM51 β -sandwichGalactose and to blood group A/B-antigens
CBM52Unknown β -1,3-glucan
CBM53UnknownStarch
CBM54UnknownXylan, yeast cell wall glucan, and chitin
CBM55UnknownChitin
CBM56Unknown β -1,3-glucan
CBM57 β -sandwichGlucose oligomers
CBM58 β -sandwichMaltoheptaose
CBM59 β -sandwichMannan, xylan, and cellulose
CBM60 β -sandwichXylan
CBM61 β -sandwich β -1,4-galactan
CBM62 β -sandwichGalactose moieties found on xyloglucan, arabinogalactan, and galactomannan
CBM63Expansin-likeCellulose
CBM64UnknownCellulose
CBM65 β -sandwich β -glucan, xyloglucan
CBM66 β -sandwichFructans
CBM67Multidomain structureL-rhamnose
CBM68UnknownMaltotriose, maltotetraose
CBM69UnknownStarch
CBM70 β -sandwichHyaluronan
CBM71 β -sandwichLactose, LacNAc
CBM72UnknownVarious polysaccharides, including cellulose, β -1,3/1,4-mixed linked glucans, xylan, and β -mannan
CBM73 β -sheet containing structureChitin
CBM74UnknownStarch
CBM75UnknownXyloglucan
CBM76Unknown β -glucan, xyloglucan, glucomannan
CBM77 β -sandwichPectin
CBM78 β -sandwichDecorated β -glucans, xyloglucan
CBM79 β -sandwich β -glucans
CBM80 β -sandwichXylocglucan, glucomannan, galactomannan, barley β -glucan
CBM81 β -sandwich β -1,4-, β -1,3-glucans, xyloglucan, avicel, cellooligosaccharides
CBM82UnknownStarch
CBM83UnknownStarch
CBM84UnknownXanthan
CBM85UnknownCellulose, glucuronoxylan, β -1,3-1,4-glucan, and glucomannan
CBM86 β -sandwichXylan
CBM87 α - β - α -domain α -1,4-N-acetylgalactosamine-rich regions of galactosaminogalactan
CBM88UnknownTerminal galactose in galactoxyloglucan and galactomannan
CBM89 β -helixBeechwood xylan and rye arabinoxylan binding
CBM90UnknownUlvan
CBM91UnknownXylans (birchwood and oat spelt)
CBM92Unknown β -1,3- and β -1,6-glucan
CBM93UnknownGlycan
CBM94UnknownN-Acetylglucosamine
CBM95UnknownPectic rhamnogalacturonan-I
CBM96UnknownAlginate
CBM97UnknownPolygalacturonic acid
CBM98UnknownAmylopectin
CBM99UnknownPorphyran
CBM100 β -sandwichChondroitin sulfate
CBM101 β -sandwichAgarose

References

  1. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer statistics, 2023. CA Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef] [PubMed]
  2. WHO. Breast Cancer. 2024. Available online: https://www.who.int/news-room/fact-sheets/detail/breast-cancer (accessed on 5 February 2024).
  3. Łukasiewicz, S.; Czeczelewski, M.; Forma, A.; Baj, J.; Sitarz, R.; Stanisławek, A. Breast Cancer-Epidemiology, Risk Factors, Classification, Prognostic Markers, and Current Treatment Strategies-An Updated Review. Cancers 2021, 13, 4287. [Google Scholar] [CrossRef] [PubMed]
  4. Corbex, M.; Burton, R.; Sancho-Garnier, H. Breast cancer early detection methods for low and middle income countries, a review of the evidence. Breast 2012, 21, 428–434. [Google Scholar] [CrossRef] [PubMed]
  5. Waks, A.G.; Winer, E.P. Breast Cancer Treatment: A Review. JAMA 2019, 321, 288–300. [Google Scholar] [CrossRef]
  6. American Cancer Society. Radiation Therapy Side Effects. 2023. Available online: https://www.cancer.org/cancer/managing-cancer/treatment-types/radiation/effects-on-different-parts-of-body.html (accessed on 7 May 2024).
  7. Swain, S.M.; Miles, D.; Kim, S.B.; Im, Y.H.; Im, S.A.; Semiglazov, V.; Ciruelos, E.; Schneeweiss, A.; Loi, S.; Monturus, E.; et al. Pertuzumab, trastuzumab, and docetaxel for HER2-positive metastatic breast cancer (CLEOPATRA): End-of-study results from a double-blind, randomised, placebo-controlled, phase 3 study. Lancet Oncol. 2020, 21, 519–530. [Google Scholar] [CrossRef] [PubMed]
  8. National Cancer Institute at the National Institutes of Health. Trastuzumab. 2006. Available online: https://www.cancer.gov/about-cancer/treatment/drugs/trastuzumab (accessed on 5 February 2024).
  9. Patel, A.; Unni, N.; Peng, Y. The Changing Paradigm for the Treatment of HER2-Positive Breast Cancer. Cancers 2020, 12, 2081. [Google Scholar] [CrossRef] [PubMed]
  10. Rugo, H.S.; Im, S.A.; Cardoso, F.; Cortes, J.; Curigliano, G.; Musolino, A.; Pegram, M.D.; Bachelot, T.; Wright, G.S.; Saura, C.; et al. Margetuximab Versus Trastuzumab in Patients With Previously Treated HER2-Positive Advanced Breast Cancer (SOPHIA): Final Overall Survival Results From a Randomized Phase 3 Trial. J. Clin. Oncol. 2022, 41, 198–205. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, R.; Pan, W.; Jin, L.; Huang, W.; Li, Y.; Wu, D.; Gao, C.; Ma, D.; Liao, S. Human papillomavirus vaccine against cervical cancer: Opportunity and challenge. Cancer Lett. 2020, 471, 88–102. [Google Scholar] [CrossRef]
  12. Papahadjopoulos, D.; Poste, G.; Schaeffer, B.E. Fusion of mammalian cells by unilamellar lipid vesicles: Influence of lipid surface charge, fluidity and cholesterol. Biochim. Biophys. Acta Biomembr. 1973, 323, 23–42. [Google Scholar] [CrossRef]
  13. Adams, D.H.; Joyce, G.; Richardson, V.J.; Ryman, B.E.; Wiśniewski, H.M. Liposome toxicity in the mouse central nervous system. J. Neurol. Sci. 1977, 31, 173–179. [Google Scholar] [CrossRef]
  14. Shi, M.; Anantha, M.; Wehbe, M.; Bally, M.B.; Fortin, D.; Roy, L.O.; Charest, G.; Richer, M.; Paquette, B.; Sanche, L. Liposomal formulations of carboplatin injected by convection-enhanced delivery increases the median survival time of F98 glioma bearing rats. J. Nanobiotechnol. 2018, 16, 77. [Google Scholar] [CrossRef] [PubMed]
  15. Fan, Y.; Sahdev, P.; Ochyl, L.J.; Akerberg, J.J.; Moon, J.J. Cationic liposome–hyaluronic acid hybrid nanoparticles for intranasal vaccination with subunit antigens. J. Control. Release 2015, 208, 121–129. [Google Scholar] [CrossRef] [PubMed]
  16. Mehta, M.; Bui, T.A.; Yang, X.; Aksoy, Y.; Goldys, E.M.; Deng, W. Lipid-Based Nanoparticles for Drug/Gene Delivery: An Overview of the Production Techniques and Difficulties Encountered in Their Industrial Development. ACS Mater. Au 2023, 3, 600–619. [Google Scholar] [CrossRef] [PubMed]
  17. Sun, L.; Liu, H.; Ye, Y.; Lei, Y.; Islam, R.; Tan, S.; Tong, R.; Miao, Y.B.; Cai, L. Smart nanoparticles for cancer therapy. Signal Transduct. Target. Ther. 2023, 8, 418. [Google Scholar] [CrossRef] [PubMed]
  18. Mitchell, M.J.; Billingsley, M.M.; Haley, R.M.; Wechsler, M.E.; Peppas, N.A.; Langer, R. Engineering precision nanoparticles for drug delivery. Nat. Rev. Drug Discov. 2021, 20, 101–124. [Google Scholar] [CrossRef] [PubMed]
  19. Hong, S.; Choi, D.W.; Kim, H.N.; Park, C.G.; Lee, W.; Park, H.H. Protein-Based Nanoparticles as Drug Delivery Systems. Pharmaceutics 2020, 12, 604. [Google Scholar] [CrossRef] [PubMed]
  20. Khramtsov, P.; Kalashnikova, T.; Bochkova, M.; Kropaneva, M.; Timganova, V.; Zamorina, S.; Rayev, M. Measuring the concentration of protein nanoparticles synthesized by desolvation method: Comparison of Bradford assay, BCA assay, hydrolysis/UV spectroscopy and gravimetric analysis. Int. J. Pharm. 2021, 599, 120422. [Google Scholar] [CrossRef] [PubMed]
  21. Shome, A.; Rather, A.M.; Manna, U. Chemically reactive protein nanoparticles for synthesis of a durable and deformable superhydrophobic material. Nanoscale Adv. 2019, 1, 1746–1753. [Google Scholar] [CrossRef]
  22. Turrina, C.; Klassen, A.; Milani, D.; Rojas-González, D.M.; Ledinski, G.; Auer, D.; Sartori, B.; Cvirn, G.; Mela, P.; Berensmeier, S.; et al. Superparamagnetic iron oxide nanoparticles for their application in the human body: Influence of the surface. Heliyon 2023, 9, e16487. [Google Scholar] [CrossRef]
  23. Barenholz, Y. (Chezy) Doxil®—The first FDA-approved nano-drug: Lessons learned. J. Control. Release 2012, 160, 117–134. [Google Scholar] [CrossRef]
  24. Veiseh, O.; Gunn, J.W.; Zhang, M. Design and fabrication of magnetic nanoparticles for targeted drug delivery and imaging. Adv. Drug Deliv. Rev. 2010, 62, 284–304. [Google Scholar] [CrossRef]
  25. Avval, Z.M.; Malekpour, L.; Raeisi, F.; Babapoor, A.; Mousavi, S.M.; Hashemi, S.A.; Salari, M. Introduction of magnetic and supermagnetic nanoparticles in new approach of targeting drug delivery and cancer therapy application. Drug Metab. Rev. 2020, 52, 157–184. [Google Scholar] [CrossRef]
