Next Article in Journal
Oxidative Stress in NAFLD: Role of Nutrients and Food Contaminants
Next Article in Special Issue
Therapeutic Nanobodies Targeting Cell Plasma Membrane Transport Proteins: A High-Risk/High-Gain Endeavor
Previous Article in Journal
The Regulation of Fat Metabolism during Aerobic Exercise
Previous Article in Special Issue
Nanobodies as Versatile Tool for Multiscale Imaging Modalities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanobodies Right in the Middle: Intrabodies as Toolbox to Visualize and Modulate Antigens in the Living Cell

by
Teresa R. Wagner
1,2 and
Ulrich Rothbauer
1,2,*
1
Pharmaceutical Biotechnology, Eberhard Karls University Tuebingen, 72076 Tuebingen, Germany
2
Natural and Medical Sciences Institute, University of Tuebingen, 72770 Reutlingen, Germany
*
Author to whom correspondence should be addressed.
Biomolecules 2020, 10(12), 1701; https://doi.org/10.3390/biom10121701
Submission received: 30 November 2020 / Revised: 15 December 2020 / Accepted: 18 December 2020 / Published: 21 December 2020
(This article belongs to the Special Issue The Therapeutic and Diagnostic Potential of Nanobodies)

Abstract

:
In biomedical research, there is an ongoing demand for new technologies to elucidate disease mechanisms and develop novel therapeutics. This requires comprehensive understanding of cellular processes and their pathophysiology based on reliable information on abundance, localization, post-translational modifications and dynamic interactions of cellular components. Traceable intracellular binding molecules provide new opportunities for real-time cellular diagnostics. Most prominently, intrabodies derived from antibody fragments of heavy-chain only antibodies of camelids (nanobodies) have emerged as highly versatile and attractive probes to study and manipulate antigens within the context of living cells. In this review, we provide an overview on the selection, delivery and usage of intrabodies to visualize and monitor cellular antigens in living cells and organisms. Additionally, we summarize recent advances in the development of intrabodies as cellular biosensors and their application to manipulate disease-related cellular processes. Finally, we highlight switchable intrabodies, which open entirely new possibilities for real-time cell-based diagnostics including live-cell imaging, target validation and generation of precisely controllable binding reagents for future therapeutic applications.

1. Selection of Intracellular Functional Nanobodies

More than thirty years after their discovery [1], antibody fragments derived from heavy-chain only antibodies of camelids, termed as variable heavy chain of heavy-chain only antibodies (VHH) or nanobodies (Nbs), have emerged as highly potent and versatile binding molecules for biomedical research, diagnostics and therapy [2]. With caplacizumab (ALX-0081), the first therapeutic Nb was approved in the U.S. and Europe for treatment of acquired thrombotic thrombocytopenic purpura in 2019 [3,4]. Compared to conventional antibodies (IgGs, ~150 kDa), Nbs are 10-times smaller (~15 kDa) and are characterized by stable folding, a compact structure, high solubility and chemical stability [2,5,6]. Due to their beneficial properties and accessibility to various cell compartments, Nbs applied as intrabodies have turned out to be highly suitable intracellular binding molecules (reviewed in [7,8,9]). Having the ability to specifically visualize and/or modulate endogenous targets within living cells, intrabodies provide distinct advantages compared to other technologies such as fluorescent fusion proteins, conditional gene expression, RNA interference and chemical or genetic knockouts. Nevertheless, when it comes to intracellular applications it has to be considered that Nbs, like any other antibody-derived binding molecule, comprise evolutionary conserved disulfide bonds. While the endoplasmic reticulum (ER) and mitochondria provide naturally suitable compartments for Nbs due to their (at least partially) oxidative environment, the reducing milieu of the cytosol strongly impairs formation of disulfide bonds. This can massively affect correct paratope formation and overall folding. Consequently, only Nbs whose functionalities do not rely on the formation of disulfide bonds are suitable candidates for the successful application as intrabodies.
Considering that most Nbs yielded from conventional screening approaches are non-functional within living cells [10], multiple approaches have been developed to improve selection of intracellularly functional binders from synthetic or immunized libraries. One obvious strategy to overcome such limitations is to remove conserved cysteine residues within antibody-derived fragments. Therefore, either complete libraries or selected high affinity binders initially identified from conventional screenings such as phage display were site directly mutated to cysteine-free derivates [11]. However, in most cases, depletion of conserved disulfide bonds causes a major loss in stability, therefore extensive functional downstream analysis is needed, limiting throughput and diversity of intracellularly functional antibody fragments [11,12]. Alternatively, Olichon and Surrey genetically combined Nbs with the disulfide bond isomerase (DsbC), which catalyzes formation of disulfide bonds in the cytoplasm. Following this strategy, they were able to produce functional tubulin-binding Nbs in the cytoplasm of E. coli [13]. However, as this approach is restricted to certain expression systems and requires the generation of complex fusion proteins, it has some severe constraints which limit its broad application. Further studies suggested that, especially, physiochemical parameters have an important influence on the development of cytoplasmic stable intrabodies [14,15]. Recently, single chain antibody fragments (scFvs) were fused to peptide-tags, conferring a strong negative charge and low isoelectric point in order to generate so-called ultra-stable cytoplasmic antibodies (STANDs) [16]. In their study, the authors demonstrated how fusion of scFvs to negatively charged s3Flag and HA peptide-tags induces a significantly increased stability of aggregation-prone cytoplasmic intrabodies [16]. Notably, many Nbs are already strongly charged, showing either a positive or negative isoelectric point (pI) and therefore are less prone to aggregate at the most neutral pH of the cytoplasm [17]. In contrast to modifying sequences of existing Nbs, there is a long history of screening strategies aiming at initial identification of intrabodies. The first described methodology for this purpose was an adopted yeast two-hybrid (Y2H) assay [18]. For selection of bait/prey interactors, antibody fragments and antigens are fused to either a DNA-binding domain (bait) or to a transcriptional activation domain (prey). In the case of successful antigen binding, a reporter gene is activated and induces an easily detectable phenotype e.g., growth on selection media [19,20]. The advantage of this method is the flexible expression of potential intrabodies in cytosol and nucleus. However, it suffers from selecting large numbers of false positives and restricted screenable library sizes (>106–107 individual clones) [21]. Consequently, the intracellular antigen capture technology (IACT) was developed, in which a pre-selection step via phage-based biopanning is performed to enrich potential candidates prior to the Y2H screen [22,23]. Since its first description, the IACT has been successfully applied to identify consensus sequences within different binding molecules optimized for intracellular expression [24]. Thereby, a new selection strategy exclusively focusing on the complementarity-determining regions (CDRs) was established (iDabs) [22,25]. Similarly, Moutel et al. exploited the idea of intracellular optimized scaffolds. Based on a humanized synthetic Nb (hs2dAb) scaffold optimized for intracellular stability, they generated a fully synthetic phage display library of humanized Nbs from which they successfully selected effectively working intrabodies against a variety of targets including fluorescent proteins and cytoskeletal structures [10].
Other intrabody screening approaches employ a bacterial-two-hybrid (B2H) system specifically tailored for the identification of intrabodies based on distinct Nb scaffolds [21]. Compared to yeast, this bacterial-based system has clear advantages coming from faster growth rates and higher transformation efficiencies, two important characteristics that are beneficial for high-throughput selection of intrabodies. The B2H encompasses fusion constructs of Nbs and antigens with the lambda repressor and the α-subunit of the RNA polymerase, respectively. In the case of antigen binding within the bacterial cytoplasm, expression of selective marker genes essential for the survival under selection pressure is initiated and intracellular functional Nbs can be selected [21,26]. Notably, as the screening is performed in bacteria, functionality and specificity of B2H-selected intrabodies has to be further validated in eukaryotic systems e.g., by employing the mammalian cell-based fluorescent-2-hybrid (F2H) approach [27].
This limitation was elegantly overcome by Schmidt et al. using a lentiviral screening technology to select Nbs that are stably producible in mammalian cells and simultaneously neutralize influenza virus or vesicular stomatitis virus. For this approach, the Nb library derived from an immunized camelid was inserted into a lentiviral vector followed by the transduction of A549 cells. Upon application of a lethal dose of the virus of interest only those cells survived, which successfully expressed an effective viral neutralizing intrabody. In a proof-of-concept study, they described successful Nb selection and outlined a potential therapeutic application of virus-specific intrabodies [28]. Notably, similar concepts were realized by screening for functional scFv-derived intrabodies in mammalian cells. For example, scFv-derived intrabodies were selected by FACS analysis of impaired degranulation as a selection marker in rat leukemic cells [29], or intracellular binding of scFvs fused to the cytoplasmic domain of a receptor tyrosine kinase to its homo-oligomeric antigen became detectable by growth signal induction in a B-cell line [30]. Indisputably, high-throughput selection of target specific Nbs is challenging and the generation of stable, intracellularly functional Nbs adds another level of complexity to the screening process. However, as shortly outlined in this chapter, different screening approaches became available which support accelerated identification of suitable intrabodies based on Nbs.

2. Delivery Systems of Intrabodies

After successful selection of intrabodies, the next challenge is to transport them into living cells or organisms. Here, one can chose between two options: transfer of the Nb-encoding cDNAs or introducing Nbs as purified proteins. While the first option is straightforward and employs broadly available and standardized expression systems in combination with effectively working transfection protocols e.g., using lipofection, electroporation or transfection of nanoparticles, these approaches suffer from substantial toxic side effects and hard-to-transfect cell types such as primary cells. Furthermore, expression levels do not directly correlate with the amount of transfected cDNA. Thus, resulting intrabody levels can be very heterogeneous and are hardly controllable. Therefore, while these approaches are easy applicable to studying intrabodies in single cells, they are less suitable for generation of intrabodies comprising cellular models which can be implemented e.g., in compound screening campaigns. Using a precise genomic insertion of intrabody-encoding sequences via CRISPR gene editing might be a more efficient approach to generate cell lines stably and homogenously expressing intrabodies [31]. Especially for in vivo analysis, gene delivery based on adeno-associated virus causing a long and stable expression, is preferred. For example, viral vector-based delivery of Nbs enabled live-cell STED microscopy of neuronal actin in mouse [32] or targeted degradation of α-synuclein in rats [33]. In consequence, combining cell-specific viral delivery with precise CRISPR gene editing could be an efficient way to introduce intrabodies into animal models for preclinical testing. Notably, gene delivery bears distinct advantages such as an easy guidance of intrabodies to defined cellular compartments by simply adding mitochondrial-, nuclear- or ER-localization and/or retention sequences [34]. However, current concerns regarding intrabodies transduced by gene delivery such as non-controllable expression yields, low cell or tissue specificity or unsolved safety issues still comprise major limitations for the development of intrabodies especially for therapeutic applications.
To avoid genetic cell manipulation one might try to deliver intrabodies directly as proteins. Multiple methods and technologies for protein transduction are now available (reviewed in [35,36,37] with focus on antibodies) which also can be applied for Nbs. Physical methods including electroporation, microinjection, sonoporation or transfer by cell squeezing [38,39,40] can be used to transfer Nbs into cells. However, with these approaches only a limited number of cells can be addressed which afterward often suffer from severe cellular damage. Alternatively, Nbs can be coupled to protein transduction domains (PTDs) or cell-penetrating peptides (CPPs) to facilitate their cellular uptake. As most prominent examples, positively charged peptides such as HIV-1 TAT [41,42] have to be mentioned, but also other CCPs from natural and synthetic sources [43,44,45,46] have been successfully employed for cellular delivery of intrabodies. More recently, further approaches such as virus-like particles [47], silica nanoparticles [48], nanocapsules [49], polycationic resurfacing [50], charge-conversional polyion complex micelles [51], liposomal carriers [52] or oligoaminoamide carriers [53], to name a few, have been described as delivery options for Nbs. Notably, cellular transduction of Nbs has the exclusive advantage that any cargo in tow e.g., synthetic dyes or small molecule inhibitors, can be delivered alongside them. However, as holds true for other proteins, protein transduction still suffers from general low efficiencies, substantial mislocalization of the transduced Nb to endosomes and cellular toxicity. Since both methods, i.e., genetic- and protein-based intrabody delivery, still have substantial drawbacks especially for in vivo applications, further efforts have to be undertaken to improve cellular delivery of intrabodies before these molecules can be efficiently used in clinical testing.