  26. Hernandes, E.P.; Lazarin-Bidóia, D.; Bini, R.D.; Nakamura, C.V.; Cótica, L.F.; de Oliveira Silva Lautenschlager, S. Doxorubicin-Loaded Iron Oxide Nanoparticles Induce Oxidative Stress and Cell Cycle Arrest in Breast Cancer Cells. Antioxidants 2023, 12, 237. [Google Scholar] [CrossRef]
  27. Dilnawaz, F.; Singh, A.; Mohanty, C.; Sahoo, S.K. Dual drug loaded superparamagnetic iron oxide nanoparticles for targeted cancer therapy. Biomaterials 2010, 31, 3694–3706. [Google Scholar] [CrossRef] [PubMed]
  28. Amsalem, O.; Nassar, T.; Benhamron, S.; Lazarovici, P.; Benita, S.; Yavin, E. Solid nano-in-nanoparticles for potential delivery of siRNA. J. Control. Release 2017, 257, 144–155. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, H.; Wang, H.; Wen, C.; Bai, S.; Wei, P.; Xu, B.; Xu, Y.; Liang, C.; Zhang, Y.; Zhang, G.; et al. Effects of iron oxide nanoparticles as T2-MRI contrast agents on reproductive system in male mice. J. Nanobiotechnol. 2022, 20, 98. [Google Scholar] [CrossRef]
  30. Oberdick, S.D.; Jordanova, K.V.; Lundstrom, J.T.; Parigi, G.; Poorman, M.E.; Zabow, G.; Keenan, K.E. Iron oxide nanoparticles as positive T1 contrast agents for low-field magnetic resonance imaging at 64 mT. Sci. Rep. 2023, 13, 11520. [Google Scholar] [CrossRef]
  31. Peng, Y.K.; Tsang, S.C.E.; Chou, P.T. Chemical design of nanoprobes for T1-weighted magnetic resonance imaging. Mater. Today 2016, 19, 336–348. [Google Scholar] [CrossRef]
  32. Chiarelli, P.A.; Revia, R.A.; Stephen, Z.R.; Wang, K.; Jeon, M.; Nelson, V.; Kievit, F.M.; Sham, J.; Ellenbogen, R.G.; Kiem, H.P.; et al. Nanoparticle Biokinetics in Mice and Nonhuman Primates. ACS Nano 2017, 11, 9514–9524. [Google Scholar] [CrossRef]
  33. Khandhar, A.P.; Ferguson, R.M.; Arami, H.; Krishnan, K.M. Monodisperse magnetite nanoparticle tracers for in vivo magnetic particle imaging. Biomaterials 2013, 34, 3837–3845. [Google Scholar] [CrossRef]
  34. Weber, C.; Coester, C.; Kreuter, J.; Langer, K. Desolvation process and surface characterisation of protein nanoparticles. Int. J. Pharm. 2000, 194, 91–102. [Google Scholar] [CrossRef]
  35. Forest, V.; Cottier, M.; Pourchez, J. Electrostatic interactions favor the binding of positive nanoparticles on cells: A reductive theory. Nano Today 2015, 10, 677–680. [Google Scholar] [CrossRef]
  36. Yang, X.; Zhang, W.; Jiang, W.; Kumar, A.; Zhou, S.; Cao, Z.; Zhan, S.; Yang, W.; Liu, R.; Teng, Y.; et al. Nanoconjugates to enhance PDT-mediated cancerimmunotherapy by targeting the indoleamine-2,3-dioxygenase pathway. J. Nanobiotechnol. 2021, 19, 182. [Google Scholar] [CrossRef] [PubMed]
  37. Shaban, M.; Hasanzadeh, M.; Solhi, E. An Fe3O4/PEDOT:PSS nanocomposite as an advanced electroconductive material for the biosensing of the prostate-specific antigen in unprocessed human plasma samples. Anal. Methods 2019, 11, 5661–5672. [Google Scholar] [CrossRef]
  38. Chomoucka, J.; Drbohlavova, J.; Huska, D.; Adam, V.; Kizek, R.; Hubalek, J. Magnetic nanoparticles and targeted drug delivering. Pharmacol. Res. 2010, 62, 144–149. [Google Scholar] [CrossRef] [PubMed]
  39. Bobo, D.; Robinson, K.J.; Islam, J.; Thurecht, K.J.; Corrie, S.R. Nanoparticle-Based Medicines: A Review of FDA-Approved Materials and Clinical Trials to Date. Pharm. Res. 2016, 33, 2373–2387. [Google Scholar] [CrossRef] [PubMed]
  40. Ranade, A.A.; Joshi, D.A.; Phadke, G.K.; Patil, P.P.; Kasbekar, R.B.; Apte, T.G.; Dasare, R.R.; Mengde, S.D.; Parikh, P.M.; Bhattacharyya, G.S.; et al. Clinical and economic implications of the use of nanoparticle paclitaxel (Nanoxel) in India. Ann. Oncol. 2013, 24, v6–v12. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, J.; Li, X.; Wu, W.; Xu, X.M.; Xu, H.; Zhang, T. Recent Progress of Paclitaxel Delivery Systems: Covalent and Noncovalent Approaches. Adv. Ther. 2023, 6, 2200281. [Google Scholar] [CrossRef]
  42. Mahmoudi, K.; Bouras, A.; Bozec, D.; Ivkov, R.; Hadjipanayis, C. Magnetic hyperthermia therapy for the treatment of glioblastoma: A review of the therapy’s history, efficacy and application in humans. Int. J. Hyperth. 2018, 34, 1316–1328. [Google Scholar] [CrossRef]
  43. Makita, M.; Manabe, E.; Kurita, T.; Takei, H.; Nakamura, S.; Kuwahata, A.; Sekino, M.; Kusakabe, M.; Ohashi, Y. Moving a neodymium magnet promotes the migration of a magnetic tracer and increases the monitoring counts on the skin surface of sentinel lymph nodes in breast cancer. BMC Med. Imaging 2020, 20, 58. [Google Scholar] [CrossRef]
  44. Jain, A.; Singh, S.K.; Arya, S.K.; Kundu, S.C.; Kapoor, S. Protein Nanoparticles: Promising Platforms for Drug Delivery Applications. ACS Biomater. Sci. Eng. 2018, 4, 3939–3961. [Google Scholar] [CrossRef] [PubMed]
  45. Yao, Y.; Zhou, Y.; Liu, L.; Xu, Y.; Chen, Q.; Wang, Y.; Wu, S.; Deng, Y.; Zhang, J.; Shao, A. Nanoparticle-Based Drug Delivery in Cancer Therapy and Its Role in Overcoming Drug Resistance. Front. Mol. Biosci. 2020, 7, 193. [Google Scholar] [CrossRef]
  46. Parodi, A.; Molinaro, R.; Sushnitha, M.; Evangelopoulos, M.; Martinez, J.O.; Arrighetti, N.; Corbo, C.; Tasciotti, E. Bio-inspired engineering of cell- and virus-like nanoparticles for drug delivery. Biomaterials 2017, 147, 155–168. [Google Scholar] [CrossRef]
  47. Silva, R.C.; Lourenço, B.G.; Ulhoa, P.H.F.; Dias, E.A.F.; da Cunha, F.L.; Tonetto, C.P.; Villani, L.G.; Vimieiro, C.B.S.; Lepski, G.A.; Monjardim, M.; et al. Biomimetic Design of a Tendon-Driven Myoelectric Soft Hand Exoskeleton for Upper-Limb Rehabilitation. Biomimetics 2023, 8, 317. [Google Scholar] [CrossRef] [PubMed]
  48. Abdelhafiz, M.H.; Andreasen Struijk, L.N.S.; Dosen, S.; Spaich, E.G. Biomimetic Tendon-Based Mechanism for Finger Flexion and Extension in a Soft Hand Exoskeleton: Design and Experimental Assessment. Sensors 2023, 23, 2272. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, Y.; Jin, S.; Luo, D.; He, D.; Shi, C.; Zhu, L.; Guan, B.; Li, Z.; Zhang, T.; Zhou, Y.; et al. Functional regeneration and repair of tendons using biomimetic scaffolds loaded with recombinant periostin. Nat. Commun. 2021, 12, 1293. [Google Scholar] [CrossRef] [PubMed]
  50. Xie, R.; Yao, H.; Mao, A.S.; Zhu, Y.; Qi, D.; Jia, Y.; Gao, M.; Chen, Y.; Wang, L.; Wang, D.A.; et al. Biomimetic cartilage-lubricating polymers regenerate cartilage in rats with early osteoarthritis. Nat. Biomed. Eng. 2021, 5, 1189–1201. [Google Scholar] [CrossRef]
  51. Zafar, M.S.; Amin, F.; Fareed, M.A.; Ghabbani, H.; Riaz, S.; Khurshid, Z.; Kumar, N. Biomimetic Aspects of Restorative Dentistry Biomaterials. Biomimetics 2020, 5, 34. [Google Scholar] [CrossRef]
  52. Palazzo, B.; Iafisco, M.; Laforgia, M.; Margiotta, N.; Natile, G.; Bianchi, C.; Walsh, D.; Mann, S.; Roveri, N. Biomimetic Hydroxyapatite–Drug Nanocrystals as Potential Bone Substitutes with Antitumor Drug Delivery Properties. Adv. Funct. Mater. 2007, 17, 2180–2188. [Google Scholar] [CrossRef]
  53. Rao, L.; Bu, L.L.; Xu, J.H.; Cai, B.; Yu, G.T.; Yu, X.; He, Z.; Huang, Q.; Li, A.; Guo, S.S.; et al. Red Blood Cell Membrane as a Biomimetic Nanocoating for Prolonged Circulation Time and Reduced Accelerated Blood Clearance. Small 2015, 11, 6225–6236. [Google Scholar] [CrossRef]
  54. Sushnitha, M.; Evangelopoulos, M.; Tasciotti, E.; Taraballi, F. Cell Membrane-Based Biomimetic Nanoparticles and the Immune System: Immunomodulatory Interactions to Therapeutic Applications. Front. Bioeng. Biotechnol. 2020, 8, 627. [Google Scholar] [CrossRef]
  55. Naahidi, S.; Jafari, M.; Edalat, F.; Raymond, K.; Khademhosseini, A.; Chen, P. Biocompatibility of engineered nanoparticles for drug delivery. J. Control. Release 2013, 166, 182–194. [Google Scholar] [CrossRef]
  56. Zhang, X.; Li, J.; Ma, C.; Zhang, H.; Liu, K. Biomimetic Structural Proteins: Modular Assembly and High Mechanical Performance. Accounts Chem. Res. 2023, 56, 2664–2675. [Google Scholar] [CrossRef]
  57. Lavickova, B.; Laohakunakorn, N.; Maerkl, S.J. A partially self-regenerating synthetic cell. Nat. Commun. 2020, 11, 6340. [Google Scholar] [CrossRef] [PubMed]
  58. Xu, X.; Chen, X.; Li, J. Natural protein bioinspired materials for regeneration of hard tissues. J. Mater. Chem. B 2020, 8, 2199–2215. [Google Scholar] [CrossRef]
  59. Bernaudat, F.; Frelet-Barrand, A.; Pochon, N.; Dementin, S.; Hivin, P.; Boutigny, S.; Rioux, J.B.; Salvi, D.; Seigneurin-Berny, D.; Richaud, P.; et al. Heterologous Expression of Membrane Proteins: Choosing the Appropriate Host. PLoS ONE 2011, 6, e29191. [Google Scholar] [CrossRef] [PubMed]
  60. Noad, R.; Roy, P. Virus-like particles as immunogens. Trends Microbiol. 2003, 11, 438–444. [Google Scholar] [CrossRef] [PubMed]
  61. Rohovie, M.J.; Nagasawa, M.; Swartz, J.R. Virus-like particles: Next-generation nanoparticles for targeted therapeutic delivery. Bioeng. Transl. Med. 2017, 2, 43–57. [Google Scholar] [CrossRef]
  62. Sari-Ak, D.; Bahrami, S.; Laska, M.J.; Drncova, P.; Fitzgerald, D.J.; Schaffitzel, C.; Garzoni, F.; Berger, I. High-Throughput Production of Influenza Virus-Like Particle (VLP) Array by Using VLP-factory™, a MultiBac Baculoviral Genome Customized for Enveloped VLP Expression. In High-Throughput Protein Production and Purification: Methods and Protocols; Vincentelli, R., Ed.; Springer: New York, NY, USA, 2019; pp. 213–226. [Google Scholar] [CrossRef]
  63. Brillault, L.; Jutras, P.V.; Dashti, N.; Thuenemann, E.C.; Morgan, G.; Lomonossoff, G.P.; Landsberg, M.J.; Sainsbury, F. Engineering Recombinant Virus-like Nanoparticles from Plants for Cellular Delivery. ACS Nano 2017, 11, 3476–3484. [Google Scholar] [CrossRef]
  64. Huo, Y.; Wan, X.; Ling, T.; Wu, J.; Wang, W.; Shen, S. Expression and purification of norovirus virus like particles in Escherichia coli and their immunogenicity in mice. Mol. Immunol. 2018, 93, 278–284. [Google Scholar] [CrossRef]
  65. Wetzel, D.; Rolf, T.; Suckow, M.; Kranz, A.; Barbian, A.; Chan, J.A.; Leitsch, J.; Weniger, M.; Jenzelewski, V.; Kouskousis, B.; et al. Establishment of a yeast-based VLP platform for antigen presentation. Microb. Cell Factories 2018, 17, 17. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, C.; Zhang, J.; Yin, J.; Gan, Y.; Xu, S.; Gu, Y.; Huang, W. Alternative approaches to target Myc for cancer treatment. Signal Transduct. Target. Ther. 2021, 6, 117. [Google Scholar] [CrossRef] [PubMed]
  67. Lambert, M.; Jambon, S.; Depauw, S.; David-Cordonnier, M.H. Targeting Transcription Factors for Cancer Treatment. Molecules 2018, 23, 1479. [Google Scholar] [CrossRef] [PubMed]
  68. Corsico, B.; Cistola, D.P.; Frieden, C.; Storch, J. The helical domain of intestinal fatty acid binding protein is critical for collisional transfer of fatty acids to phospholipid membranes. Proc. Natl. Acad. Sci. USA 1998, 95, 12174–12178. [Google Scholar] [CrossRef] [PubMed]
  69. Boraston, A.B.; Bolam, D.N.; Gilbert, H.J.; Davies, G.J. Carbohydrate-binding modules: Fine-tuning polysaccharide recognition. Biochem. J. 2004, 382, 769–781. [Google Scholar] [CrossRef] [PubMed]
  70. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Molecular Biology of the Cell, 4th ed.; Garland Science: New York, NY, USA, 2002. [Google Scholar]
  71. Hashimoto, M.; Ikegami, T.; Seino, S.; Ohuchi, N.; Fukada, H.; Sugiyama, J.; Shirakawa, M.; Watanabe, T. Expression and Characterization of the Chitin-Binding Domain of Chitinase A1 from Bacillus circulans WL-12. J. Bacteriol. 2000, 182, 3045–3054. [Google Scholar] [CrossRef] [PubMed]
  72. Nampally, M.; Moerschbacher, B.M.; Kolkenbrock, S. Fusion of a Novel Genetically Engineered Chitosan Affinity Protein and Green Fluorescent Protein for Specific Detection of Chitosan In Vitro and In Situ. Appl. Environ. Microbiol. 2012, 78, 3114–3119. [Google Scholar] [CrossRef]
  73. Hudson, W.H.; Ortlund, E.A. The structure, function and evolution of proteins that bind DNA and RNA. Nat. Rev. Mol. Cell Biol. 2014, 15, 749–760. [Google Scholar] [CrossRef] [PubMed]
  74. Smith, M.R.; Costa, G. RNA-binding proteins and translation control in angiogenesis. FEBS J. 2022, 289, 7788–7809. [Google Scholar] [CrossRef]
  75. Eigenfeld, M.; Kerpes, R.; Becker, T. Recombinant protein linker production as a basis for non-invasive determination of single-cell yeast age in heterogeneous yeast populations. RSC Adv. 2021, 11, 31923–31932. [Google Scholar] [CrossRef]
  76. Vogt, S.; Kelkenberg, M.; Nöll, T.; Steinhoff, B.; Schönherr, H.; Merzendorfer, H.; Nöll, G. Rapid determination of binding parameters of chitin binding domains using chitin-coated quartz crystal microbalance sensor chips. Analyst 2018, 143, 5255–5263. [Google Scholar] [CrossRef] [PubMed]
  77. Azuma, K.; Osaki, T.; Minami, S.; Okamoto, Y. Anticancer and Anti-Inflammatory Properties of Chitin and Chitosan Oligosaccharides. J. Funct. Biomater. 2015, 6, 33–49. [Google Scholar] [CrossRef] [PubMed]
  78. Zhou, X.; Liu, D.; Liu, H.; Yang, Q.; Yao, K.; Wang, X.; Wang, L.; Yang, X. Effect of Low Molecular Weight Chitosans on Drug Permeation through Mouse Skin: 1. Transdermal Delivery of Baicalin. J. Pharm. Sci. 2010, 99, 2991–2998. [Google Scholar] [CrossRef] [PubMed]
  79. Zheng, B.; Wen, Z.S.; Huang, Y.J.; Xia, M.S.; Xiang, X.W.; Qu, Y.L. Molecular Weight-Dependent Immunostimulative Activity of Low Molecular Weight Chitosan via Regulating NF-kB and AP-1 Signaling Pathways in RAW264.7 Macrophages. Mar. Drugs 2016, 14, 169. [Google Scholar] [CrossRef] [PubMed]
  80. Madu, C.O.; Lu, Y. Novel diagnostic biomarkers for prostate cancer. J. Cancer 2010, 1, 150–177. [Google Scholar] [CrossRef] [PubMed]
  81. Zheng, X.; Liu, X.; Lei, Y.; Wang, G.; Liu, M. Glypican-3: A Novel and Promising Target for the Treatment of Hepatocellular Carcinoma. Front. Oncol. 2022, 12, 824208. [Google Scholar] [CrossRef]
  82. Janski, N.; Masoud, K.; Batzenschlager, M.; Herzog, E.; Evrard, J.L.; Houlné, G.; Bourge, M.; Chabouté, M.E.; Schmit, A.C. The GCP3-interacting proteins GIP1 and GIP2 are required for γ-tubulin complex protein localization, spindle integrity, and chromosomal stability. Plant Cell 2012, 24, 1171–1187. [Google Scholar] [CrossRef]
  83. Aggarwal, D.; Yang, J.; Salam, M.A.; Sengupta, S.; Al-Amin, M.Y.; Mustafa, S.; Khan, M.A.; Huang, X.; Pawar, J.S. Antibody-drug conjugates: The paradigm shifts in the targeted cancer therapy. Front. Immunol 2023, 14, 1203073. [Google Scholar] [CrossRef] [PubMed]
  84. Rassy, E.; Heard, J.M.; Andre, F. The paradigm shift to precision oncology between political will and cultural acceptance. ESMO Open 2023, 8, 101622. [Google Scholar] [CrossRef]
  85. Rhee, S.; Martin, R.G.; Rosner, J.L.; Davies, D.R. A novel DNA-binding motif in MarA: The first structure for an AraC family transcriptional activator. Proc. Natl. Acad. Sci. USA 1998, 95, 10413–10418. [Google Scholar] [CrossRef]
  86. Gonzalez, D.H. Chapter 1—Introduction to Transcription Factor Structure and Function. In Plant Transcription Factors; Gonzalez, D.H., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 3–11. [Google Scholar] [CrossRef]
  87. Corbella, M.; Liao, Q.; Moreira, C.; Parracino, A.; Kasson, P.M.; Kamerlin, S.C.L. The N-terminal Helix-Turn-Helix Motif of Transcription Factors MarA and Rob Drives DNA Recognition. J. Phys. Chem. B 2021, 125, 6791–6806. [Google Scholar] [CrossRef] [PubMed]
  88. McColl, D.J.; Honchell, C.D.; Frankel, A.D. Structure-based design of an RNA-binding zinc finger. Proc. Natl. Acad. Sci. USA 1999, 96, 9521–9526. [Google Scholar] [CrossRef] [PubMed]
  89. Ransom, M.; Dennehey, B.K.; Tyler, J.K. Chaperoning Histones during DNA Replication and Repair. Cell 2010, 140, 183–195. [Google Scholar] [CrossRef] [PubMed]
  90. Stracy, M.; Schweizer, J.; Sherratt, D.J.; Kapanidis, A.N.; Uphoff, S.; Lesterlin, C. Transient non-specific DNA binding dominates the target search of bacterial DNA-binding proteins. Mol. Cell 2021, 81, 1499–1514.e6. [Google Scholar] [CrossRef] [PubMed]
  91. Keith, J.M. Bioinformatics; Book Section 7—The Classification of Protein Domains; Springer: New York, NY, USA, 2017; pp. 137–164. [Google Scholar]
  92. Charoensawan, V.; Wilson, D.; Teichmann, S.A. Genomic repertoires of DNA-binding transcription factors across the tree of life. Nucleic Acids Res. 2010, 38, 7364–7377. [Google Scholar] [CrossRef] [PubMed]