3. Intrabodies to Visualize Antigens in Living Cells

To visualize native cell structures and dynamic processes within living cells, imaging probes must not influence their target structures or intracellular processes. Thus, transiently binding intrabodies have become universal and versatile tools for live-cell imaging. Different strategies for how intrabodies have been used as imaging probes are summarized in the following (Figure 1).

3.1. Chromobody Technology

To generate detectable intrabodies for live-cell imaging, Nbs have been genetically coupled to fluorescent proteins such as eGFP or RFP and introduced as DNA-encoded expression constructs in living cells. Reflecting their chimeric structure, those fluorescently labeled imaging probes were termed “chromobodies” (Cbs) [54] (Figure 1a). With ~40 kDa, Cbs are rather small and diffusive binding molecules that can access different cellular compartments. In contrast to genetic fusions or covalent labeling, intracellular antigen detection is distinguished by distinct on- and off-rates of Cbs, whereby the protein of interest (POI) is not constantly bound by the immuno-label. As a first example, a GFP-specific Cb was generated by labeling a GFP-binding Nb with mRFP1 followed by functional expression in the cytoplasm of mammalian cells. In a proof-of-concept study, real-time fluorescent co-localization analysis successfully visualized dynamic redistribution of GFP-tagged antigens in different cellular compartments following the GFP-Cb signal. Additionally, the detection of endogenous structures using Cbs against LaminA/C and human cytokeratin-8 paved the way to establish Cbs as novel and versatile probes for live-cell imaging [54].
Reflecting the broad accessibility of GFP-fusion proteins, a diverse range of toolkits using the GFP-Nb as an effectively working intrabody have been developed throughout the last decade. For example, GFP-Nb was integrated into different Nb-sensor fusions for detection and manipulation of a variety of intracellular dynamics. Thereby, a range of fluorescent sensors for Ca2+ (RGECO), H+ (SepH, pHuji) and ATP/ADP (PHR) or multimerization proteins were genetically coupled to the GFP-Nb and functionally applied in the cytoplasm, mitochondria and ER of mammalian cells [55]. Furthermore, GFP-Nb was implemented in sophisticated protein degradation mechanisms either generating complete knockouts or fine-tunable and inducible mechanisms controlling the degradation process [56,57,58,59]. Likewise, targeted gene induction was realized by employing GFP-Nbs in GFP-dependent transcription systems [60,61]. Notably, GFP-Nb has also been applied for in vivo studies e.g., as part of the so-called LlamaTag used to image transcription factors in fly embryos [62], or as a binding moiety to facilitate targeted protein mislocalization, summarized as FrabFP [63]. As extensively reported, the GFP-Nb/GFP-Cb system can be easily implemented in different applications and is continuously used for proof-of-principle studies demonstrating the suitability of intrabodies to manipulate cells and organisms. However, it suffers from the need of GFP-labeled POIs and as a consequence, the addition of a relatively large protein portion (~25 kDa) either to the N- or the C-terminus of the POI, which may considerably affect its expression level, activity and localization compared to endogenous counterparts [64,65,66].

3.2. Tag-Specific Intrabodies

Considering that shorter epitopes might have only a minor impact on native protein folding, function and protein–protein interactions, extensive efforts have been undertaken to identify Nbs specifically recognizing small peptide-tags. Although most Nbs preferably address conformational epitopes [2,67,68,69], more recently a handful of Nbs recognizing linear epitopes were identified, mostly as byproducts from screening campaigns against full-sized proteins. Some of these Nbs have been exploited as novel capture and detection systems. For purification of proteins comprising the small EPEA-tag (also known as C-tag) [70], Myc-tag or BC2- (Spot-) tag [71,72], Nbs were converted into affinity matrices such as the CaptureSelectTM C-tag affinity matrix (www.thermofisher.com), the Myc-trap© or Spot-trap® resin (www.chromotek.com), respectively.
Notably, beside their functionality to capture and detect peptide-tags, Nbs against the ALFA-tag [73], Moon-tag [74,75] and Pep-tag [76] have also been successfully applied as intrabodies. In the Cb format, these binders visualize cytoskeletal structures like vimentin and actin, nuclear structures like PCNA and mitochondrial outer membrane proteins such as Miro1 in live cells [73,76]. In addition, real-time single mRNA translation of ribosomes could be optically monitored in individual cells using the Moon-tag system. In combination with the Sun-tag based on a scFv-derived intrabody, start site selection of ribosomes was visualized, uncovering the extensive heterogeneity of mRNA decoding [75]. Such Cbs targeting protein- or peptide-tags substantially expanded the broad applicability of recombinant protein tagging. The development of different tag-specific intrabodies now further offers the possibility to visualize different cellular antigens in a multiplex manner within living cells simultaneously. Most importantly, the usage of universal tagging strategies in combination with well-defined intrabodies is a simple and easy adaptable option to study many different cellular POIs, as one does not need to generate intrabodies for each individual POI. However, it has to be considered that these approaches still rely on genetic manipulation and in most cases employ an artificial expression of tagged recombinant POIs.

3.3. Intrabodies Targeting Endogenous Antigens

In the light of the above-mentioned drawbacks of protein tagging, in some cases intrabodies targeting endogenous POIs are highly beneficial. Here, components of the cytoskeleton have to be especially mentioned. Due to the highly coordinated structure of the cytoskeletal network and its dynamic adaption to intra- and extracellular stimuli, even small interferences derived from the addition of large fluorescent protein-tags or covalently binding molecules can disturb their correct formation and dynamic relocalization dramatically [64,77,78,79]. To visualize these components in their native condition, a variety of Cbs have been developed. Prominent examples for cytoskeletal imaging probes are lamin-Cb [80], actin-Cb [81] or vimentin-Cb [82]. These Cbs have successfully been used e.g., to image compound-induced apoptosis by monitoring the integrity of the nuclear lamina [83], study nuclear actin in mammalian cells [84] or to trace induction of the epithelial–mesenchymal transition in response to external stimuli using vimentin as a cellular biomarker visualized by vimentin-specific Cbs [82,85]. Notably, actin-Cbs additionally visualized cytoskeleton remodeling in vivo e.g., during the different developmental stages in zebrafish embryos [81] or neuronal actin plasticity in mouse [32].
In combination with high-content real-time imaging, Cbs and Cb comprising cell models were further applied in screening campaigns to study compound effects on dynamic cellular processes tightly regulated by endogenous proteins. In this context, details of S phase progression and endogenous DNA replication was impressively revealed by using a PCNA-Cb [86] which also enabled high-throughput screening libraries for cell-cycle-affecting compounds [87]. Similarly, DNA-damaging agents were identified by monitoring the signal of a PARP1-Cb in live cells [88]. Moreover, histone-targeting Cbs visualized DNA-damaging sites upon detection of post-translational modifications in cells [89] or labelled chromatin structures in organisms [90]. To monitor real-time induction of the Wnt/β-catenin signaling pathway, which is highly important for embryonic development and is a key player in carcinogenesis, a Cb (BC1-Cb) which addresses cancer-related hypophosphorylated β-catenin was implemented in cellular models to study the effect of small molecules modulating the Wnt/β-catenin pathway [91,92].
To visualize the dynamics and native distribution of endogenous antigens, Cbs have to fulfil specific requirements: most importantly, they must not modulate the intrinsic function and localization of their antigen e.g., by binding catalytic sites or displacing central interaction partners. Furthermore, high-affinity Cbs can affect the dynamics and cellular distribution of their addressed antigens. In consequence, intracellular binding properties of Cbs have to be carefully analyzed. This can be addressed e.g., by targeted selection of Cbs addressing inert epitopes or selecting transiently binding Cbs, detectable by FRAP (fluorescent recovery after photobleaching) analysis [54,76,81,82]. Additionally, intracellular immune-precipitations (ICIPs) can be performed to analyze the proteome of Cbs [82,91,93]. In the case of stable Cb cell lines, these further require detailed phenotypic analysis evaluating their morphology, proliferation and signaling pathways in comparison to respective wild-type cells [82]. Notably, cellular Cb levels are only partially adjustable and the signal of bound Cbs can be outshined by the diffuse signal derived from non-bound Cbs. In this context, recently a phenomenon describing reduced proteasomal degradation of antigen-bound Cbs, summarized as antigen-mediated chromobody stabilization (AMCBS), was observed [93]. Recent studies further reveal how minor changes within the Cb sequence lead to turnover-accelerated [93], conditionally stable [94] or enhancer Cbs [95]. By implementation of these developments, Cbs have become more versatile imaging probes which not only visualize their endogenous antigens but can also be used to quantify changes in endogenous protein concentration in living cells [76,93].

3.4. Intrabodies as Biosensors

Considering that intracellular processes are highly complex and visualization is not always sufficient to study this their multifaceted nature, Nb-based biosensors have been developed which exhibit their “function” only after activation of the target protein. Such cellular biosensors have been successfully developed and applied e.g., to monitor activation of GPCRs in cellular models. In a pioneering work, Irannejad et al. generated a Cb (Nb80-GFP) specifically recognizing the conformational active form of the β2-adrenergic receptor (β2AR) in living cells. Upon activation of β2-AR with isoprenaline, this intracellular biosensor relocalizes from a diffuse fraction to the plasma membrane. Prolonged time-lapse imaging further revealed a displacement of the Cb to internalized β2-AR after binding of β-arrestin, followed by its redistribution to β2-AR-containing endocytic vesicles [96]. To generate biosensors against CCR7, the CDR1 and CDR3 regions of the Nb80 were randomized and three novel intrabodies specific for the cytoplasmic domain of CCR7 were selected from the synthetic Nb library. Upon fusion of these Nbs to split fragments of YFP and following the signal derived from bi-fluorescent complementation (BiFC), ligand-induced trafficking of CCR7 became detectable with high spatiotemporal resolution in living cells [97]. Similarly, intrabodies specific for active/agonist-bound state of opioid receptors (ORs) were identified, which monitored ligand-induced activation of ORs and revealed distortion of neuronal opioid receptor distribution caused by neuromodulating drugs [98]. Furthermore, via the RHO-Nb-GFP fusion protein, a BRET-based biosensor was developed, dynamically monitoring real-time RHO activation [99]. Another approach exploited Nbs as a diagnostic tool for sensing influenza A virus infection in a Nb-based sandwich reporter system. Via the nucleoprotein (NP)-dependent reporter gene transcription activation using NP-specific Nbs, infection of various influenza A subtypes in living cells has been monitored [100].