  93. Wingender, E. Available online: http://gene-regulation.com/ (accessed on 29 April 2024).
  94. Liptak, C.; Loria, J.P. Movement and Specificity in a Modular DNA Binding Protein. Structure 2015, 23, 973–974. [Google Scholar] [CrossRef] [PubMed]
  95. Bochkarev, A.; Barwell, J.A.; Pfuetzner, R.A.; Furey, W.; Edwards, A.M.; Frappier, L. Crystal structure of the DNA-binding domain of the Epstein-Barr virus origin-binding protein EBNA1. Cell 1995, 83, 39–46. [Google Scholar] [CrossRef] [PubMed]
  96. Bochkarev, A.; Barwell, J.A.; Pfuetzner, R.A.; Bochkareva, E.; Frappier, L.; Edwards, A.M. Crystal Structure of the DNA-Binding Domain of the Epstein-Barr Virus Origin-Binding Protein, EBNA1, Bound to DNA. Cell 1996, 84, 791–800. [Google Scholar] [CrossRef]
  97. Schleif, R. DNA Binding by Proteins. Science 1988, 241, 1182–1187. [Google Scholar] [CrossRef]
  98. Chaible, L.M.; Kinoshita, D.; Finzi Corat, M.A.; Zaidan Dagli, M.L. Chapter 27—Genetically Modified Animal Models. In Animal Models for the Study of Human Disease, 2nd ed.; Conn, P.M., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 703–726. [Google Scholar] [CrossRef]
  99. Suzuki, T.; Kimura, A.; Nagai, R.; Horikoshi, M. Regulation of interaction of the acetyltransferase region of p300 and the DNA-binding domain of Sp1 on and through DNA binding. Genes Cells 2002, 5, 29–41. [Google Scholar] [CrossRef]
  100. Grove, A.; Lim, L. High-affinity DNA binding of HU protein from the hyperthermophile Thermotoga maritima11Edited by T. Richmond. J. Mol. Biol. 2001, 311, 491–502. [Google Scholar] [CrossRef] [PubMed]
  101. Yang, Y.; Sass, L.E.; Du, C.; Hsieh, P.; Erie, D.A. Determination of protein-DNA binding constants and specificities from statistical analyses of single molecules: MutS-DNA interactions. Nucleic Acids Res. 2005, 33, 4322–4334. [Google Scholar] [CrossRef] [PubMed]
  102. Radaeva, M.; Ton, A.T.; Hsing, M.; Ban, F.; Cherkasov, A. Drugging the ‘undruggable’. Therapeutic targeting of protein-DNA interactions with the use of computer-aided drug discovery methods. Drug Discov. Today 2021, 26, 2660–2679. [Google Scholar] [CrossRef] [PubMed]
  103. Chahrour, M.; Zoghbi, H.Y. The Story of Rett Syndrome: From Clinic to Neurobiology. Neuron 2007, 56, 422–437. [Google Scholar] [CrossRef] [PubMed]
  104. Islam, Z.; Ali, A.M.; Naik, A.; Eldaw, M.; Decock, J.; Kolatkar, P.R. Transcription Factors: The Fulcrum Between Cell Development and Carcinogenesis. Front. Oncol. 2021, 11, 681377. [Google Scholar] [CrossRef] [PubMed]
  105. Bushweller, J.H. Targeting transcription factors in cancer—From undruggable to reality. Nat. Rev. Cancer 2019, 19, 611–624. [Google Scholar] [CrossRef] [PubMed]
  106. Herz, H.M.; Hu, D.; Shilatifard, A. Enhancer Malfunction in Cancer. Mol. Cell 2014, 53, 859–866. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, Z.; Xue, S.t.; Gao, Y.; Li, Y.; Zhou, Z.; Wang, J.; Li, Z.; Liu, Z. Small molecule targeting FOXM1 DNA binding domain exhibits anti-tumor activity in ovarian cancer. Cell Death Discov. 2022, 8, 280. [Google Scholar] [CrossRef] [PubMed]
  108. Shiroma, Y.; Takahashi, R.U.; Yamamoto, Y.; Tahara, H. Targeting DNA binding proteins for cancer therapy. Cancer Sci. 2020, 111, 1058–1064. [Google Scholar] [CrossRef]
  109. Yingling, J.M.; Datto, M.B.; Wong, C.; Frederick, J.P.; Liberati, N.T.; Wang, X.F. Tumor Suppressor Smad4 Is a Transforming Growth Factor β-Inducible DNA Binding Protein. Mol. Cell. Biol. 1997, 17, 7019–7028. [Google Scholar] [CrossRef]
  110. Stefanoudakis, D.; Kathuria-Prakash, N.; Sun, A.W.; Abel, M.; Drolen, C.E.; Ashbaugh, C.; Zhang, S.; Hui, G.; Tabatabaei, Y.A.; Zektser, Y.; et al. The Potential Revolution of Cancer Treatment with CRISPR Technology. Cancers 2023, 15, 1813. [Google Scholar] [CrossRef] [PubMed]
  111. Ketron, A.C.; Denny, W.A.; Graves, D.E.; Osheroff, N. Amsacrine as a Topoisomerase II Poison: Importance of Drug–DNA Interactions. Biochemistry 2012, 51, 1730–1739. [Google Scholar] [CrossRef] [PubMed]
  112. Finlay, G.J.; Riou, J.F.; Baguley, B.C. From amsacrine to DACA (N-[2-(dimethylamino)ethyl]acridine-4-carboxamide): Selectivity for topoisomerases I and II among acridine derivatives. Eur. J. Cancer 1996, 32, 708–714. [Google Scholar] [CrossRef] [PubMed]
  113. Baguley, B.C.; Drummond, C.J.; Chen, Y.Y.; Finlay, G.J. DNA-Binding Anticancer Drugs: One Target, Two Actions. Molecules 2021, 26, 552. [Google Scholar] [CrossRef] [PubMed]
  114. Weber, G.F. DNA Damaging Drugs. In Molecular Therapies of Cancer; Springer International Publishing: Cham, Switzerland, 2015; pp. 9–112. [Google Scholar] [CrossRef]
  115. Ivanov, A.A. Explore Protein–Protein Interactions for Cancer Target Discovery Using the OncoPPi Portal. In Protein-Protein Interaction Networks: Methods and Protocols; Canzar, S., Ringeling, F.R., Eds.; Springer US: New York, NY, USA, 2020; pp. 145–164. [Google Scholar] [CrossRef]
  116. Hugo, W.; Sung, W.K.; Ng, S.K. Discovering Interacting Domains and Motifs in Protein–Protein Interactions. In Data Mining for Systems Biology: Methods and Protocols; Mamitsuka, H., DeLisi, C., Kanehisa, M., Eds.; Humana Press: Totowa, NJ, USA, 2013; pp. 9–20. [Google Scholar] [CrossRef]
  117. Bardwell, V.J.; Treisman, R. The POZ domain: A conserved protein-protein interaction motif. Genes Dev. 1994, 8, 1664–1677. [Google Scholar] [CrossRef] [PubMed]
  118. Zollman, S.; Godt, D.; Privé, G.G.; Couderc, J.L.; Laski, F.A. The BTB domain, found primarily in zinc finger proteins, defines an evolutionarily conserved family that includes several developmentally regulated genes in Drosophila. Proc. Natl. Acad. Sci. USA 1994, 91, 10717–10721. [Google Scholar] [CrossRef] [PubMed]
  119. Brayer, K.J.; Kulshreshtha, S.; Segal, D.J. The Protein-Binding Potential of C2H2 Zinc Finger Domains. Cell Biochem. Biophys. 2008, 51, 9–19. [Google Scholar] [CrossRef]
  120. Das, S.; Chakrabarti, S. Classification and prediction of protein–protein interaction interface using machine learning algorithm. Sci. Rep. 2021, 11, 1761. [Google Scholar] [CrossRef] [PubMed]
  121. Park, S.H.; Reyes, J.A.; Gilbert, D.R.; Kim, J.W.; Kim, S. Prediction of protein-protein interaction types using association rule based classification. BMC Bioinform. 2009, 10, 36. [Google Scholar] [CrossRef]
  122. Urquiza, J.M.; Rojas, I.; Pomares, H.; Herrera, J.; Florido, J.P.; Valenzuela, O.; Cepero, M. Using machine learning techniques and genomic/proteomic information from known databases for defining relevant features for PPI classification. Comput. Biol. Med. 2012, 42, 639–650. [Google Scholar] [CrossRef]
  123. Dunne, M.; Hupfeld, M.; Klumpp, J.; Loessner, M.J. Molecular Basis of Bacterial Host Interactions by Gram-Positive Targeting Bacteriophages. Viruses 2018, 10, 397. [Google Scholar] [CrossRef] [PubMed]
  124. Dunne, M.; Prokhorov, N.S.; Loessner, M.J.; Leiman, P.G. Reprogramming bacteriophage host range: Design principles and strategies for engineering receptor binding proteins. Curr. Opin. Biotechnol. 2021, 68, 272–281. [Google Scholar] [CrossRef] [PubMed]
  125. Taslem Mourosi, J.; Awe, A.; Guo, W.; Batra, H.; Ganesh, H.; Wu, X.; Zhu, J. Understanding Bacteriophage Tail Fiber Interaction with Host Surface Receptor: The Key “Blueprint” for Reprogramming Phage Host Range. Int. J. Mol. Sci. 2022, 23, 12146. [Google Scholar] [CrossRef] [PubMed]
  126. Bertozzi Silva, J.; Storms, Z.; Sauvageau, D. Host receptors for bacteriophage adsorption. FEMS Microbiol. Lett. 2016, 363. [Google Scholar] [CrossRef] [PubMed]
  127. Dunstan, R.A.; Pickard, D.; Dougan, S.; Goulding, D.; Cormie, C.; Hardy, J.; Li, F.; Grinter, R.; Harcourt, K.; Yu, L.; et al. The flagellotropic bacteriophage YSD1 targets Salmonella Typhi with a Chi-like protein tail fibre. Mol. Microbiol. 2019, 112, 1831–1846. [Google Scholar] [CrossRef] [PubMed]
  128. Berg, H.C.; Purcell, E.M. Physics of chemoreception. Biophys. J. 1977, 20, 193–219. [Google Scholar] [CrossRef]
  129. Axelrod, D.; Wang, M.D. Reduction-of-dimensionality kinetics at reaction-limited cell surface receptors. Biophys. J. 1994, 66, 588–600. [Google Scholar] [CrossRef] [PubMed]
  130. Langer, M.; Malykhin, A.; Maeda, K.; Chakrabarty, K.; Williamson, K.S.; Feasley, C.L.; West, C.M.; Metcalf, J.P.; Coggeshall, K.M. Bacillus anthracis peptidoglycan stimulates an inflammatory response in monocytes through the p38 mitogen-activated protein kinase pathway. PLoS ONE 2008, 3, e3706. [Google Scholar] [CrossRef]
  131. Aucher, W.; Davison, S.; Fouet, A. Characterization of the Sortase Repertoire in Bacillus anthracis. PLoS ONE 2011, 6, e27411. [Google Scholar] [CrossRef]
  132. Davison, S.; Couture-Tosi, E.; Candela, T.; Mock, M.; Fouet, A. Identification of the Bacillus anthracis lambda Phage Receptor. J. Bacteriol. 2005, 187, 6742–6749. [Google Scholar] [CrossRef]
  133. Wang, F.; Yang, W.; Hu, X. Discovery of High Affinity Receptors for Dityrosine through Inverse Virtual Screening and Docking and Molecular Dynamics. Int. J. Mol. Sci. 2019, 20, 115. [Google Scholar] [CrossRef] [PubMed]
  134. Smathers, R.L.; Petersen, D.R. The human fatty acid-binding protein family: Evolutionary divergences and functions. Human Genom. 2011, 5, 170. [Google Scholar] [CrossRef] [PubMed]
  135. Furuhashi, M.; Hotamisligil, G.S. Fatty acid-binding proteins: Role in metabolic diseases and potential as drug targets. Nat. Rev. Drug Discov. 2008, 7, 489–503. [Google Scholar] [CrossRef] [PubMed]
  136. Richieri, G.V.; Ogata, R.T.; Zimmerman, A.W.; Veerkamp, J.H.; Kleinfeld, A.M. Fatty Acid Binding Proteins from Different Tissues Show Distinct Patterns of Fatty Acid Interactions. Biochemistry 2000, 39, 7197–7204. [Google Scholar] [CrossRef] [PubMed]
  137. Toelzer, C.; Gupta, K.; Yadav, S.K.N.; Borucu, U.; Davidson, A.D.; Kavanagh Williamson, M.; Shoemark, D.K.; Garzoni, F.; Staufer, O.; Milligan, R.; et al. Free fatty acid binding pocket in the locked structure of SARS-CoV-2 spike protein. Science 2020, 370, 725–730. [Google Scholar] [CrossRef] [PubMed]
  138. Curry, S.; Mandelkow, H.; Brick, P.; Franks, N. Crystal structure of human serum albumin complexed with fatty acid reveals an asymmetric distribution of binding sites. Nat. Struct. Biol. 1998, 5, 827–835. [Google Scholar] [CrossRef] [PubMed]
  139. Lu, Y.; Yang, G.Z.; Yang, D. Effects of ligand binding on dynamics of fatty acid binding protein and interactions with membranes. Biophys. J. 2022, 121, 4024–4032. [Google Scholar] [CrossRef]
  140. Ghosh, S.; Dey, J. Binding of Fatty Acid Amide Amphiphiles to Bovine Serum Albumin: Role of Amide Hydrogen Bonding. J. Phys. Chem. B 2015, 119, 7804–7815. [Google Scholar] [CrossRef] [PubMed]
  141. Armstrong, E.H.; Goswami, D.; Griffin, P.R.; Noy, N.; Ortlund, E.A. Structural Basis for Ligand Regulation of the Fatty Acid-binding Protein 5, Peroxisome Proliferator-activated Receptor β/δ (FABP5-PPAR β/δ) Signaling Pathway. J. Biol. Chem. 2014, 289, 14941–14954. [Google Scholar] [CrossRef]
  142. Guillén, D.; Sánchez, S.; Rodríguez-Sanoja, R. Carbohydrate-binding domains: Multiplicity of biological roles. Appl. Microbiol. Biotechnol. 2010, 85, 1241–1249. [Google Scholar] [CrossRef]
  143. Henrissat, B.; Terrapon, N.; Coutinho, P.M.; Lombard, V.; Drula, E.; Garron, M.L.; Boulinguiez, M. Carbohydrate-Binding-Modules; Université d’Aix-Marseille: Marseille, France, 1998; Available online: http://www.cazy.org/Carbohydrate-Binding-Modules.html (accessed on 5 February 2024).
  144. Abbott, D.W.; Boraston, A.B. Chapter eleven—Quantitative Approaches to The Analysis of Carbohydrate-Binding Module Function. In Methods in Enzymology; Gilbert, H.J., Ed.; Academic Press: Cambridge, MA, USA, 2012; Volume 510, pp. 211–231. [Google Scholar] [CrossRef]
  145. Das, S.N.; Wagenknecht, M.; Nareddy, P.K.; Bhuvanachandra, B.; Niddana, R.; Balamurugan, R.; Swamy, M.J.; Moerschbacher, B.M.; Podile, A.R. Amino Groups of Chitosan Are Crucial for Binding to a Family 32 Carbohydrate Binding Module of a Chitosanase from Paenibacillus elgii. J. Biol. Chem. 2016, 291, 18977–18990. [Google Scholar] [CrossRef]
  146. Mathieu, S.V.; Aragão, K.S.; Imberty, A.; Varrot, A. Discoidin I from Dictyostelium discoideum and Interactions with Oligosaccharides: Specificity, Affinity, Crystal Structures, and Comparison with Discoidin II. J. Mol. Biol. 2010, 400, 540–554. [Google Scholar] [CrossRef]
  147. Kimoto, H.; Kusaoke, H.; Yamamoto, I.; Fujii, Y.; Onodera, T.; Taketo, A. Biochemical and Genetic Properties of Paenibacillus Glycosyl Hydrolase Having Chitosanase Activity and Discoidin Domain. J. Biol. Chem. 2002, 277, 14695–14702. [Google Scholar] [CrossRef]
  148. Desai, N.; Rana, D.; Salave, S.; Gupta, R.; Patel, P.; Karunakaran, B.; Sharma, A.; Giri, J.; Benival, D.; Kommineni, N. Chitosan: A Potential Biopolymer in Drug Delivery and Biomedical Applications. Pharmaceutics 2023, 15, 1313. [Google Scholar] [CrossRef]
  149. Banki, M.R.; Wood, D.W. Inteins and affinity resin substitutes for protein purification and scale up. Microb. Cell Fact. 2005, 4, 32. [Google Scholar] [CrossRef] [PubMed]
  150. Noren, C.J.; Wang, J.; Perler, F.B. Dissecting the Chemistry of Protein Splicing and Its Applications. Angew. Chem. Int. Ed. 2000, 39, 450–466. [Google Scholar] [CrossRef]
  151. Kikkawa, Y.; Tokuhisa, H.; Shingai, H.; Hiraishi, T.; Houjou, H.; Kanesato, M.; Imanaka, T.; Tanaka, T. Interaction Force of Chitin-Binding Domains onto Chitin Surface. Biomacromolecules 2008, 9, 2126–2131. [Google Scholar] [CrossRef] [PubMed]
  152. Eigenfeld, M.; Kerpes, R.; Whitehead, I.; Becker, T. Autofluorescence prediction model for fluorescence unmixing and age determination. Biotechnol. J. 2022, 17, 2200091. [Google Scholar] [CrossRef] [PubMed]
  153. Eigenfeld, M.; Wittmann, L.; Kerpes, R.; Schwaminger, S.P.; Becker, T. Studying the impact of cell age on the yeast growth behaviour of Saccharomyces pastorianus var. carlsbergensis by magnetic separation. Biotechnol. J. 2023, 18, 2200610. [Google Scholar] [CrossRef]
  154. Manjeet, K.; Purushotham, P.; Neeraja, C.; Podile, A.R. Bacterial chitin binding proteins show differential substrate binding and synergy with chitinases. Microbiol. Res. 2013, 168, 461–468. [Google Scholar] [CrossRef]
  155. Vaaje-Kolstad, G.; Horn, S.J.; van Aalten, D.M.F.; Synstad, B.; Eijsink, V.G.H. The Non-catalytic Chitin-binding Protein CBP21 from Serratia marcescens Is Essential for Chitin Degradation. J. Biol. Chem. 2005, 280, 28492–28497. [Google Scholar] [CrossRef] [PubMed]
  156. Nimlos, M.R.; Beckham, G.T.; Matthews, J.F.; Bu, L.; Himmel, M.E.; Crowley, M.F. Binding preferences, surface attachment, diffusivity, and orientation of a family 1 carbohydrate-binding module on cellulose. J. Biol. Chem. 2012, 287, 20603–20612. [Google Scholar] [CrossRef] [PubMed]
  157. Pinto, R.; Carvalho, J.; Mota, M.; Gama, M. Large-scale production of cellulose-binding domains. Adsorption studies using CBD-FITC conjugates. Cellulose 2006, 13, 557–569. [Google Scholar] [CrossRef]
  158. Linder, M.; Winiecka-Krusnell, J.; Linder, E. Use of Recombinant Cellulose-Binding Domains of Trichoderma reesei Cellulase as a Selective Immunocytochemical Marker for Cellulose in Protozoa. Appl. Environ. Microbiol. 2002, 68, 2503–2508. [Google Scholar] [CrossRef] [PubMed]