4. Intrabodies to Modulate and Manipulate Intracellular Antigens

In addition to visualization, intrabodies can be selected to influence signaling pathways or function of cellular targets. Activating or blocking intrabodies have been developed for a range of POIs including enzymes, oncogenic proteins, proteins of the nervous system, virus proteins and toxins [101]. So far, many of those targets are still not druggable with small molecules. Therefore, intrabodies with high affinities and specificities are considered as interesting alternatives. In the following, a summary of promising intrabodies are depicted underlining their potential for preclinical research and clinical application (Figure 2).

4.1. Intrabodies in Oncology

The most common strategies to modulate key drivers and intracellular signaling pathways in tumor therapy still rely on small molecules which often suffer from off-target binding which can cause severe side effects and overall toxicity [102,103]. Here, the specificity of intrabodies could be highly beneficial. A main driver in many cancer types is the deregulation of the tumor suppressor p53. By equipping an Nb binding the N-terminal transactivation domain of p53 with a mitochondrial localization signal, wild-type p53 was delocalized to mitochondria which affects the viability of tumor cell lines [104]. In contrast, a Nb targeting the DNA-binding domain of p53 was shown to stabilize endogenous p53 in a HPV-infected cervical cancer cell model [105]. In addition, GPCRs are well known to regulate a variety of cancer-associated signaling pathways like the PI3-protein kinase/AKT and MAPK pathway, therefore tight regulation of GPCRs is essential to block disease-related signaling. With Nb5, an intrabody which tightly binds to the Gβγ dimer and responds to all combinations of β- and γ subtypes competing with endogenous regulatory proteins, a control switch was identified. Due its potency to prevent Gβγ-mediated signaling events, this intrabody can be considered as a potential candidate for the treatment of different diseases driven by deregulated GPCRs [106]. In a similar context, in vitro studies were performed with protein kinase Cε modulating Nbs, either increasing or inhibiting its kinase activity on a cellular basis which therefore also might represent promising reagents for therapeutic applications [107]. Invasive and metastatic cancer types pointedly damaging surrounding tissues are non-operable and are therefore highly challenging. Consequently, advanced therapeutic approaches to stop cell migration and invasion are heavily needed. Recently, intrabodies targeting important regulators of formation of invadopodia such as fascin, cortactin and N-WASp (neural Wiskott–Aldrich syndrome protein) have been identified and applied in cellular live-cell studies. These studies showed that in the presence of such intrabodies, actin bundling can efficiently be disrupted in different cancer cell lines (including breast, prostate, head and neck cancer), hindering the formation of properly organized invadopodia [108,109,110,111]. Many of the examples mentioned above exhibit promising potential for cancer treatment in cellulo, however, before their maturation into clinical testing, these intrabodies have to prove their potency in living organisms. An encouraging example represents the F-actin caping protein (CapG)-specific Nb, which efficiently prevented migration of breast cancer metastasis in immune-deficient mice [112]. Another recently described STAT-3 binding intrabody suppresses the function of phosphorylated STAT-3 and distinctly reduces cell proliferation in vitro and breast cancer growth in mouse xenografts [113]. In summary, multiple Nb-based intrabodies are emerging as interesting candidates for the treatment of various malignancies, emphasizing the potential of these binding molecules for future clinical application in oncology.

4.2. Intrabodies as Immune Modulators

According to the strong link between inflammation and development of malignancies, many intrabodies initially identified and analyzed in cancer models could also be applicable in the modulation of key players of immune signaling pathways. A classical target for immune modulators is the inflammasome, which in case of dysregulation causes a range of inflammatory diseases like arthritis, gout or diabetes. Therefore, obstruction of this innate immune regulator is highly interesting for many applications. A CARD (caspase activation and recruitment domain)-specific intrabody was shown to suppress assembly of the inflammasome which leads to a reduced secretion of proinflammatory cytokines. Furthermore, applied as Cb this intrabody visualized for the first time the filamentous structure of the adaptor protein ASC in living cells [114]. Additionally, activation of immune cells strongly relies on extensive cytoskeleton remodeling. Notable examples are podosome formation in macrophages and formation of an immunological synapse between T- and antigen-presenting cells. In this process, L-plastin, an actin-bundling protein specific to leukocytes, orchestrates assembly and turnover of actin filaments. De Clerque et al. showed that intrabodies trapping L-plastin in the active conformation induced T cell proliferation as well as cytokine secretion [115], whereas inhibiting intrabodies caused severe dysfunction of macrophages [116,117]. However, gelsolin-specific intrabodies applied in a similar approach showed no direct effect on podosome formation [116,118]. In summary, intrabodies manipulating immune-related targets offer new opportunities and advanced options exceeding current standards of care.

4.3. Intrabodies to Address Neurological Disorders

Up to now, despite intensive research efforts, curative therapies are available only for a limited number of neurological disorders and neurodegenerative diseases. Thus, unconventional alternatives came into focus such as the application of intrabodies for the most prominent examples like Parkinson’s disease and Alzheimer’s disease [119]. Consequently, selection strategies have been adapted and validation pipelines for intrabody application have been developed to generate intrabodies specifically addressing targets in mammalian brain neurons [120]. Clearly, these developments are currently still in their infancy but encouraging examples are summarized in the following.
Actuators of many neurodegenerative diseases are mutated and misfolded proteins prone to aggregation continuously destroying parts of the brain and mental functions. In Parkinson’s disease, aggregated α-synuclein (α-syn) is the disruptive element and therefore the most promising target for therapeutic approaches. Ongoing research efforts including the application of computational affinity maturation [121] have led to the development of advanced α-syn intrabodies recognizing pathogenic aggregates at different stages of fibril maturation during disease progression [122] as well as post-translationally modified forms of α-syn [123]. The most promising Nb (NbSyn87) was linked to a PEST motif and induces a targeted degradation of α-syn aggregates in cellular models showing beneficial effects on proteostatic stress and cellular toxicity [124]. By this approach, the authors impressively demonstrated the general applicability of intrabodies fused to PEST domains for degradation of their respective antigens [124]. Moreover, upon viral vector-mediated delivery of this intrabody into a rat Parkinson’s disease model, an efficient removal of misfolded and aggregated α-syn in vivo was shown, turning this strategy into a highly interesting treatment option for synucleinopathies like Parkinson’s disease [33]. Regarding another aspect, oxidative stress-induced apoptosis in non-regenerative tissues such as the brain causes severe damage due to loss of functional cells. This cell death mechanism is associated with pathologies like Alzheimer’s and Parkinson’s disease. In this context, intrabodies targeting the pro-apoptotic protein Bax were demonstrated to block both mitochondrial membrane potential collapse and apoptosis after oxidative stress, bringing those tools into play as novel therapeutics [125]. Dysregulation of ion channel functions are also associated with several neurological and neuromuscular diseases. Many of these ion channels are directly regulated by GPCRs, therefore the Gβ-inhibiting intrabody identified by Gulati et al. also might be a potential candidate for therapy of Parkinson’s and Alzheimer’s disease or multiple sclerosis [106]. High-voltage-activated calcium channels (HVACCs) represent another class of ion channels important for neurological diseases like epilepsy, chronic pain and Parkinson’s disease. Here, HVACC-specific intrabodies in combination with an E3 ubiquitin ligase were successfully applied for targeted ion channel depletion. Expression of these intrabodies results in efficient current ablation, proving the potential of intrabodies to modulate ion channels [126]. To address vesicular glutamate transporters (VGLUTs) playing a key role in excitatory neurotransmission, intrabodies addressing the cytosolic domain of ratVGLUT1 have been developed and have been successfully shown to reduce the uptake of glutamate in reconstituted liposomes and subcellular fractions enriched with synaptic vesicles in vitro [127].

4.4. Intrabodies Inhibiting Viral and Bacterial Pathogens in Live Cells

Considering intrabodies as inhibitors of viral or bacterial infections, human immunodeficiency virus (HIV) is the most prominent target when acknowledging the perpetual lack of a virus-depleting cure [128]. Up to now, a range intrabodies targeting different HIV-specific antigens including Rev [129,130], Vpr [131] and Nef [132,133] have been developed. Notably, Rev-specific intrabodies were shown to successfully inhibit viral replication and partly lower infectivity of viral particles as demonstrated in different cellular models [129,130]. With Nef intrabodies, even in vivo Nef-mediated effects could be reversed in transgenic mice, underlining the potential of intrabodies to fight against amplification and propagation of HIV in living organisms [132].
Due to mutational escape, the influenza virus represents a recurring seasonal predator. Therefore, in addition to vaccination, further possibilities to contain its spread are under development. Here, the nucleoprotein turned out to be an efficient target for generating intrabody-based inhibitors of influenza and a variety of different candidates were selected for disrupting virus replication in living cells [134,135]. Along this line, advanced treatment options fighting viral-induced hepatitis are urgently needed especially with the lack of efficient vaccinations e.g., for hepatitis virus C (HCV) which causes chronic infections with the risk of hepatocellular carcinoma. Several studies convincingly show that intrabodies targeting non-structural (NS) active proteins like NS3, NS4B and NS5B or the core protein, inhibit HCV replication [136,137,138]. Additionally, intrabodies addressing the core antigen of hepatitis B virus negatively affected expression and trafficking of the target antigen within live cells [139]. Facing an increasing number of antibiotic-resistant bacteria strains, a change of thinking towards alternative options is necessary. Thereby, intrabodies neutralizing bacterial toxins could become interesting for this field. Two promising examples of intrabodies addressing the SpvB toxin secreted by Salmonella typhimurium [140] and Botulinum neurotoxin [141] showed how such binders can rescue host cells and therefore function as potential antidotes.
In summary, this chapter comprises only a short and non-comprehensive summary of recent findings on inhibitory intrabodies and their application in different preclinical disease models. Nevertheless, these developments impressively demonstrate how modulating intrabodies can successfully address a plethora of different cellular targets and contribute not only to a more detailed understanding of disease-related processes but also illustrate their potential for future therapeutic applications.