  159. Yang, J.K.; Xiong, W.; Chen, F.Y.; Xu, L.; Han, Z.G. Aromatic amino acids in the cellulose binding domain of Penicillium crustosum endoglucanase EGL1 differentially contribute to the cellulose affinity of the enzyme. PLoS ONE 2017, 12, e0176444. [Google Scholar] [CrossRef] [PubMed]
  160. Carrard, G.; Koivula, A.; Söderlund, H.; Béguin, P. Cellulose-binding domains promote hydrolysis of different sites on crystalline cellulose. Proc. Natl. Acad. Sci. USA 2000, 97, 10342–10347. [Google Scholar] [CrossRef] [PubMed]
  161. Griffo, A.; Rooijakkers, B.J.M.; Hähl, H.; Jacobs, K.; Linder, M.B.; Laaksonen, P. Binding Forces of Cellulose Binding Modules on Cellulosic Nanomaterials. Biomacromolecules 2019, 20, 769–777. [Google Scholar] [CrossRef] [PubMed]
  162. Richins, R.D.; Mulchandani, A.; Chen, W. Expression, immobilization, and enzymatic characterization of cellulose-binding domain-organophosphorus hydrolase fusion enzymes. Biotechnol. Bioeng. 2000, 69, 591–596. [Google Scholar] [CrossRef] [PubMed]
  163. Tomme, P.; Boraston, A.; McLean, B.; Kormos, J.; Creagh, A.L.; Sturch, K.; Gilkes, N.R.; Haynes, C.A.; Warren, R.A.J.; Kilburn, D.G. Characterization and affinity applications of cellulose-binding domains. Presented at the 2nd Conference on Affinity Technology, Arlington, VA, USA, 29–30 September 1997. J. Chromatogr. B Biomed. Sci. Appl. 1998, 715, 283–296. [Google Scholar] [CrossRef]
  164. Terpe, K. Overview of tag protein fusions: From molecular and biochemical fundamentals to commercial systems. Appl. Microbiol. Biotechnol. 2003, 60, 523–533. [Google Scholar] [CrossRef]
  165. Gardner, K.H.; Blackwell, J. The structure of native cellulose. Biopolymers 1974, 13, 1975–2001. [Google Scholar] [CrossRef]
  166. Hong, S. RNA Binding Protein as an Emerging Therapeutic Target for Cancer Prevention and Treatment. J. Cancer Prev. 2017, 22, 203–210. [Google Scholar] [CrossRef] [PubMed]
  167. Corley, M.; Burns, M.C.; Yeo, G.W. How RNA-Binding Proteins Interact with RNA: Molecules and Mechanisms. Mol. Cell 2020, 78, 9–29. [Google Scholar] [CrossRef] [PubMed]
  168. Lunde, B.M.; Moore, C.; Varani, G. RNA-binding proteins: Modular design for efficient function. Nat. Rev. Mol. Cell Biol. 2007, 8, 479–490. [Google Scholar] [CrossRef] [PubMed]
  169. Jahandideh, S.; Srinivasasainagendra, V.; Zhi, D. Comprehensive comparative analysis and identification of RNA-binding protein domains: Multi-class classification and feature selection. J. Theor. Biol. 2012, 312, 65–75. [Google Scholar] [CrossRef] [PubMed]
  170. Treger, M.; Westhof, E. Statistical analysis of atomic contacts at RNA–protein interfaces. J. Mol. Recognit. 2001, 14, 199–214. [Google Scholar] [CrossRef] [PubMed]
  171. Yu, Q.; Ye, W.; Jiang, C.; Luo, R.; Chen, H.F. Specific Recognition Mechanism between RNA and the KH3 Domain of Nova-2 Protein. J. Phys. Chem. B 2014, 118, 12426–12434. [Google Scholar] [CrossRef] [PubMed]
  172. Park, S.; Phukan, P.D.; Zeeb, M.; Martinez-Yamout, M.A.; Dyson, H.J.; Wright, P.E. Structural Basis for Interaction of the Tandem Zinc Finger Domains of Human Muscleblind with Cognate RNA from Human Cardiac Troponin T. Biochemistry 2017, 56, 4154–4168. [Google Scholar] [CrossRef] [PubMed]
  173. Liu, G.; Zhang, Q.; Xia, L.; Shi, M.; Cai, J.; Zhang, H.; Li, J.; Lin, G.; Xie, W.; Zhang, Y.; et al. RNA-binding protein CELF6 is cell cycle regulated and controls cancer cell proliferation by stabilizing p21. Cell Death Dis. 2019, 10, 688. [Google Scholar] [CrossRef]
  174. Nasiri-Aghdam, M.; Garcia-Garduño, T.C.; Jave-Suárez, L.F. CELF Family Proteins in Cancer: Highlights on the RNA-Binding Protein/Noncoding RNA Regulatory Axis. Int. J. Mol. Sci. 2021, 22, 11056. [Google Scholar] [CrossRef]
  175. Lin, G.; Li, J.; Cai, J.; Zhang, H.; Xin, Q.; Wang, N.; Xie, W.; Zhang, Y.; Xu, N. RNA-binding Protein MBNL2 regulates Cancer Cell Metastasis through MiR-182-MBNL2-AKT Pathway. J. Cancer 2021, 12, 6715–6726. [Google Scholar] [CrossRef] [PubMed]
  176. Vaishali; Dimitrova-Paternoga, L.; Haubrich, K.; Sun, M.; Ephrussi, A.; Hennig, J. Validation and classification of RNA binding proteins identified by mRNA interactome capture. RNA 2021, 27, 1173–1185. [Google Scholar] [CrossRef]
  177. Dallastella, M.; Oliveira, W.K.d.; Rodrigues, M.L.; Goldenberg, S.; Alves, L.R. The characterization of RNA-binding proteins and RNA metabolism-related proteins in fungal extracellular vesicles. Front. Cell. Infect. Microbiol. 2023, 13, 1247329. [Google Scholar] [CrossRef]
  178. Sidali, A.; Teotia, V.; Solaiman, N.S.; Bashir, N.; Kanagaraj, R.; Murphy, J.J.; Surendranath, K. AU-Rich Element RNA Binding Proteins: At the Crossroads of Post-Transcriptional Regulation and Genome Integrity. Int. J. Mol. Sci. 2022, 23, 96. [Google Scholar] [CrossRef] [PubMed]
  179. Löblein, M.T.; Falke, I.; Eich, H.T.; Greve, B.; Götte, M.; Troschel, F.M. Dual Knockdown of Musashi RNA-Binding Proteins MSI-1 and MSI-2 Attenuates Putative Cancer Stem Cell Characteristics and Therapy Resistance in Ovarian Cancer Cells. Int. J. Mol. Sci. 2021, 22, 11502. [Google Scholar] [CrossRef]
  180. Dolicka, D.; Foti, M.; Sobolewski, C. The Emerging Role of Stress Granules in Hepatocellular Carcinoma. Int. J. Mol. Sci. 2021, 22, 9428. [Google Scholar] [CrossRef]
  181. Kang, D.; Lee, Y.; Lee, J.S. RNA-Binding Proteins in Cancer: Functional and Therapeutic Perspectives. Cancers 2020, 12, 2699. [Google Scholar] [CrossRef]
  182. Niehrs, C.; Luke, B. Regulatory R-loops as facilitators of gene expression and genome stability. Nat. Rev. Mol. Cell Biol. 2020, 21, 167–178. [Google Scholar] [CrossRef] [PubMed]
  183. Ouyang, J.; Yadav, T.; Zhang, J.M.; Yang, H.; Rheinbay, E.; Guo, H.; Haber, D.A.; Lan, L.; Zou, L. RNA transcripts stimulate homologous recombination by forming DR-loops. Nature 2021, 594, 283–288. [Google Scholar] [CrossRef]
  184. Murphy, J.J.; Surendranath, K.; Kanagaraj, R. RNA-Binding Proteins and Their Emerging Roles in Cancer: Beyond the Tip of the Iceberg. Int. J. Mol. Sci. 2023, 24, 9612. [Google Scholar] [CrossRef]
  185. Adachi, T.; Nakamura, Y. Aptamers: A Review of Their Chemical Properties and Modifications for Therapeutic Application. Molecules 2019, 24, 4229. [Google Scholar] [CrossRef] [PubMed]
  186. Limsirichai, P.; Gaj, T.; Schaffer, D.V. CRISPR-mediated Activation of Latent HIV-1 Expression. Mol. Ther. 2016, 24, 499–507. [Google Scholar] [CrossRef] [PubMed]
  187. Nimjee, S.M.; White, R.R.; Becker, R.C.; Sullenger, B.A. Aptamers as Therapeutics. Annu. Rev. Pharmacol. Toxicol. 2017, 57, 61–79. [Google Scholar] [CrossRef] [PubMed]
  188. Gao, F.; Yin, J.; Chen, Y.; Guo, C.; Hu, H.; Su, J. Recent advances in aptamer-based targeted drug delivery systems for cancer therapy. Front. Bioeng. Biotechnol. 2022, 10, 972933. [Google Scholar] [CrossRef] [PubMed]
  189. Friedman, A.D.; Kim, D.; Liu, R. Highly stable aptamers selected from a 2’-fully modified fGmH RNA library for targeting biomaterials. Biomaterials 2015, 36, 110–123. [Google Scholar] [CrossRef] [PubMed]
  190. Song, K.M.; Lee, S.; Ban, C. Aptamers and their biological applications. Sensors 2012, 12, 612–631. [Google Scholar] [CrossRef] [PubMed]
  191. Mascini, M. Aptamers and their applications. Anal. Bioanal. Chem. 2008, 390, 987–988. [Google Scholar] [CrossRef] [PubMed]
  192. Ma, H.; Ó’Fágáin, C.; O’Kennedy, R. Antibody stability: A key to performance—Analysis, influences and improvement. Biochimie 2020, 177, 213–225. [Google Scholar] [CrossRef]
  193. Hayashi, T.; Oshima, H.; Mashima, T.; Nagata, T.; Katahira, M.; Kinoshita, M. Binding of an RNA aptamer and a partial peptide of a prion protein: Crucial importance of water entropy in molecular recognition. Nucleic Acids Res. 2014, 42, 6861–6875. [Google Scholar] [CrossRef]