5. Switchable Intrabodies as Upcoming Tools

A major disadvantage of intrabodies is their non-adjustable binding after expression. Therefore, switchable intrabodies are not only interesting for investigating cellular mechanisms but also open new opportunities for a controlled treatment which increases safety measurements substantially. Recently, different types of switchable systems were developed using light or small molecules as activators, termed as optogenetic or chemogenetic controls, respectively. Optogenetic switchable intrabodies were either realized by Nb fusions with light-responsive proteins [57,142] or by engineering light-controllable opto-intrabodies [143]. The first approach was developed for tunable light-induced depletion of target proteins and employs an E3 ubiquitin ligase fusion with CIBN (cryptochrome-interacting basic-helix-loop-helix 1 N-terminus) in combination with POI-specific intrabodies fused to Cryptochrome 2 (CRYP2). In the presence of short light pulses, conformational changes of the cryptochrome-associated proteins were triggered, thereby causing heterodimerization of both fusion proteins. Using this approach, intrabodies binding GFP, nuclear LaminA/C and PCNA were linked to the E3 ubiquitin ligase causing depletion of the respective target proteins [57]. Alternatively, to induce (switch on) intracellular binding, intrabodies targeting different POIs were split in two parts and fused to light-responsive proteins. Upon application of respective light-derived stimuli, their function was reactivated by bringing both halves together [142]. Similarly, Gil et al. engineered light-switchable Nbs by inserting the light-sensitive AsLOV2 domain into a solvent-exposed loop of an intrabody. Depending on the insertion position, light-induced binding or dissociation due to conformational changes was observed. With these developments a highly versatile photoswitchable nanobody toolbox enabling programmable regulation was established [143].
In contrast to light switchable systems, chemogenetic controllable systems bear the advantage of a stepwise regulation through precise titration of the active compound. Recently, such ligand-modulated binders have been generated by insertion of permutated bacterial dihydrofolate reductase (cpDHFR) in the CDR3 region of intrabodies. While the unstructured cpDHFR does not affect binding of the intrabody, addition of small compounds such as NADPH or trimethoprim changes the conformation of the cpDHFR, which results in the release of the bound antigen. This reversible process was nicely shown for several intrabodies inducing e.g., targeted antigen relocalization in living cells. It impressively demonstrates how intrabodies can be turned into switchable binders by the addition of small, non-toxic compounds [144]. Additionally, controlling protein function via switchable intrabodies in vivo is of great interest, thus providing the possibility of tunable phenotypes. This was successfully demonstrated by Deng et al., showing rapamycin-induced depletion of GFP-tagged CED-3 using the GFP-binding intrabody in combination with the E3 ubiquitin ligase, each fused to a rapamycin-responsive element. Following this strategy in a C. elegans model, the authors demonstrated how such an approach can be applied to influence e.g., induction of apoptosis in developing animals [57].
In summary, the field of Nbs validated for intracellular applications has widely increased over recent years, not only because of the ongoing development of advanced screening technologies for selection and generation of tailor-made intrabodies, but also due to their unique features and capabilities to monitor and manipulate cellular processes, thereby providing distinct advantages compared to conventional technologies such as fluorescent fusion proteins, siRNA, knock down genetics or chemical strategies using small molecules. Whereas non-modifying intrabodies for visualization purposes like Cbs have become state of the art versatile research tools, intrabodies for functional manipulation of disease-related antigens gain more and more importance for preclinical research. The most recent examples of intrabodies which modulate their target structures in living organisms using a switchable binding mechanism impressively provide a perspective on how intrabodies can be applied in advanced clinical research and can be implemented in novel therapeutic strategies, which are urgently needed for more personalized medicine in future.

Funding

This research was funded by the German Research Foundation (DFG), grant number RTG 2364 “MOMbrane”.

Acknowledgments

The authors gratefully acknowledge support by Open Access Publishing Fund of University of Tuebingen.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hamers-Casterman, C.; Atarhouch, T.; Muyldermans, S.; Robinson, G.; Hamers, C.; Songa, E.B.; Bendahman, N.; Hamers, R. Naturally occurring antibodies devoid of light chains. Nature 1993, 363, 446–448. [Google Scholar] [CrossRef] [PubMed]
  2. Muyldermans, S. Nanobodies: Natural single-domain antibodies. Annu. Rev. Biochem. 2013, 82, 775–797. [Google Scholar] [CrossRef] [Green Version]
  3. Peyvandi, F.; Scully, M.; Kremer Hovinga, J.A.; Cataland, S.; Knobl, P.; Wu, H.; Artoni, A.; Westwood, J.P.; Mansouri Taleghani, M.; Jilma, B.; et al. Caplacizumab for acquired thrombotic thrombocytopenic purpura. N. Engl. J. Med. 2016, 374, 511–522. [Google Scholar] [CrossRef] [PubMed]
  4. Morrison, C. Nanobody approval gives domain antibodies a boost. Nat. Rev. Drug Discov. 2019, 18, 485–487. [Google Scholar] [CrossRef] [PubMed]
  5. Dumoulin, M.; Conrath, K.; Van Meirhaeghe, A.; Meersman, F.; Heremans, K.; Frenken, L.G.; Muyldermans, S.; Wyns, L.; Matagne, A. Single-domain antibody fragments with high conformational stability. Protein Sci. 2002, 11, 500–515. [Google Scholar] [CrossRef] [PubMed]
  6. Kunz, P.; Zinner, K.; Mücke, N.; Bartoschik, T.; Muyldermans, S.; Hoheisel, J.D. The structural basis of nanobody unfolding reversibility and thermoresistance. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef] [PubMed]
  7. Kaiser, P.D.; Maier, J.; Traenkle, B.; Emele, F.; Rothbauer, U. Recent progress in generating intracellular functional antibody fragments to target and trace cellular components in living cells. Biochim. Biophys. Acta 2014, 1844, 1933–1942. [Google Scholar] [CrossRef]
  8. Helma, J.; Cardoso, M.C.; Muyldermans, S.; Leonhardt, H. Nanobodies and recombinant binders in cell biology. J. Cell Biol. 2015, 209, 633–644. [Google Scholar] [CrossRef] [Green Version]
  9. Traenkle, B.; Rothbauer, U. Under the microscope: Single-domain antibodies for live-cell imaging and super-resolution microscopy. Front. Immunol. 2017, 8, 1030. [Google Scholar] [CrossRef] [Green Version]
  10. Moutel, S.; Bery, N.; Bernard, V.; Keller, L.; Lemesre, E.; de Marco, A.; Ligat, L.; Rain, J.C.; Favre, G.; Olichon, A.; et al. Nali-h1: A universal synthetic library of humanized nanobodies providing highly functional antibodies and intrabodies. eLife 2016, 5, e16228. [Google Scholar] [CrossRef]
  11. Wörn, A.; Plückthun, A. Mutual stabilization of vl and vh in single-chain antibody fragments, investigated with mutants engineered for stability. Biochemistry 1998, 37, 13120–13127. [Google Scholar] [CrossRef] [PubMed]
  12. Proba, K.; WoÈrn, A.; Honegger, A.; PluÈckthun, A. Antibody scfv fragments without disulfide bonds, made by molecular evolution. J. Mol. Biol. 1998, 275, 245–253. [Google Scholar] [CrossRef] [PubMed]
  13. Olichon, A.; Surrey, T. Selection of genetically encoded fluorescent single domain antibodies engineered for efficient expression in escherichia coli. J. Biol. Chem. 2007, 282, 36314–36320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kvam, E.; Sierks, M.R.; Shoemaker, C.B.; Messer, A. Physico-chemical determinants of soluble intrabody expression in mammalian cell cytoplasm. Protein Eng. Des. Sel. 2010, 23, 489–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Joshi, S.N.; Butler, D.C.; Messer, A. Fusion to a highly charged proteasomal retargeting sequence increases soluble cytoplasmic expression and efficacy of diverse anti-synuclein intrabodies. In MAbs; Taylor & Francis: London, UK, 2012; pp. 686–693. [Google Scholar] [CrossRef] [Green Version]
  16. Kabayama, H.; Takeuchi, M.; Tokushige, N.; Muramatsu, S.-I.; Kabayama, M.; Fukuda, M.; Yamada, Y.; Mikoshiba, K. An ultra-stable cytoplasmic antibody engineered for in vivo applications. Nat. Commun. 2020, 11, 1–20. [Google Scholar] [CrossRef] [PubMed]
  17. Li, T.; Bourgeois, J.P.; Celli, S.; Glacial, F.; Le Sourd, A.M.; Mecheri, S.; Weksler, B.; Romero, I.; Couraud, P.O.; Rougeon, F.; et al. Cell-penetrating anti-gfap vhh and corresponding fluorescent fusion protein vhh-gfp spontaneously cross the blood-brain barrier and specifically recognize astrocytes: Application to brain imaging. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2012, 26, 3969–3979. [Google Scholar] [CrossRef]
  18. Fields, S.; Song, O. A novel genetic system to detect protein-protein interactions. Nature 1989, 340, 245–246. [Google Scholar] [CrossRef]
  19. Young, K. Yeast two-hybrid: So many interactions, (in) so little time…. Biol. Reprod. 1998, 58, 302–311. [Google Scholar] [CrossRef] [Green Version]
  20. Brückner, A.; Polge, C.; Lentze, N.; Auerbach, D.; Schlattner, U. Yeast two-hybrid, a powerful tool for systems biology. Int. J. Mol. Sci. 2009, 10, 2763–2788. [Google Scholar] [CrossRef] [Green Version]
  21. Pellis, M.; Muyldermans, S.; Vincke, C. Bacterial two hybrid: A versatile one-step intracellular selection method. Methods Mol. Biol. 2012, 911, 135–150. [Google Scholar]
  22. Tanaka, T.; Rabbitts, T.H. Intracellular antibody capture (iac) methods for single domain antibodies. In Single Domain Antibodies; Springer: Cham, Switzerland, 2012; pp. 151–173. [Google Scholar]
  23. Visintin, M.; Tse, E.; Axelson, H.; Rabbitts, T.H.; Cattaneo, A. Selection of antibodies for intracellular function using a two-hybrid in vivo system. Proc. Natl. Acad. Sci. USA 1999, 96, 11723–11728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Visintin, M.; Settanni, G.; Maritan, A.; Graziosi, S.; Marks, J.D.; Cattaneo, A. The intracellular antibody capture technology (iact): Towards a consensus sequence for intracellular antibodies. J. Mol. Biol. 2002, 317, 73–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Tanaka, T.; Lobato, M.N.; Rabbitts, T.H. Single domain intracellular antibodies: A minimal fragment for direct in vivo selection of antigen-specific intrabodies. J. Mol. Biol. 2003, 331, 1109–1120. [Google Scholar] [CrossRef]
  26. Pellis, M.; Pardon, E.; Zolghadr, K.; Rothbauer, U.; Vincke, C.; Kinne, J.; Dierynck, I.; Hertogs, K.; Leonhardt, H.; Messens, J.; et al. A bacterial-two-hybrid selection system for one-step isolation of intracellularly functional nanobodies. Arch. Biochem. Biophys. 2012, 526, 114–123. [Google Scholar] [CrossRef] [PubMed]
  27. Zolghadr, K.; Mortusewicz, O.; Rothbauer, U.; Kleinhans, R.; Goehler, H.; Wanker, E.E.; Cardoso, M.C.; Leonhardt, H. A fluorescent two-hybrid assay for direct visualization of protein interactions in living cells. Mol. Cell. Proteom. MCP 2008, 7, 2279–2287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Schmidt, F.I.; Hanke, L.; Morin, B.; Brewer, R.; Brusic, V.; Whelan, S.P.; Ploegh, H.L. Phenotypic lentivirus screens to identify functional single domain antibodies. Nat. Microbiol. 2016, 1, 16080. [Google Scholar] [CrossRef] [Green Version]
  29. Mazuc, E.; Guglielmi, L.; Bec, N.; Parez, V.; Hahn, C.S.; Mollevi, C.; Parrinello, H.; Desvignes, J.-P.; Larroque, C.; Jupp, R. In-cell intrabody selection from a diverse human library identifies c12orf4 protein as a new player in rodent mast cell degranulation. PLoS ONE 2014, 9, e104998. [Google Scholar] [CrossRef]
  30. Lee, S.; Kaku, Y.; Inoue, S.; Nagamune, T.; Kawahara, M. Growth signalobody selects functional intrabodies in the mammalian cytoplasm. Biotechnol. J. 2016, 11, 565–573. [Google Scholar] [CrossRef]
  31. Keller, B.-M.; Maier, J.; Weldle, M.; Segan, S.; Traenkle, B.; Rothbauer, U. A strategy to optimize the generation of stable chromobody cell lines for visualization and quantification of endogenous proteins in living cells. Antibodies 2019, 8, 10. [Google Scholar] [CrossRef] [Green Version]
  32. Wegner, W.; Ilgen, P.; Gregor, C.; Van Dort, J.; Mott, A.C.; Steffens, H.; Willig, K.I. In vivo mouse and live cell sted microscopy of neuronal actin plasticity using far-red emitting fluorescent proteins. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  33. Chatterjee, D.; Bhatt, M.; Butler, D.; De Genst, E.; Dobson, C.M.; Messer, A.; Kordower, J.H. Proteasome-targeted nanobodies alleviate pathology and functional decline in an α-synuclein-based parkinson’s disease model. npj Park. Dis. 2018, 4, 1–10. [Google Scholar] [CrossRef] [Green Version]
  34. Hammond, C.; Helenius, A. Quality control in the secretory pathway: Retention of a misfolded viral membrane glycoprotein involves cycling between the er, intermediate compartment, and golgi apparatus. J. Cell Biol. 1994, 126, 41–52. [Google Scholar] [CrossRef]
  35. Slastnikova, T.A.; Ulasov, A.V.; Rosenkranz, A.A.; Sobolev, A.S. Targeted intracellular delivery of antibodies: The state of the art. Front. Pharmacol. 2018, 9, 1208. [Google Scholar] [CrossRef] [Green Version]
  36. Singh, K.; Ejaz, W.; Dutta, K.; Thayumanavan, S. Antibody delivery for intracellular targets: Emergent therapeutic potential. Bioconjugate Chem. 2019, 30, 1028–1041. [Google Scholar] [CrossRef]
  37. Li, Y.; Li, P.; Li, R.; Xu, Q. Intracellular antibody delivery mediated by lipids, polymers, and inorganic nanomaterials for therapeutic applications. Adv. Ther. 2020, 2000178. [Google Scholar] [CrossRef]
  38. Conic, S.; Desplancq, D.; Ferrand, A.; Fischer, V.; Heyer, V.; Reina San Martin, B.; Pontabry, J.; Oulad-Abdelghani, M.; Babu, N.K.; Wright, G.D.; et al. Imaging of native transcription factors and histone phosphorylation at high resolution in live cells. J. Cell Biol. 2018, 217, 1537–1552. [Google Scholar] [CrossRef] [Green Version]
  39. Dixon, C.R.; Platani, M.; Makarov, A.A.; Schirmer, E.C. Microinjection of antibodies targeting the lamin a/c histone-binding site blocks mitotic entry and reveals separate chromatin interactions with hp1, cenpb and pml. Cells 2017, 6, 9. [Google Scholar] [CrossRef] [Green Version]
  40. Klein, A.; Hank, S.; Raulf, A.; Joest, E.; Tissen, F.; Heilemann, M.; Wieneke, R.; Tampé, R. Live-cell labeling of endogenous proteins with nanometer precision by transduced nanobodies. Chem. Sci. 2018, 9, 7835–7842. [Google Scholar] [CrossRef] [Green Version]
  41. Frankel, A.D.; Pabo, C.O. Cellular uptake of the tat protein from human immunodeficiency virus. Cell 1988, 55, 1189–1193. [Google Scholar] [CrossRef]
  42. Green, M.; Loewenstein, P.M. Autonomous functional domains of chemically synthesized human immunodeficiency virus tat trans-activator protein. Cell 1988, 55, 1179–1188. [Google Scholar] [CrossRef]
  43. Joliot, A.; Pernelle, C.; Deagostini-Bazin, H.; Prochiantz, A. Antennapedia homeobox peptide regulates neural morphogenesis. Proc. Natl. Acad. Sci. USA 1991, 88, 1864–1868. [Google Scholar] [CrossRef] [Green Version]
  44. Futaki, S.; Suzuki, T.; Ohashi, W.; Yagami, T.; Tanaka, S.; Ueda, K.; Sugiura, Y. Arginine-rich peptides an abundant source of membrane-permeable peptides having potential as carriers for intracellular protein delivery. J. Biol. Chem. 2001, 276, 5836–5840. [Google Scholar] [CrossRef] [Green Version]
  45. Futaki, S. Arginine-rich peptides: Potential for intracellular delivery of macromolecules and the mystery of the translocation mechanisms. Int. J. Pharm. 2002, 245, 1–7. [Google Scholar] [CrossRef]
  46. Prochiantz, A. Messenger proteins: Homeoproteins, tat and others. Curr. Opin. Cell Biol. 2000, 12, 400–406. [Google Scholar] [CrossRef]
  47. Kaczmarczyk, S.J.; Sitaraman, K.; Young, H.A.; Hughes, S.H.; Chatterjee, D.K. Protein delivery using engineered virus-like particles. Proc. Natl. Acad. Sci. USA 2011, 108, 16998–17003. [Google Scholar] [CrossRef] [Green Version]
  48. Méndez, J.; Morales Cruz, M.; Delgado, Y.; Figueroa, C.M.; Orellano, E.A.; Morales, M.; Monteagudo, A.; Griebenow, K. Delivery of chemically glycosylated cytochrome c immobilized in mesoporous silica nanoparticles induces apoptosis in hela cancer cells. Mol. Pharm. 2014, 11, 102–111. [Google Scholar] [CrossRef] [Green Version]
  49. Ray, M.; Tang, R.; Jiang, Z.; Rotello, V.M. Quantitative tracking of protein trafficking to the nucleus using cytosolic protein delivery by nanoparticle-stabilized nanocapsules. Bioconjugate Chem. 2015, 26, 1004–1007. [Google Scholar] [CrossRef] [Green Version]
  50. Bruce, V.J.; Lopez-Islas, M.; McNaughton, B.R. Resurfaced cell-penetrating nanobodies: A potentially general scaffold for intracellularly targeted protein discovery. Protein Sci. 2016, 25, 1129–1137. [Google Scholar] [CrossRef] [Green Version]
  51. Lee, Y.; Ishii, T.; Kim, H.J.; Nishiyama, N.; Hayakawa, Y.; Itaka, K.; Kataoka, K. Efficient delivery of bioactive antibodies into the cytoplasm of living cells by charge-conversional polyion complex micelles. Angew. Chem. Int. Ed. 2010, 49, 2552–2555. [Google Scholar] [CrossRef]
  52. Sarker, S.R.; Hokama, R.; Takeoka, S. Intracellular delivery of universal proteins using a lysine headgroup containing cationic liposomes: Deciphering the uptake mechanism. Mol. Pharm. 2014, 11, 164–174. [Google Scholar] [CrossRef]
  53. Röder, R.; Helma, J.; Preiß, T.; Rädler, J.O.; Leonhardt, H.; Wagner, E. Intracellular delivery of nanobodies for imaging of target proteins in live cells. Pharm. Res. 2017, 34, 161–174. [Google Scholar] [CrossRef] [PubMed]
  54. Rothbauer, U.; Zolghadr, K.; Tillib, S.; Nowak, D.; Schermelleh, L.; Gahl, A.; Backmann, N.; Conrath, K.; Muyldermans, S.; Cardoso, M.C. Targeting and tracing antigens in live cells with fluorescent nanobodies. Nat. Methods 2006, 3, 887–889. [Google Scholar] [CrossRef] [PubMed]
  55. Prole, D.L.; Taylor, C.W. A genetically encoded toolkit of functionalized nanobodies against fluorescent proteins for visualizing and manipulating intracellular signalling. BMC Biol. 2019, 17, 1–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Caussinus, E.; Kanca, O.; Affolter, M. Fluorescent fusion protein knockout mediated by anti-gfp nanobody. Nat. Struct. Mol. Biol. 2011, 19, 117–121. [Google Scholar] [CrossRef] [PubMed]
  57. Deng, W.; Bates, J.A.; Wei, H.; Bartoschek, M.D.; Conradt, B.; Leonhardt, H. Tunable light and drug induced depletion of target proteins. Nat. Commun. 2020, 11, 1–13. [Google Scholar] [CrossRef] [Green Version]
  58. Daniel, K.; Icha, J.; Horenburg, C.; Müller, D.; Norden, C.; Mansfeld, J. Conditional control of fluorescent protein degradation by an auxin-dependent nanobody. Nat. Commun. 2018, 9, 1–13. [Google Scholar] [CrossRef]
  59. Clift, D.; McEwan, W.A.; Labzin, L.I.; Konieczny, V.; Mogessie, B.; James, L.C.; Schuh, M. A method for the acute and rapid degradation of endogenous proteins. Cell 2017, 171, 1692–1706.e18. [Google Scholar] [CrossRef] [Green Version]
  60. Tang, J.C.; Rudolph, S.; Dhande, O.S.; Abraira, V.E.; Choi, S.; Lapan, S.W.; Drew, I.R.; Drokhlyansky, E.; Huberman, A.D.; Regehr, W.G.; et al. Cell type-specific manipulation with gfp-dependent cre recombinase. Nat. Neurosci. 2015, 18, 1334–1341. [Google Scholar] [CrossRef] [Green Version]
  61. Tang, J.C.; Szikra, T.; Kozorovitskiy, Y.; Teixiera, M.; Sabatini, B.L.; Roska, B.; Cepko, C.L. A nanobody-based system using fluorescent proteins as scaffolds for cell-specific gene manipulation. Cell 2013, 154, 928–939. [Google Scholar] [CrossRef] [Green Version]
  62. Bothma, J.P.; Norstad, M.R.; Alamos, S.; Garcia, H.G. Llamatags: A versatile tool to image transcription factor dynamics in live embryos. Cell 2018, 173, 1810–1822.e16. [Google Scholar] [CrossRef]
  63. Harmansa, S.; Alborelli, I.; Bieli, D.; Caussinus, E.; Affolter, M. A nanobody-based toolset to investigate the role of protein localization and dispersal in drosophila. eLife 2017, 6, e22549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hosein, R.E.; Williams, S.A.; Haye, K.; Gavin, R.H. Expression of gfp-actin leads to failure of nuclear elongation and cytokinesis in tetrahymena thermophila. J. Eukaryot. Microbiol. 2003, 50, 403–408. [Google Scholar] [CrossRef] [PubMed]
  65. Snapp, E.L. Fluorescent proteins: A cell biologist’s user guide. Trends Cell Biol. 2009, 19, 649–655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Stadler, C.; Rexhepaj, E.; Singan, V.R.; Murphy, R.F.; Pepperkok, R.; Uhlen, M.; Simpson, J.C.; Lundberg, E. Immunofluorescence and fluorescent-protein tagging show high correlation for protein localization in mammalian cells. Nat. Methods 2013, 10, 315–323. [Google Scholar] [CrossRef]
  67. De Genst, E.; Silence, K.; Decanniere, K.; Conrath, K.; Loris, R.; Kinne, J.; Muyldermans, S.; Wyns, L. Molecular basis for the preferential cleft recognition by dromedary heavy-chain antibodies. Proc. Natl. Acad. Sci. USA 2006, 103, 4586–4591. [Google Scholar] [CrossRef] [Green Version]
  68. Pardon, E.; Laeremans, T.; Triest, S.; Rasmussen, S.G.; Wohlkonig, A.; Ruf, A.; Muyldermans, S.; Hol, W.G.; Kobilka, B.K.; Steyaert, J. A general protocol for the generation of nanobodies for structural biology. Nat. Protoc. 2014, 9, 674–693. [Google Scholar] [CrossRef]
  69. Nunes-Silva, S.; Gangnard, S.; Vidal, M.; Vuchelen, A.; Dechavanne, S.; Chan, S.; Pardon, E.; Steyaert, J.; Ramboarina, S.; Chêne, A. Llama immunization with full-length var2csa generates cross-reactive and inhibitory single-domain antibodies against the dbl1x domain. Sci. Rep. 2014, 4, 7373. [Google Scholar] [CrossRef]
  70. De Genst, E.J.; Guilliams, T.; Wellens, J.; O’Day, E.M.; Waudby, C.A.; Meehan, S.; Dumoulin, M.; Hsu, S.-T.D.; Cremades, N.; Verschueren, K.H. Structure and properties of a complex of α-synuclein and a single-domain camelid antibody. J. Mol. Biol. 2010, 402, 326–343. [Google Scholar] [CrossRef]
  71. Braun, M.B.; Traenkle, B.; Koch, P.A.; Emele, F.; Weiss, F.; Poetz, O.; Stehle, T.; Rothbauer, U. Peptides in headlock–a novel high-affinity and versatile peptide-binding nanobody for proteomics and microscopy. Sci. Rep. 2016, 6, 19211. [Google Scholar] [CrossRef] [Green Version]
  72. Virant, D.; Traenkle, B.; Maier, J.; Kaiser, P.D.; Bodenhofer, M.; Schmees, C.; Vojnovic, I.; Pisak-Lukats, B.; Endesfelder, U.; Rothbauer, U. A peptide tag-specific nanobody enables high-quality labeling for dstorm imaging. Nat. Commun. 2018, 9, 930. [Google Scholar] [CrossRef]
  73. Götzke, H.; Kilisch, M.; Martínez-Carranza, M.; Sograte-Idrissi, S.; Rajavel, A.; Schlichthaerle, T.; Engels, N.; Jungmann, R.; Stenmark, P.; Opazo, F. The alfa-tag is a highly versatile tool for nanobody-based bioscience applications. Nat. Commun. 2019, 10, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Hulsik, D.L.; Liu, Y.-Y.; Strokappe, N.M.; Battella, S.; El Khattabi, M.; McCoy, L.E.; Sabin, C.; Hinz, A.; Hock, M.; Macheboeuf, P. A gp41 mper-specific llama vhh requires a hydrophobic cdr3 for neutralization but not for antigen recognition. PLoS Pathog. 2013, 9, e1003202. [Google Scholar]
  75. Boersma, S.; Khuperkar, D.; Verhagen, B.M.; Sonneveld, S.; Grimm, J.B.; Lavis, L.D.; Tanenbaum, M.E. Multi-color single-molecule imaging uncovers extensive heterogeneity in mrna decoding. Cell 2019, 178, 458–472.e19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Traenkle, B.; Segan, S.; Fagbadebo, F.O.; Kaiser, P.D.; Rothbauer, U. A novel epitope tagging system to visualize and monitor antigens in live cells with chromobodies. Sci. Rep. 2020, 10, 1–13. [Google Scholar] [CrossRef] [PubMed]
  77. Mendez, M.G.; Kojima, S.; Goldman, R.D. Vimentin induces changes in cell shape, motility, and adhesion during the epithelial to mesenchymal transition. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2010, 24, 1838–1851. [Google Scholar] [CrossRef] [Green Version]
  78. Belin, B.J.; Goins, L.M.; Mullins, R.D. Comparative analysis of tools for live cell imaging of actin network architecture. Bioarchitecture 2014, 4, 189–202. [Google Scholar] [CrossRef]
  79. Lemieux, M.G.; Janzen, D.; Hwang, R.; Roldan, J.; Jarchum, I.; Knecht, D.A. Visualization of the actin cytoskeleton: Different f-actin-binding probes tell different stories. Cytoskeleton 2014, 71, 157–169. [Google Scholar] [CrossRef]
  80. Schmidthals, K.; Helma, J.; Zolghadr, K.; Rothbauer, U.; Leonhardt, H. Novel antibody derivatives for proteome and high-content analysis. Anal. Bioanal. Chem. 2010, 397, 3203–3208. [Google Scholar] [CrossRef] [Green Version]
  81. Panza, P.; Maier, J.; Schmees, C.; Rothbauer, U.; Sollner, C. Live imaging of endogenous protein dynamics in zebrafish using chromobodies. Development 2015, 142, 1879–1884. [Google Scholar] [CrossRef] [Green Version]
  82. Maier, J.; Traenkle, B.; Rothbauer, U. Real-time analysis of epithelial-mesenchymal transition using fluorescent single-domain antibodies. Sci. Rep. 2015, 5, 13402. [Google Scholar] [CrossRef] [Green Version]
  83. Zolghadr, K.; Gregor, J.; Leonhardt, H.; Rothbauer, U. Case study on live cell apoptosis-assay using lamin-chromobody cell-lines for high-content analysis. Methods Mol. Biol. 2012, 911, 569–575. [Google Scholar]
  84. Plessner, M.; Melak, M.; Chinchilla, P.; Baarlink, C.; Grosse, R. Nuclear f-actin formation and reorganization upon cell spreading. J. Biol. Chem. 2015, 290, 11209–11216. [Google Scholar] [CrossRef] [Green Version]
  85. Maier, J.; Traenkle, B.; Rothbauer, U. Visualizing epithelial-mesenchymal transition using the chromobody technology. Cancer Res. 2016, 76, 5592–5596. [Google Scholar] [CrossRef] [Green Version]
  86. Burgess, A.; Lorca, T.; Castro, A. Quantitative live imaging of endogenous DNA replication in mammalian cells. PLoS ONE 2012, 7, e45726. [Google Scholar] [CrossRef] [Green Version]
  87. Schorpp, K.; Rothenaigner, I.; Maier, J.; Traenkle, B.; Rothbauer, U.; Hadian, K. A multiplexed high-content screening approach using the chromobody technology to identify cell cycle modulators in living cells. J. Biomol. Screen. 2016, 21, 965–977. [Google Scholar] [CrossRef] [Green Version]
  88. Buchfellner, A.; Yurlova, L.; Nuske, S.; Scholz, A.M.; Bogner, J.; Ruf, B.; Zolghadr, K.; Drexler, S.E.; Drexler, G.A.; Girst, S.; et al. A new nanobody-based biosensor to study endogenous parp1 in vitro and in live human cells. PLoS ONE 2016, 11, e0151041. [Google Scholar] [CrossRef] [Green Version]
  89. Rajan, M.; Mortusewicz, O.; Rothbauer, U.; Hastert, F.D.; Schmidthals, K.; Rapp, A.; Leonhardt, H.; Cardoso, M.C. Generation of an alpaca-derived nanobody recognizing gamma-h2ax. FEBS Open Bio 2015, 5, 779–788. [Google Scholar] [CrossRef] [Green Version]
  90. Jullien, D.; Vignard, J.; Fedor, Y.; Bery, N.; Olichon, A.; Crozatier, M.; Erard, M.; Cassard, H.; Ducommun, B.; Salles, B.; et al. Chromatibody, a novel non-invasive molecular tool to explore and manipulate chromatin in living cells. J. Cell Sci. 2016, 129, 2673–2683. [Google Scholar] [CrossRef] [Green Version]
  91. Traenkle, B.; Emele, F.; Anton, R.; Poetz, O.; Haeussler, R.S.; Maier, J.; Kaiser, P.D.; Scholz, A.M.; Nueske, S.; Buchfellner, A.; et al. Monitoring interactions and dynamics of endogenous beta-catenin with intracellular nanobodies in living cells. Mol. Cell. Proteom. MCP 2015, 14, 707–723. [Google Scholar] [CrossRef] [Green Version]
  92. Dietrich, J.; Sommersdorf, C.; Gohlke, S.; Poetz, O.; Traenkle, B.; Rothbauer, U.; Hessel-Pras, S.; Lampen, A.; Braeuning, A. Okadaic acid activates wnt/beta-catenin-signaling in human heparg cells. Arch. Toxicol. 2019, 93, 1927–1939. [Google Scholar] [CrossRef]
  93. Keller, B.-M.; Maier, J.; Secker, K.-A.; Egetemaier, S.-M.; Parfyonova, Y.; Rothbauer, U.; Traenkle, B. Chromobodies to quantify changes of endogenous protein concentration in living cells. Mol. Cell. Proteom. 2018, 17, 2518–2533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Tang, J.C.; Drokhlyansky, E.; Etemad, B.; Rudolph, S.; Guo, B.; Wang, S.; Ellis, E.G.; Li, J.Z.; Cepko, C.L. Detection and manipulation of live antigen-expressing cells using conditionally stable nanobodies. eLife 2016, 5, e15312. [Google Scholar] [CrossRef] [PubMed]
  95. Roebroek, T.; Duwé, S.; Vandenberg, W.; Dedecker, P. Reduced fluorescent protein switching fatigue by binding-induced emissive state stabilization. Int. J. Mol. Sci. 2017, 18, 2015. [Google Scholar] [CrossRef] [Green Version]
  96. Irannejad, R.; Tomshine, J.C.; Tomshine, J.R.; Chevalier, M.; Mahoney, J.P.; Steyaert, J.; Rasmussen, S.G.; Sunahara, R.K.; El-Samad, H.; Huang, B.; et al. Conformational biosensors reveal gpcr signalling from endosomes. Nature 2013, 495, 534–538. [Google Scholar] [CrossRef] [Green Version]
  97. Jakobs, B.D.; Spannagel, L.; Purvanov, V.; Uetz-von Allmen, E.; Matti, C.; Legler, D.F. Engineering of nanobodies recognizing the human chemokine receptor ccr7. Int. J. Mol. Sci. 2019, 20, 2597. [Google Scholar] [CrossRef] [Green Version]
  98. Stoeber, M.; Jullié, D.; Lobingier, B.T.; Laeremans, T.; Steyaert, J.; Schiller, P.W.; Manglik, A.; von Zastrow, M. A genetically encoded biosensor reveals location bias of opioid drug action. Neuron 2018, 98, 963–976.e5. [Google Scholar] [CrossRef] [Green Version]
  99. Bery, N.; Keller, L.; Soulié, M.; Gence, R.; Iscache, A.-L.; Cherier, J.; Cabantous, S.; Sordet, O.; Lajoie-Mazenc, I.; Pedelacq, J.-D. A targeted protein degradation cell-based screening for nanobodies selective toward the cellular rhob gtp-bound conformation. Cell Chem. Biol. 2019, 26, 1544–1558.e1546. [Google Scholar] [CrossRef]
  100. Cao, J.; Zhong, N.; Wang, G.; Wang, M.; Zhang, B.; Fu, B.; Wang, Y.; Zhang, T.; Zhang, Y.; Yang, K. Nanobody-based sandwich reporter system for living cell sensing influenza a virus infection. Sci. Rep. 2019, 9, 1–8. [Google Scholar] [CrossRef]
  101. Böldicke, T. Single domain antibodies for the knockdown of cytosolic and nuclear proteins. Protein Sci. 2017, 26, 925–945. [Google Scholar] [CrossRef] [Green Version]
  102. Lodish, M.B. Kinase inhibitors: Adverse effects related to the endocrine system. J. Clin. Endocrinol. Metab. 2013, 98, 1333–1342. [Google Scholar] [CrossRef] [Green Version]
  103. Ferguson, F.M.; Gray, N.S. Kinase inhibitors: The road ahead. Nat. Rev. Drug Discov. 2018, 17, 353. [Google Scholar] [CrossRef] [PubMed]
  104. Steels, A.; Verhelle, A.; Zwaenepoel, O.; Gettemans, J. Intracellular displacement of p53 using transactivation domain (p53 tad) specific nanobodies. In MAbs; Taylor & Francis: London, UK, 2018; pp. 1045–1059. [Google Scholar] [CrossRef]
  105. Steels, A.; Vannevel, L.; Zwaenepoel, O.; Gettemans, J. Nb-induced stabilisation of p53 in hpv-infected cells. Sci. Rep. 2019, 9, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Gulati, S.; Jin, H.; Masuho, I.; Orban, T.; Cai, Y.; Pardon, E.; Martemyanov, K.A.; Kiser, P.D.; Stewart, P.L.; Ford, C.P. Targeting g protein-coupled receptor signaling at the g protein level with a selective nanobody inhibitor. Nat. Commun. 2018, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  107. Summanen, M.; Granqvist, N.; Tuominen, R.K.; Yliperttula, M.; Verrips, C.T.; Boonstra, J.; Blanchetot, C.; Ekokoski, E. Kinetics of pkcε activating and inhibiting llama single chain antibodies and their effect on pkcε translocation in hela cells. PLoS ONE 2012, 7, e35630. [Google Scholar] [CrossRef]
  108. Van Audenhove, I.; Boucherie, C.; Pieters, L.; Zwaenepoel, O.; Vanloo, B.; Martens, E.; Verbrugge, C.; Hassanzadeh-Ghassabeh, G.; Vandekerckhove, J.; Cornelissen, M. Stratifying fascin and cortactin function in invadopodium formation using inhibitory nanobodies and targeted subcellular delocalization. FASEB J. 2014, 28, 1805–1818. [Google Scholar] [CrossRef]
  109. Bertier, L.; Boucherie, C.; Zwaenepoel, O.; Vanloo, B.; Van Troys, M.; Van Audenhove, I.; Gettemans, J. Inhibitory cortactin nanobodies delineate the role of nta-and sh3-domain–specific functions during invadopodium formation and cancer cell invasion. FASEB J. 2017, 31, 2460–2476. [Google Scholar] [CrossRef] [Green Version]
  110. Bertier, L.; Hebbrecht, T.; Mettepenningen, E.; De Wit, N.; Zwaenepoel, O.; Verhelle, A.; Gettemans, J. Nanobodies targeting cortactin proline rich, helical and actin binding regions downregulate invadopodium formation and matrix degradation in scc-61 cancer cells. Biomed. Pharmacother. 2018, 102, 230–241. [Google Scholar] [CrossRef]
  111. Hebbrecht, T.; Van Audenhove, I.; Zwaenepoel, O.; Verhelle, A.; Gettemans, J. Vca nanobodies target n-wasp to reduce invadopodium formation and functioning. PLoS ONE 2017, 12, e0185076. [Google Scholar] [CrossRef] [Green Version]
  112. Van Impe, K.; Bethuyne, J.; Cool, S.; Impens, F.; Ruano-Gallego, D.; De Wever, O.; Vanloo, B.; Van Troys, M.; Lambein, K.; Boucherie, C. A nanobody targeting the f-actin capping protein capg restrains breast cancer metastasis. Breast Cancer Res. 2013, 15, R116. [Google Scholar] [CrossRef] [Green Version]
  113. Singh, S.; Murillo, G.; Chen, D.; Parihar, A.S.; Mehta, R.G. Suppression of breast cancer cell proliferation by selective single-domain antibody for intracellular stat3. Breast Cancer Basic Clin. Res. 2018, 12, 1178223417750858. [Google Scholar] [CrossRef]
  114. Schmidt, F.I.; Lu, A.; Chen, J.W.; Ruan, J.; Tang, C.; Wu, H.; Ploegh, H.L. A single domain antibody fragment that recognizes the adaptor asc defines the role of asc domains in inflammasome assembly. J. Exp. Med. 2016, 213, 771–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. De Clercq, S.; Zwaenepoel, O.; Martens, E.; Vandekerckhove, J.; Guillabert, A.; Gettemans, J. Nanobody-induced perturbation of lfa-1/l-plastin phosphorylation impairs mtoc docking, immune synapse formation and t cell activation. Cell. Mol. Life Sci. 2013, 70, 909–922. [Google Scholar] [CrossRef] [PubMed]
  116. De Clercq, S.; Boucherie, C.; Vandekerckhove, J.; Gettemans, J.; Guillabert, A. L-plastin nanobodies perturb matrix degradation, podosome formation, stability and lifetime in thp-1 macrophages. PLoS ONE 2013, 8, e78108. [Google Scholar] [CrossRef] [Green Version]
  117. Delanote, V.; Vanloo, B.; Catillon, M.; Friederich, E.; Vandekerckhove, J.; Gettemans, J. An alpaca single-domain antibody blocks filopodia formation by obstructing l-plastin-mediated f-actin bundling. FASEB J. 2010, 24, 105–118. [Google Scholar] [CrossRef]
  118. Van den Abbeele, A.; De Clercq, S.; De Ganck, A.; De Corte, V.; Van Loo, B.; Soror, S.H.; Srinivasan, V.; Steyaert, J.; Vandekerckhove, J.; Gettemans, J. A llama-derived gelsolin single-domain antibody blocks gelsolin–g-actin interaction. Cell. Mol. Life Sci. 2010, 67, 1519–1535. [Google Scholar] [CrossRef]
  119. Messer, A.; Butler, D.C. Optimizing intracellular antibodies (intrabodies/nanobodies) to treat neurodegenerative disorders. Neurobiol. Dis. 2020, 134, 104619. [Google Scholar] [CrossRef]
  120. Dong, J.-X.; Lee, Y.; Kirmiz, M.; Palacio, S.; Dumitras, C.; Moreno, C.M.; Sando, R.; Santana, L.F.; Südhof, T.C.; Gong, B. A toolbox of nanobodies developed and validated for use as intrabodies and nanoscale immunolabels in mammalian brain neurons. eLife 2019, 8, e48750. [Google Scholar] [CrossRef]
  121. Mahajan, S.P.; Meksiriporn, B.; Waraho-Zhmayev, D.; Weyant, K.B.; Kocer, I.; Butler, D.C.; Messer, A.; Escobedo, F.A.; DeLisa, M.P. Computational affinity maturation of camelid single-domain intrabodies against the nonamyloid component of alpha-synuclein. Sci. Rep. 2018, 8, 1–14. [Google Scholar] [CrossRef]
  122. Guilliams, T.; El-Turk, F.; Buell, A.K.; O’Day, E.M.; Aprile, F.A.; Esbjörner, E.K.; Vendruscolo, M.; Cremades, N.; Pardon, E.; Wyns, L. Nanobodies raised against monomeric α-synuclein distinguish between fibrils at different maturation stages. J. Mol. Biol. 2013, 425, 2397–2411. [Google Scholar] [CrossRef] [Green Version]
  123. El Turk, F.; De Genst, E.; Guilliams, T.; Fauvet, B.; Hejjaoui, M.; Di Trani, J.; Chiki, A.; Mittermaier, A.; Vendruscolo, M.; Lashuel, H.A. Exploring the role of post-translational modifications in regulating α-synuclein interactions by studying the effects of phosphorylation on nanobody binding. Protein Sci. 2018, 27, 1262–1274. [Google Scholar] [CrossRef] [Green Version]
  124. Butler, D.C.; Joshi, S.N.; Genst, E.D.; Baghel, A.S.; Dobson, C.M.; Messer, A. Bifunctional anti-non-amyloid component α-synuclein nanobodies are protective in situ. PLoS ONE 2016, 11, e0165964. [Google Scholar] [CrossRef] [Green Version]
  125. Gueorguieva, D.; Li, S.; Walsh, N.; Mukerji, A.; Tanha, J.; Pandey, S.; Gueorguieva, D.; Li, S.; Walsh, N.; Mukerji, A. Identification of single-domain, bax-specific intrabodies that confer resistance to mammalian cells against oxidative-stress-induced apoptosis. FASEB J. 2006, 20, 2636–2638. [Google Scholar] [CrossRef] [Green Version]
  126. Morgenstern, T.J.; Park, J.; Fan, Q.R.; Colecraft, H.M. A potent voltage-gated calcium channel inhibitor engineered from a nanobody targeted to auxiliary cavβ subunits. eLife 2019, 8, e49253. [Google Scholar] [CrossRef]
  127. Schenck, S.; Kunz, L.; Sahlender, D.; Pardon, E.; Geertsma, E.R.; Savtchouk, I.; Suzuki, T.; Neldner, Y.; Štefanić, S.A.; Steyaert, J. Generation and characterization of anti-vglut nanobodies acting as inhibitors of transport. Biochemistry 2017, 56, 3962–3971. [Google Scholar] [CrossRef]
  128. Dieleman, J.L.; Haakenstad, A.; Micah, A.; Moses, M.; Abbafati, C.; Acharya, P.; Adhikari, T.B.; Adou, A.K.; Kiadaliri, A.A.; Alam, K. Spending on health and hiv/aids: Domestic health spending and development assistance in 188 countries, 1995–2015. Lancet 2018, 391, 1799–1829. [Google Scholar] [CrossRef] [Green Version]
  129. Vercruysse, T.; Pardon, E.; Vanstreels, E.; Steyaert, J.; Daelemans, D. An intrabody based on a llama single-domain antibody targeting the n-terminal α-helical multimerization domain of hiv-1 rev prevents viral production. J. Biol. Chem. 2010, 285, 21768–21780. [Google Scholar] [CrossRef] [Green Version]
  130. Boons, E.; Li, G.; Vanstreels, E.; Vercruysse, T.; Pannecouque, C.; Vandamme, A.-M.; Daelemans, D. A stably expressed llama single-domain intrabody targeting rev displays broad-spectrum anti-hiv activity. Antivir. Res. 2014, 112, 91–102. [Google Scholar] [CrossRef]
  131. Matz, J.; Hérate, C.; Bouchet, J.; Dusetti, N.; Gayet, O.; Baty, D.; Benichou, S.; Chames, P. Selection of intracellular single-domain antibodies targeting the hiv-1 vpr protein by cytoplasmic yeast two-hybrid system. PLoS ONE 2014, 9, e113729. [Google Scholar] [CrossRef] [Green Version]
  132. Bouchet, J.; Basmaciogullari, S.E.; Chrobak, P.; Stolp, B.; Bouchard, N.; Fackler, O.T.; Chames, P.; Jolicoeur, P.; Benichou, S.; Baty, D. Inhibition of the nef regulatory protein of hiv-1 by a single-domain antibody. Blood 2011, 117, 3559–3568. [Google Scholar] [CrossRef] [Green Version]
  133. Bouchet, J.; Hérate, C.; Guenzel, C.A.; Vérollet, C.; Järviluoma, A.; Mazzolini, J.; Rafie, S.; Chames, P.; Baty, D.; Saksela, K. Single-domain antibody-sh3 fusions for efficient neutralization of hiv-1 nef functions. J. Virol. 2012, 86, 4856–4867. [Google Scholar] [CrossRef] [Green Version]
  134. Ashour, J.; Schmidt, F.I.; Hanke, L.; Cragnolini, J.; Cavallari, M.; Altenburg, A.; Brewer, R.; Ingram, J.; Shoemaker, C.; Ploegh, H.L. Intracellular expression of camelid single-domain antibodies specific for influenza virus nucleoprotein uncovers distinct features of its nuclear localization. J. Virol. 2015, 89, 2792–2800. [Google Scholar] [CrossRef] [Green Version]
  135. Hanke, L.; Knockenhauer, K.E.; Brewer, R.C.; van Diest, E.; Schmidt, F.I.; Schwartz, T.U.; Ploegh, H.L. The antiviral mechanism of an influenza a virus nucleoprotein-specific single-domain antibody fragment. MBio 2016, 7, e01569-16. [Google Scholar] [CrossRef] [Green Version]
  136. Glab-Ampai, K.; Malik, A.A.; Chulanetra, M.; Thanongsaksrikul, J.; Thueng-In, K.; Srimanote, P.; Tongtawe, P.; Chaicumpa, W. Inhibition of hcv replication by humanized-single domain transbodies to ns4b. Biochem. Biophys. Res. Commun. 2016, 476, 654–664. [Google Scholar] [CrossRef]
  137. Jittavisutthikul, S.; Thanongsaksrikul, J.; Thueng-In, K.; Chulanetra, M.; Srimanote, P.; Seesuay, W.; Malik, A.A.; Chaicumpa, W. Humanized-vhh transbodies that inhibit hcv protease and replication. Viruses 2015, 7, 2030–2056. [Google Scholar] [CrossRef] [Green Version]
  138. Thueng-In, K.; Thanongsaksrikul, J.; Srimanote, P.; Bangphoomi, K.; Poungpair, O.; Maneewatch, S.; Choowongkomon, K.; Chaicumpa, W. Cell penetrable humanized-vh/vhh that inhibit rna dependent rna polymerase (ns5b) of hcv. PLoS ONE 2012, 7, e49254. [Google Scholar] [CrossRef] [Green Version]
  139. Serruys, B.; Van Houtte, F.; Farhoudi-Moghadam, A.; Leroux-Roels, G.; Vanlandschoot, P. Production, characterization and in vitro testing of hbcag-specific vhh intrabodies. J. Gen. Virol. 2010, 91, 643–652. [Google Scholar] [CrossRef]
  140. Alzogaray, V.; Danquah, W.; Aguirre, A.; Urrutia, M.; Berguer, P.; Véscovi, E.G.; Haag, F.; Koch-Nolte, F.; Goldbaum, F.A. Single-domain llama antibodies as specific intracellular inhibitors of spvb, the actin adp-ribosylating toxin of salmonella typhimurium. FASEB J. 2011, 25, 526–534. [Google Scholar] [CrossRef]
  141. Tremblay, J.M.; Kuo, C.-L.; Abeijon, C.; Sepulveda, J.; Oyler, G.; Hu, X.; Jin, M.M.; Shoemaker, C.B. Camelid single domain antibodies (vhhs) as neuronal cell intrabody binding agents and inhibitors of clostridium botulinum neurotoxin (bont) proteases. Toxicon 2010, 56, 990–998. [Google Scholar] [CrossRef] [Green Version]
  142. Yu, D.; Lee, H.; Hong, J.; Jung, H.; Jo, Y.; Oh, B.-H.; Park, B.O.; Do Heo, W. Optogenetic activation of intracellular antibodies for direct modulation of endogenous proteins. Nat. Methods 2019, 16, 1095–1100. [Google Scholar] [CrossRef]
  143. Gil, A.A.; Carrasco-López, C.; Zhu, L.; Zhao, E.M.; Ravindran, P.T.; Wilson, M.Z.; Goglia, A.G.; Avalos, J.L.; Toettcher, J.E. Optogenetic control of protein binding using light-switchable nanobodies. Nat. Commun. 2020, 11, 1–12. [Google Scholar] [CrossRef] [PubMed]
  144. Farrants, H.; Tarnawski, M.; Müller, T.G.; Otsuka, S.; Hiblot, J.; Koch, B.; Kueblbeck, M.; Kräusslich, H.-G.; Ellenberg, J.; Johnsson, K. Chemogenetic control of nanobodies. Nat. Methods 2020, 17, 279–282. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Intrabodies for visualization. (a) Schematic depiction of nano- and chromobodies derived from heavy-chain antibodies of Camelidae (b) Illustration of different labelling strategies for visualization of the protein of interest (POI) using intrabodies coupled to a detectable moiety and summary of their characteristics: POI genetically coupled to GFP recognized by a GFP intrabody, POI genetically coupled to a short peptide-tag recognized by the tag-specific intrabody and POI directly recognized by a specific intrabody. Model protein structures adapted from PDBs 3OGO, 5H88, 1EMA and 5LEW. Representative images of living Hela cells transiently expressing PNCAGFP and the GFP-specific chromobody (tagRFP), PNCAPEP-tag and the PEP-tag-specific chromobody (tagRFP) or PCNA-specific chromobody (tagRFP) as examples for the described labeling strategies are shown. Scale bar: 25 µm.
Figure 1. Intrabodies for visualization. (a) Schematic depiction of nano- and chromobodies derived from heavy-chain antibodies of Camelidae (b) Illustration of different labelling strategies for visualization of the protein of interest (POI) using intrabodies coupled to a detectable moiety and summary of their characteristics: POI genetically coupled to GFP recognized by a GFP intrabody, POI genetically coupled to a short peptide-tag recognized by the tag-specific intrabody and POI directly recognized by a specific intrabody. Model protein structures adapted from PDBs 3OGO, 5H88, 1EMA and 5LEW. Representative images of living Hela cells transiently expressing PNCAGFP and the GFP-specific chromobody (tagRFP), PNCAPEP-tag and the PEP-tag-specific chromobody (tagRFP) or PCNA-specific chromobody (tagRFP) as examples for the described labeling strategies are shown. Scale bar: 25 µm.
Biomolecules 10 01701 g001
Figure 2. Intrabodies as intracellular modulators. Graphical representation how intrabodies can address their POI (protein of interest) for functional modulation. Shown are models of blocking and activating intrabodies, intrabodies inducing degradation and switchable intrabody systems. Additionally, commonly applied testing systems and potential clinical applications of intrabodies are listed. Model protein structures adapted from PDB 3OGO, 1EMA, 5LEW, 2EPE and parts of schematic art pieces are freely available from Servier Medical Art (https://smart.servier.com; Servier Medical Art by Servier is licensed under a Creative Commons Attribution 3.0 Unported License).
Figure 2. Intrabodies as intracellular modulators. Graphical representation how intrabodies can address their POI (protein of interest) for functional modulation. Shown are models of blocking and activating intrabodies, intrabodies inducing degradation and switchable intrabody systems. Additionally, commonly applied testing systems and potential clinical applications of intrabodies are listed. Model protein structures adapted from PDB 3OGO, 1EMA, 5LEW, 2EPE and parts of schematic art pieces are freely available from Servier Medical Art (https://smart.servier.com; Servier Medical Art by Servier is licensed under a Creative Commons Attribution 3.0 Unported License).
Biomolecules 10 01701 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wagner, T.R.; Rothbauer, U. Nanobodies Right in the Middle: Intrabodies as Toolbox to Visualize and Modulate Antigens in the Living Cell. Biomolecules 2020, 10, 1701. https://doi.org/10.3390/biom10121701

AMA Style

Wagner TR, Rothbauer U. Nanobodies Right in the Middle: Intrabodies as Toolbox to Visualize and Modulate Antigens in the Living Cell. Biomolecules. 2020; 10(12):1701. https://doi.org/10.3390/biom10121701

Chicago/Turabian Style

Wagner, Teresa R., and Ulrich Rothbauer. 2020. "Nanobodies Right in the Middle: Intrabodies as Toolbox to Visualize and Modulate Antigens in the Living Cell" Biomolecules 10, no. 12: 1701. https://doi.org/10.3390/biom10121701

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