  194. Cai, S.; Yan, J.; Xiong, H.; Liu, Y.; Peng, D.; Liu, Z. Investigations on the interface of nucleic acid aptamers and binding targets. Analyst 2018, 143, 5317–5338. [Google Scholar] [CrossRef]
  195. Johansson, H.E.; Liljas, L.; Uhlenbeck, O.C. RNA Recognition by the MS2 Phage Coat Protein. Semin. Virol. 1997, 8, 176–185. [Google Scholar] [CrossRef]
  196. Miyakawa, S.; Nomura, Y.; Sakamoto, T.; Yamaguchi, Y.; Kato, K.; Yamazaki, S.; Nakamura, Y. Structural and molecular basis for hyperspecificity of RNA aptamer to human immunoglobulin G. RNA 2008, 14, 1154–1163. [Google Scholar] [CrossRef] [PubMed]
  197. Murakami, K.; Izuo, N.; Bitan, G. Aptamers targeting amyloidogenic proteins and their emerging role in neurodegenerative diseases. J. Biol. Chem. 2022, 298, 101478. [Google Scholar] [CrossRef] [PubMed]
  198. Kim, S.T.; Kim, H.G.; Kim, Y.M.; Han, H.S.; Cho, J.H.; Lim, S.C.; Lee, T.; Jahng, G.H. An aptamer-based magnetic resonance imaging contrast agent for detecting oligomeric amyloid-β in the brain of an Alzheimer’s disease mouse model. NMR Biomed. 2023, 36, e4862. [Google Scholar] [CrossRef] [PubMed]
  199. Kohlberger, M.; Gadermaier, G. SELEX: Critical factors and optimization strategies for successful aptamer selection. Biotechnol. Appl. Biochem. 2022, 69, 1771–1792. [Google Scholar] [CrossRef] [PubMed]
  200. Peinetti, A.S.; Lake, R.J.; Cong, W.; Cooper, L.; Wu, Y.; Ma, Y.; Pawel, G.T.; Toimil-Molares, M.E.; Trautmann, C.; Rong, L.; et al. Direct detection of human adenovirus or SARS-CoV-2 with ability to inform infectivity using DNA aptamer-nanopore sensors. Sci. Adv. 2021, 7, eabh2848. [Google Scholar] [CrossRef]
  201. Belikov, S.; Berg, O.G.; Wrange, Ö. Quantification of transcription factor-DNA binding affinity in a living cell. Nucleic Acids Res. 2015, 44, 3045–3058. [Google Scholar] [CrossRef]
  202. Hanaoka, S.; Nagadoi, A.; Nishimura, Y. Comparison between TRF2 and TRF1 of their telomeric DNA-bound structures and DNA-binding activities. Protein Sci. 2005, 14, 119–130. [Google Scholar] [CrossRef] [PubMed]
  203. Pääkkönen, J.; Jänis, J.; Rouvinen, J. Calculation and Visualization of Binding Equilibria in Protein Studies. ACS Omega 2022, 7, 10789–10795. [Google Scholar] [CrossRef]
  204. Zhang, M.; Wang, H.; Foster, E.R.; Nikolov, Z.L.; Fernando, S.D.; King, M.D. Binding behavior of spike protein and receptor binding domain of the SARS-CoV-2 virus at different environmental conditions. Sci. Rep. 2022, 12, 789. [Google Scholar] [CrossRef]
  205. Ikegami, T.; Okada, T.; Hashimoto, M.; Seino, S.; Watanabe, T.; Shirakawa, M. Solution Structure of the Chitin-binding Domain of Bacillus circulans WL-12 Chitinase A1. J. Biol. Chem. 2000, 275, 13654–13661. [Google Scholar] [CrossRef] [PubMed]
  206. Madland, E.; Forsberg, Z.; Wang, Y.; Lindorff-Larsen, K.; Niebisch, A.; Modregger, J.; Eijsink, V.G.H.; Aachmann, F.L.; Courtade, G. Structural and functional variation of chitin-binding domains of a lytic polysaccharide monooxygenase from Cellvibrio japonicus. J. Biol. Chem. 2021, 297, 101084. [Google Scholar] [CrossRef] [PubMed]
  207. Zeltins, A.; Schrempp, H. Specific Interaction of the Streptomyces Chitin-Binding Protein Chb1 with α-Chitin. Eur. J. Biochem. 1997, 246, 557–564. [Google Scholar] [CrossRef] [PubMed]
  208. Schnellmann, J.; Zeltins, A.; Blaak, H.; Schrempf, H. The novel lectin-like protein CHB1 is encoded by a chitin-inducible Streptomyces olivaceoviridis gene and binds specifically to crystalline α-chitin of fungi and other organisms. Mol. Microbiol. 1994, 13, 807–819. [Google Scholar] [CrossRef] [PubMed]
  209. Shinya, S.; Fukamizo, T. Interaction between chitosan and its related enzymes: A review. Int. J. Biol. Macromol. 2017, 104, 1422–1435. [Google Scholar] [CrossRef] [PubMed]
  210. Goldstein, M.A.; Takagi, M.; Hashida, S.; Shoseyov, O.; Doi, R.H.; Segel, I.H. Characterization of the cellulose-binding domain of the Clostridium cellulovorans cellulose-binding protein A. J. Bacteriol. 1993, 175, 5762–5768. [Google Scholar] [CrossRef] [PubMed]
  211. Consortium, T.C. Ten years of CAZypedia: A living encyclopedia of carbohydrate-active enzymes. Glycobiology 2017, 28, 3–8. [Google Scholar] [CrossRef] [PubMed]
  212. Boraston, A.B.; Warren, R.A.J.; Kilburn, D.G. β-1,3-Glucan Binding by a Thermostable Carbohydrate-Binding Module from Thermotoga maritima. Biochemistry 2001, 40, 14679–14685. [Google Scholar] [CrossRef] [PubMed]
  213. Hurlburt, N.K.; Chen, L.H.; Stergiopoulos, I.; Fisher, A.J. Structure of the Cladosporium fulvum Avr4 effector in complex with (GlcNAc)6 reveals the ligand-binding mechanism and uncouples its intrinsic function from recognition by the Cf-4 resistance protein. PLoS Pathog. 2018, 14, e1007263. [Google Scholar] [CrossRef]
  214. Georgelis, N.; Tabuchi, A.; Nikolaidis, N.; Cosgrove, D.J. Structure-Function Analysis of the Bacterial Expansin EXLX1. J. Biol. Chem. 2011, 286, 16814–16823. [Google Scholar] [CrossRef]
  215. Forsberg, Z.; Nelson, C.E.; Dalhus, B.; Mekasha, S.; Loose, J.S.M.; Crouch, L.I.; Røhr, Å.K.; Gardner, J.G.; Eijsink, V.G.H.; Vaaje-Kolstad, G. Structural and Functional Analysis of a Lytic Polysaccharide Monooxygenase Important for Efficient Utilization of Chitin in Cellvibrio japonicus. J. Biol. Chem. 2016, 291, 7300–7312. [Google Scholar] [CrossRef] [PubMed]
  216. Leth, M.L.; Ejby, M.; Workman, C.; Ewald, D.A.; Pedersen, S.S.; Sternberg, C.; Bahl, M.I.; Licht, T.R.; Aachmann, F.L.; Westereng, B.; et al. Differential bacterial capture and transport preferences facilitate co-growth on dietary xylan in the human gut. Nat. Microbiol. 2018, 3, 570–580. [Google Scholar] [CrossRef] [PubMed]
  217. Jalak, J.; Väljamäe, P. Multi-Mode Binding of Cellobiohydrolase Cel7A from Trichoderma reesei to Cellulose. PLoS ONE 2014, 9, e108181. [Google Scholar] [CrossRef] [PubMed]
  218. Baumann, M.J.; Borch, K.; Westh, P. Xylan oligosaccharides and cellobiohydrolase I (TrCel7A) interaction and effect on activity. Biotechnol. Biofuels 2011, 4, 45. [Google Scholar] [CrossRef] [PubMed]
  219. Mitsumori, M.; Xu, L.; Kajikawa, H.; Kurihara, M. Properties of cellulose-binding modules in endoglucanase F from Fibrobacter succinogenes S85 by means of surface plasmon resonance. FEMS Microbiol. Lett. 2002, 214, 277–281. [Google Scholar] [CrossRef]
  220. Mitsumori, M.; Minato, H. Identification of the cellulose-binding domain of Fibrobacter succinogenes endoglucanase F. FEMS Microbiol. Lett. 2000, 183, 99–103. [Google Scholar] [CrossRef]
  221. Jouravleva, K.; Vega-Badillo, J.; Zamore, P.D. Principles and pitfalls of high-throughput analysis of microRNA-binding thermodynamics and kinetics by RNA Bind-n-Seq. Cell Rep. Methods 2022, 2, 100185. [Google Scholar] [CrossRef]
  222. Cléry, A.; Blatter, M.; Allain, F.H.T. RNA recognition motifs: Boring? Not quite. Curr. Opin. Struct. Biol. 2008, 18, 290–298. [Google Scholar] [CrossRef] [PubMed]
  223. Lyu, Y.; Teng, I.T.; Zhang, L.; Guo, Y.; Cai, R.; Zhang, X.; Qiu, L.; Tan, W. Comprehensive Regression Model for Dissociation Equilibria of Cell-Specific Aptamers. Anal. Chem. 2018, 90, 10487–10493. [Google Scholar] [CrossRef]
  224. Chang, A.L.; McKeague, M.; Liang, J.C.; Smolke, C.D. Kinetic and Equilibrium Binding Characterization of Aptamers to Small Molecules using a Label-Free, Sensitive, and Scalable Platform. Anal. Chem. 2014, 86, 3273–3278. [Google Scholar] [CrossRef]
Figure 1. An illustrative comparison of particle modification techniques for selective orientation in catalytic processes on the left side; on the right, the indiscriminate attachment of cells due to interactions between cell wall proteins and surface charges.
Figure 1. An illustrative comparison of particle modification techniques for selective orientation in catalytic processes on the left side; on the right, the indiscriminate attachment of cells due to interactions between cell wall proteins and surface charges.
Life 14 00630 g001
Figure 2. Schematic diagram of an application scenario for the use of natural binding proteins in targeted drug delivery and MRI targeting. Created using biorender.com.
Figure 2. Schematic diagram of an application scenario for the use of natural binding proteins in targeted drug delivery and MRI targeting. Created using biorender.com.
Life 14 00630 g002
Figure 3. Diagram illustrating various binding proteins and motifs within the complete protein complex. Data of the protein structure were obtained from www.rcsb.org (datasets 2BSD, 1ED7, 3FYS, 3RHI, and 2W4S; accessed on 2 May 2024). Created using biorender.com.
Figure 3. Diagram illustrating various binding proteins and motifs within the complete protein complex. Data of the protein structure were obtained from www.rcsb.org (datasets 2BSD, 1ED7, 3FYS, 3RHI, and 2W4S; accessed on 2 May 2024). Created using biorender.com.
Life 14 00630 g003
Figure 4. Visual representation of carbohydrate-binding domain families (CBMs) and their ligands according to data of [144] and modified. Differentiation is performed between amorphous and crystalline cellulose.
Figure 4. Visual representation of carbohydrate-binding domain families (CBMs) and their ligands according to data of [144] and modified. Differentiation is performed between amorphous and crystalline cellulose.
Life 14 00630 g004
Table 1. Classification of DNA-binding domains; adopted from Gonzales et al. [86] and modified according to Wingender [93].
Table 1. Classification of DNA-binding domains; adopted from Gonzales et al. [86] and modified according to Wingender [93].
SuperclassClassFamily
Basic domain
Leucine zipper factors
AP-1(-like) components
CREB
C/EBP-like factors
bZIP/PAR
Plant-G-box binding factors
ZIP only
Other bZIP factors
Helix–loop–helix factors (bHLH)
Ubiquitous (class A) factors
Myogenic transcription factors
Achaete–scute
Tal/Twist/Atonal/Hen
Hairy
Factors with PAS domain
INO
HLH domain only
Other bHLH factors
Helix–loop–helix/leucine zipper factors (bHLH-ZIP)
Ubiquitous bHLH-ZIP factors
Cell-cycle controlling factors
NF-1
NF-1
RF-X
RF-X
bHSH
AP-2
Zinc-coordinating DNA-binding domains
Cys4 zinc finger of nuclear receptor type
Cys4 zinc finger of nuclear receptor type
Thyroid hormone receptor-like factors
Diverse Cys4 zinc fingers
GATA factors
Trithorax
Other factors
Cys2His2 zinc finger domain
Ubiquitous factors
Developmental/cell cycle regulators
Metabolic regulators in fungi
Large factors with NF-6B-like binding properties
Viral regulator
Cys6 cysteine–zinc cluster
Metabolic regulators in fungi
Zinc fingers of alternating composition
Cx7Hx8Cx4C zinc fingers
Cx2Hx4Hx4C zinc fingers
Helix–turn–helix
Homeodomain
Homeodomain only
POU domain factors
Homeodomain with LIM region
Homeodomain plus zinc finger motifs
Paired box
Paired plus homeodomain
Paired domain only
Fork head/winged helix
Developmental regulators
Tissue-specific regulators
Cell-cycle controlling factors
Other regulators
Heat shock factors
HSF
Tryptophan clusters
Myb
Ets-type
Interferon-regulating factors
TEA domain
TEA
Beta scaffold factors with minor groove contacts
Rel homology region (RHR)
Rel/ankyrin
Ankyrin only
NF-AT
STAT
STAT
P53
P53
MADS box
Regulators of differentiation
Responders to external signals
Metabolic regulators
β -Barrel α -helix transcription factors
E2
TATA-binding proteins
TBP
HMG
SOX
TCF-1
HMG2-related
UBF
MATA
Other HMG box factors
Heteromeric CCAAT factors
Heteromeric CCAAT factors
Grainyhead
Grainyhead
Cold-shock domain factors
csd
Runt
Runt
Other transcription factors
HMGI(Y)
HMGI(Y)
Pocket domain
Rb
CBP
E1 A-like factors
E1A
AP2/EREBP-related factors
AP2
EREBP
AP2/B3
Table 3. Classification of RNA-binding proteins based on their recognition surfaces, based on data of Lunde et al. [168] and modified.
Table 3. Classification of RNA-binding proteins based on their recognition surfaces, based on data of Lunde et al. [168] and modified.
DomainTopologyRNA Recognition Surface Notes
RRM α β Surface of β -sheet
KH type I α β Hydrophobic cleft formed by variable loop between β 2 , β 3 , and GXXG loop
KH type II α β Same as type I, except variable loop is between α 2 and β 2
dsRBD α β Helix α 1 , N-terminal of helix α 2 , and loop between β 1 and β 2
Znf-CCHH α β Primarily residues in α -helices
Znf-CCHHLittle regular secondary structureAromatic side chains form hydrophobic binding pockets for bases that make direct hydrogen bonds to protein backbone
S1 β Core formed by two β -strands with contributions from surrounding loops
PAZ α β Hydrophobic pocket formed by OB-like β -barrel and small α β motif
PIWI α β Highly conserved pocket, including a metal ion that is bound to the exposed C-terminal carboxylate
TRAP β Edges of β -sheets between each of the 11 subunits that form the entire protein structure
Pumilio α Two repeats combine to form binding pocket for individual bases, helix α 2 provides specificity-determining residues
SAM α Hydrophobic cavity between three helices surrounded by an electropositive region
Table 4. Comparison of different protein classes and their corresponding binding affinities.
Table 4. Comparison of different protein classes and their corresponding binding affinities.
Protein ClassSpecific ProteinBinding Affinity  K D [nM]Size [Amino Acids]Reference
DNA-binding domainDNA binding by glucocorticoid receptor1.000  [201]
DNA binding by androgen receptor130  [201]
DNA-binding proteins telomer repeat binding factor TRF1 and TRF2200 and 75063 [202]
Prokaryotic transcriptional regulators of multiple antibiotic resistance in E. coli 129 [85]
Protein-binding domainCompetitive binding of a ligand to two receptors100–80,000Simulation data [203]
Spike protein and receptor-binding domain314–3137  [204]
Fatty acid-binding proteinHuman FABP112717–23 [134]
Carbohydrate-binding domainChitin-binding domain of chitinase A1 from Bacillus circulans149–22845 [75,205]
Chitin-binding domain of a lytic polysaccharide monooxygenase from Cellvibrio japonicus2900–850058 [206]
Chitin-binding domain from Streptomyces110–2170100/200/201 [207,208]
Chitosan-binding module from Paenibacillus elgii 132 [209]
Chitosan-binding module from Paenibacillus sp. 1K-5 260 [209]
DD127,200–3,770,000  [146]
Clostridium cellulovorans cellulose-binding protein A500–1400161 [210]
Scaffoldin (CipA) containing a CBM3 family domain of Gram-positive bacterias such as Clostridium thermocellulum400150 [211,212]
CBM4 glycanases from thermophilic and mesophilic bacteria000– 0,000150 [211,212]
CBM10 families 4000 towards cellulose 45[211,212]
CBM14 from fungal tomato pathogen Cladosporium fulvum towards ( G l c N A c 6 )  6700 70[211,213]
CBM63 based on C-terminus of expansin BsEXLX1 from Bacillus subtilis2100 towards cellulose100[211,214]
CBM73 of trimodular LPMO4300 towards α -chitin60[211,215,216]
CBM86 of xylanase in Roseburia intestinalis480,000 towards xylohexaose, 490,000 towards xylopentaose, 998,000 towards xylotetraose, and 1,900,000 towards xylotriose138[211,216]
Cellobiohydrolase TrCel7A from Trichoderma reesei2.936 [217,218]
AD2 from Fibrobacter succinogenes S85397.95411 [219,220]
AD4 from Fibrobacter succinogenes S85838.51207 [219,220]
RNA-binding domainAGO2 let-7a0.004–0.8  [221]
90 [222]
AptamerJHIT-1–JHIT-7; LZH-1–LZH-17 against HepG2 target cells3.9-2516.3  [223]
Target: flavin mononucleotide1100 ± 400  [224]
Malachite green950 ± 340  [224]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Eigenfeld, M.; Lupp, K.F.M.; Schwaminger, S.P. Role of Natural Binding Proteins in Therapy and Diagnostics. Life 2024, 14, 630. https://doi.org/10.3390/life14050630

AMA Style

Eigenfeld M, Lupp KFM, Schwaminger SP. Role of Natural Binding Proteins in Therapy and Diagnostics. Life. 2024; 14(5):630. https://doi.org/10.3390/life14050630

Chicago/Turabian Style

Eigenfeld, Marco, Kilian F. M. Lupp, and Sebastian P. Schwaminger. 2024. "Role of Natural Binding Proteins in Therapy and Diagnostics" Life 14, no. 5: 630. https://doi.org/10.3390/life14050630

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop