Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 DOI 10.1002/biot.201300498 www.biotechnology-journal.com Review RNA aptamers as genetic control devices: The potential of riboswitches as synthetic elements for regulating gene expression Christian Berens1, Florian Groher2 and Beatrix Suess2 1 Institute of Molecular Pathogenesis, Friedrich-Loeffler-Institut, Federal Research Institute for Animal Health, Jena, Germany Biology and Genetics, Department of Biology, Technical University Darmstadt, Germany 2 Synthetic RNA utilizes many different mechanisms to control gene expression. Among the regulatory elements that respond to external stimuli, riboswitches are a prominent and elegant example. They consist solely of RNA and couple binding of a small molecule ligand to the so-called “aptamer domain” with a conformational change in the downstream “expression platform” which then determines system output. The modular organization of riboswitches and the relative ease with which ligand-binding RNA aptamers can be selected in vitro against almost any molecule have led to the rapid and widespread adoption of engineered riboswitches as artificial genetic control devices in biotechnology and synthetic biology over the past decade. This review highlights proofof-principle applications to demonstrate the versatility and robustness of engineered riboswitches in regulating gene expression in pro- and eukaryotes. It then focuses on strategies and parameters to identify aptamers that can be integrated into synthetic riboswitches that are functional in vivo, before finishing with a reflection on how to improve the regulatory properties of engineered riboswitches, so that we can not only further expand riboswitch applicability, but also finally fully exploit their potential as control elements in regulating gene expression. Received 14 OCT 2014 Revised 23 DEC 2014 Accepted 15 JAN 2015 Keywords: Engineered riboswitch · Regulatory circuits · RNA aptamer · RNA aptazyme · SELEX 1 Introduction Working with RNA can be like opening a “goodie bag” – for one, you never know what you might find until you look, but once in a while you will discover something exciting and unknown you never had thought about before. Similarly, our knowledge of what RNA can do in a Correspondence: Prof. Dr. Beatrix Suess, Synthetic Biology and Genetics, Department of Biology, Technical University Darmstadt, Schnittspahnstr. 10, 64287 Darmstadt, Germany. E-mail: bsuess@bio.tu-darmstadt.de Abbreviations: CRISPR, clustered regularly interspaced short palindromic repeats; GFP, green fluorescent protein; miRNA, micro RNA; mRNA, messenger RNA; PPDA, pyrimido[4,5-d]pyrimidine-2,4-diamine; SD, Shine-Dalgarno; SELEX, systematic evolution of ligands by exponential enrichment; shRNA, short hairpin RNA; siRNA, short interfering RNA; sRNA, small RNA; UTR, untranslated leader region 246 cell has changed dramatically since the “central dogma of molecular biology”, in which RNA served primarily as an information carrier, was proposed about 45 years ago [1]. And fortunately, our perception of what RNA might be capable of doing in a cell or in a test-tube is still expanding, since new regulatory functions of RNA are continuously being discovered [2]. Circular exonic RNA molecules, for example, are an abundant and differentially expressed RNA species that have recently been proposed to serve as molecular “sinks” for trans-acting short RNA regulators, as mRNA traps in regulating protein expression, or as interaction partners of RNA-binding proteins [3]. Another newly identified and apparently ubiquitous RNA species is long noncoding RNAs. This quite heterogeneous population of RNA molecules participates in many different molecular and cellular processes involved in regulating nuclear organization, development, viability, immunity, and disease [4]. CRISPR/Cas systems are used by various bacteria and archaea as an adaptive immune © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotecvisions.com www.biotechnology-journal.com system to mediate defence against bacteriophages and other foreign nucleic acids. They use a “guide RNA” with complementarity to a specific DNA target site to direct a Cas9 nuclease which then initiates a double-strand break resulting in degradation of the target DNA molecule [5]. Many of these types of RNA-mediated regulation of gene expression have proven to be very attractive for biotechnological and even therapeutic applications. RNA interference [6], using miRNA, shRNA and siRNA, is employed routinely in the lab for loss-of-function studies in eukaryotes [7] and also shows potential as therapeutic molecule in disease treatment [8]. Circular RNAs have biotechnological potential due to their exceptional intracellular stability and their ability to titrate endogenous RNA species or to support internal ribosome entry sitemediated translation [3]. Finally, CRISPR/Cas9 systems are very successful at precisely manipulating the genomes of both pro- and eukaryotes [9, 10]. Sophisticated molecular systems that sense external stimuli and respond to these by modulating the output of a cellular process are another important application in biotechnology. Genetically, this problem can be tackled by several different approaches of which one is based on allosteric RNA elements termed riboswitches. Typically, riboswitches are genetic control elements that consist of two domains – an aptamer domain that binds a small molecule ligand and an expression platform that converts ligand binding into a change in gene expression by adopting an alternative RNA structure [11]. Interestingly, such allosteric aptamer-containing devices were designed and shown to be active in vitro [12, 13] and in vivo [14] about five years before the first naturally occurring riboswitches were discovered [15–17]. Since then, a plethora of different riboswitches have been identified in all three domains of life and have been reviewed comprehensively [18, 19]; see also the recent special edition on “Riboswitches” in Biochemica et Biophysica Acta – Gene Regulatory Mechanisms (2014) [133]. There are several reasons why riboswitches are an attractive tool in biotechnology. Most importantly, they consist solely of RNA. So far, protein components have not been found to be necessary for their activity. Therefore, they are easy to implement, because they involve the transfer of only a single genetic control element into an organism. They are modular, so that different aptamer domains can be combined with different expression platforms [20]. Finally, their use allows spatial, temporal and dosage control over target gene expression. It is therefore not astonishing that engineered riboswitches were introduced early to the fields of genetic engineering and synthetic biology [14, 21] and have seen widespread application since then [22–26]. This review is not intended to serve as a comprehensive overview of all aspects of riboswitch engineering. For example, we do not cover the wide fields of diagnostic and therapeutic aptamers and riboswitches, which have © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim been addressed elsewhere [23, 27–30]; see also the September 2014 special issue on Aptamers published in Molecular Therapy – Nucleic Acids. Our aim is rather to present three topics we consider to be of prime importance for riboswitch biotechnology. We not only demonstrate how far the field has progressed since the early days of artificial in vitro “riboregulators” [12], but also present strategies we believe are important for isolating in vivo active riboswitches, as well as challenges that will be faced in designing more versatile and effective riboswitches. We hope this format will stimulate discussions that will lead to new and improved approaches and selection systems and, ultimately and hopefully, to the development of new and more efficient synthetic riboswitches. 2 Engineered riboswitches: An overview of the status quo 2.1 The toolbox of engineered riboswitches recapitulates the functionality of natural riboswitches Natural riboswitches control translation, transcription, ribozyme activity, and splicing [15–17, 31, 32]. In the majority of these cases, they act in cis as part of the mRNA transcript. But they can also regulate in trans either as a processed, terminated riboswitch that functions like an sRNA [33] or as a full-length transcript containing a binding site for sequestering an activated RNA-binding protein [34, 35] or by acting as an antisense RNA [36]. In synthetic biology, creating complex and more realistic genetic circuits requires that we eventually identify and design “BioBricks” that not only reflect the entire spectrum of genetic regulatory elements, but that can also be tailored and fine-tuned to obtain the desired output level of the target gene. The first step towards achieving this goal is the construction of proof-of-principle applications to demonstrate that we are able to reproduce these genetic elements using artificial and/or heterologous components. The following section will provide an overview of such archetype constructs. More detailed descriptions can be found in several other reviews [25, 26]. In bacteria, engineered riboswitches most often target translation initiation, either by controlling access to the ribosomal binding site through helix slippage [37] or by sequestering this Shine-Dalgarno (SD) element (Fig. 1A) [38]. Regulation of gene expression by a theophylline-controlled translational switch was efficient enough to allow the control of chemotaxis in Escherichia coli [39]. In contrast, transcriptionally-controlled synthetic riboswitches have only recently been introduced to the toolbox by the Mörl [40], Batey [41, 42], and Micklefield [43] groups. Here, a transcriptional terminator is formed depending on the binding state of the aptamer (Fig. 1B). It will be interest- 247 Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com www.biotecvisions.com Figure 1. Common mechanisms of engineered riboswitches in bacteria. (A) Regulation of translation initiation: In the absence of a ligand, a stem-loop structure is formed between the aptamer domain (depicted in green) and a sequence element complementary to the Shine-Dalgarno (SD) sequence (magenta). Thus, the SD sequence (orange) is accessible for 30S binding and translation initiation occurs. As a consequence of ligand binding (blue circle) and folding of the aptamer domain, an alternative stem-loop is formed which sequesters the SD sequence, blocking the binding of the 30S ribosomal subunit. (B) Regulation of transcription termination: The aptamer domain is fused to a short spacer region (grey), followed by a sequence complementary to the 3’ part of the aptamer (magenta) and a U stretch. In the absence of ligand, the complementary 3’ part is base-paired with the aptamer forming a terminator structure; thus, RNA polymerase (RNAP) dissociates and transcription is terminated. Upon ligand binding, terminator structure formation is inhibited and transcription can proceed, resulting in expression of the reporter gene. (C) Control of gene expression with catalytic riboswitches: A ligand-dependent aptazyme is inserted into the 5’-UTR of an mRNA in such a way that ligand-induced self-cleavage liberates a sequestered SD sequence to induce translation initiation. ing to see how these devices perform with respect to regulatory efficiency, basal activity, noise, and kinetic parameters, in comparison to analogous, but translationally controlled riboswitches in similar experimental settings. The cleavage activity of small ribozymes, like the hammerhead ribozyme, was made ligand-dependent by inserting an aptamer domain into one of its three stemloops [44]. Luckily, all three stem-loops of the hammerhead ribozyme are amenable to aptamer insertion [45–48], greatly facilitating the applicability of such “aptazymes”, because the “best fit” between ribozyme and target sequence can now be exploited. In bacteria, such aptazymes were used to liberate the ribosomal binding site after ligand-dependent ribozyme cleavage (Fig. 1C). With respect to the thermally-controlled riboswitches mentioned below in Section 2.3 [49], it will be interesting to see if the recently introduced “thermozymes” [50] can also be modified to include an additional aptamer control element. The examples listed here indicate that in bacteria engineered riboswitches function in regulatory settings that are also found for natural riboswitches. In eukaryotes, translational regulation by engineered riboswitches functions differently. Here, insertion of the 248 riboswitch into the 5’-untranslated leader region leads to inhibition of translation initiation, either by preventing binding of the small ribosomal subunit to the mRNA cap structure or by interfering with ribosomal subunit scanning for the AUG start codon (Fig. 2A). Here, aptamers which specifically recognize a Hoechst dye [14], malachite green [51], tetracycline [21, 52], neomycin [53], biotin or theophylline [54] (Fig. 2A) have been engineered to serve as regulatory elements. The riboswitches were active in yeast, Xenopus oocytes and mammalian cell culture [55]. Cap-independent translation initiation by an internal ribosome entry site has also been controlled by engineered riboswitches. However, up to now this has only been demonstrated in vitro [56]. The only naturally occurring eukaryotic riboswitches identified so far have been discovered in filamentous fungi, green algae and higher plants where they control gene expression via regulation of mRNA splicing [57–60]. Analogously, theophylline- or tetracycline-binding aptamers have been used to control constitutive and alternative splicing by placing them close to either the 5’-splice site, the 3’-splice site or the intron internal branch point (Fig. 2B) [61–63]. © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com www.biotecvisions.com Figure 2. Common mechanisms of engineered riboswitches in eukaryotes. (A) Regulation of ribosome scanning: After insertion of the aptamer (depicted in green) into the 5’UTR of a mRNA, ligand binding (blue circle) to the aptamer prevents scanning of the 40S ribosomal subunit and, consequently, translation initiation. (B) Regulation of mRNA splicing: An aptamer is integrated into an intron of a eukaryotic mRNA to control the accessibility of the 5’ splice site (SS). Ligand binding inhibits splicing. (C) Control of gene expression with catalytic riboswitches: A ligand-dependent aptazyme is inserted into the 5’-UTR of a eukaryotic mRNA. Ligand-induced selfcleavage triggers RNA degradation (magenta pacman). (D) Regulation of RNA interference: An integrated aptamer domain (green) interferes in its ligandbound state with the enzymatic activity of Dicer. Aptazyme-mediated control of gene expression in yeast and in mammalian cell lines using hammerhead ribozymes (Fig. 2C) has been demonstrated by several groups [46, 48, 64–66]. Here, ligand-dependent ribozyme cleavage leads to mRNA degradation. Besides the hammerhead ribozyme, the hepatitis delta virus ribozyme has also been successfully applied to RNA engineering [67]. RNA interference is an important and widely used mechanism in eukaryotes to repress gene expression by © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim base-pairing to mRNA molecules. RNA interference controls a broad range of developmental and physiological processes [68–70]. It is also a standard method in molecular biology with a possible therapeutic option which is being tested in clinical trials [71]. Consequently, attempts to control the biogenesis pathway of RNA interference were made by introducing theophylline-dependent aptamers [72, 73] or aptazymes [74] at sites of siRNA or shRNA molecules where their presence can interfere with 249 Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com processing by either Drosha or Dicer (Fig. 2D). So far, extremely high ligand concentrations are needed for efficient regulation. Optimization of the experimental strategies is clearly needed and it will be challenging to see if endogenous miRNA molecules can also become subject to riboswitch control. 2.2 Riboswitch regulation is becoming more robust and versatile While proof-of-principle applications are important for demonstrating the functionality of any kind of experimental approach, they do not, per se, demonstrate their general transferability to other systems, organisms or even across the kingdoms of life. Undoubtedly, the general applicability and the ease with which system functionality can be achieved is an important parameter for determining success and acceptance of an application as a tool by the respective community. So far, most of the regulatory strategies presented in the previous section were tested in the well-characterized model organisms E. coli for Gram-negative bacteria, Bacillus subtilis for Grampositive bacteria, Saccharomyces cerevisiae for lower eukaryotes and mammalian cell lines for higher eukaryotes [19]. In developmental biology, biotechnology and in medicine, other organisms are frequently employed as physiological, developmental and disease models or as cell factories for the production of materials [75–80]. Since endogenous riboswitches have been identified in the genomes of many different organisms [81], it seem reasonable to assume that engineered riboswitches, based on similar regulatory principles, should also be active in these species. In bacteria, the Gallivan lab, consequently, designed a set of theophylline-inducible translationally controlled riboswitches, initially termed A-E, which vary in the strength of their ribosomal binding sites and the stability of their secondary structures [38]. Later, riboswitches designated E* (supplemental material in [38]), F (clone 8.1 from [82]), and G (clone D2 from [83]) were added to the set. These riboswitches were then tested for activity in diverse bacterial species, including the developmental model organism and major producer of antibiotics, anticancer and antiviral drugs or enzymes, Streptomyces coelicolor [84], cyanobacteria, model organisms for photosynthesis and potential material-producing cell factories [85, 86] or tobacco chloroplasts, which are of interest in plant biotechnology [87]. The riboswitch set was also evaluated in important pathogens like Agrobacterium tumefaciens, Acinetobacter baumannii, Streptococcus pyogenes [38], Mycobacterium [88] and Francisella [89]. The tetracycline aptamer was shown to be functional in the anaerobic model organism Methanosarcina acetivorans, an archaeon [90]. In all organisms tested, at least one riboswitch was active enough to establish a conditional knockout of an endogenous gene. In some cases, the reg- 250 www.biotecvisions.com ulatory factors obtained were similar to those obtained with transcriptional regulation. Effector molecule penetration across several membranes was also effective enough to permit target gene expression control in chloroplasts and in intracellular infections models. Theophylline was the preferred ligand of choice. It can be added at high concentrations (2–4 mM) to the medium without causing obvious toxicity. In contrast, tetracycline and neomycin are antibiotics in bacteria and the concentrations needed for riboswitch activity can be close to or above the respective minimal inhibitory concentrations. Unfortunately, anhydrotetracycline, a less toxic derivative of tetracycline, does not bind to the tetracycline aptamer [91]. Translational ON-switches are the most frequently used type of expression platform. They appear the easiest to integrate into the sequence of the target gene. And finally, there is no single “optimal” riboswitch. Depending on the species and the specific target gene, different riboswitches displayed the best regulatory properties. So far, several constructs have to be tested to find the right combination for a given application. Regulation of gene expression by synthetic riboswitches is not limited to bacteria. Viral replication was inhibited in eukaryotic DNA (adenovirus, adeno-associated virus) and RNA viruses (measles) by flanking essential genes with theophylline-dependent aptazymes [92, 93]. Taken together, gene expression regulation by riboswitches has been demonstrated to be stringent, reversible, and dose-dependent in viruses, organelles, archaea, and bacteria from a broad taxonomic range of species, firmly establishing this technology as robust and universally applicable. 2.3 Riboswitches can be integrated into complex regulatory circuits Genetic circuits are frequently more complex than simple one input/one output type networks. They can have complex architectures like toggle switches [94], oscillators [95], Boolean logic gates that integrate several external signals [96], and binary or graded responses [97]. They can also include positive [98] or negative [99] feedback loops or linearizers [100]. These examples of engineered circuits were all realized with protein-based transcription factors. Consequently, the interesting question is to determine if RNA-based devices can be used similarly to assemble multiple input or complex regulatory circuits. Nature itself has come up with several combinatorial solutions to tackle diverse regulatory problems. In B. subtilis, the glmS riboswitch probably represents the simplest solution to regulation by multiple signals. The ligand-binding pocket of the aptamer domain not only binds intracellular glucosamine-6-phosphate as an activating agent. It also binds glucose-6-phosphate, which acts as an inhibitor of ribozyme activity. The riboswitch, thus, integrates metabolite information to sense the overall © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotecvisions.com www.biotechnology-journal.com metabolic state of the cell [101]. A second solitary element that senses and integrates two different input signals is the adenine riboswitch encoded by the add gene from Vibrio vulnificus. This translational regulatory element senses temperature and ligand concentration to confer efficient regulation over the range of physiologically relevant temperatures, allowing gene expression control to adapt to the differing temperatures experienced in the free-living and in the infectious environments [49]. In Bacillus clausii, two different riboswitches, which responded independently to S-adenosylmethionine and vitamin B12, respectively, were found in tandem in the 5’ untranslated leader region (UTR) of the metE mRNA. System output fits a Boolean NOR gate, because transcription is terminated in the presence of only one or of both ligands [102]. The Yokobayashi group was first to exploit this strategy in riboswitch engineering. They embedded aptamers that recognize theophylline and thiamine pyrophosphate, respectively, in tandem in the 5’-UTR of a bacterial mRNA and selected for translational riboswitches that responded to ligand presence like AND or NAND logic gates. Reporter gene activity was observed either only in the presence of both ligands (AND gate), or, in the case of the NAND gate, when neither or only one ligand was present [103]. In a subsequent, different experimental setup, they combined a transcriptional-OFF riboswitch with a translational-ON switch. Because both switches respond to the same ligand, thiamine pyrophosphate, albeit with different sensitivity, a band pass filter is the result, with maximum target gene expression obtained at intermediate ligand concentrations. At low ligand concentrations, gene expression is repressed translationally, while at high concentrations, premature transcription termination occurs [104]. Boolean logic gates can also be constructed with aptazymes. The Hartig group [105] combined ribozyme switches responsive to thiamine pyrophosphate and theophylline to yield AND, NOR, and ANDNOT gates. Win and Smolke [106] also generated an entire set of Boolean logic gates by combining theophylline- and tetracyclineresponsive aptamers with ON- and OFF-responsive ribozymes. For example, such gates were constructed either by assembling two different aptazymes in series (AND gate and NOR gate) or by, very ingeniously, exploiting the possibility of joining aptamer-responsive control elements to all stems of the hammerhead ribozyme. Coupling two ON-switches responsive to different inputs to different ribozyme stems led to an output signal only if no or one of the ligands was present, thus generating a NAND gate [106]. © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 3 What makes a riboswitch out of an aptamer? So far, in vivo active riboswitches rely on a small set of ligands – theophylline, tetracycline, neomycin [53], 2,4-dinitrotoluene [107], ammeline, 5-aza-cytosine [108] or pyrimido[4,5-d]pyrimidine-2,4-diamine (PPDA) [43], although several dozen small molecule-binding aptamers have been selected by a process called SELEX (systematic evolution of ligands by exponential enrichment), an in vitro technology invented about a decade before the discovery of natural riboswitches [109, 110]. During SELEX, a huge combinatorial library of RNA (and sometimes DNA) molecules is subjected to affinity of a small RNA chromatography to the ligand of choice. Non-binding sequences are removed during washing steps, whereas sequences with a certain binding affinity to the target of interest are eluted. These molecules are amplified and subjected to further rounds of selection. During these iterative cycles, the selection pressure can be increased steadily, finally resulting in molecules with sometimes extraordinarily high binding affinities and specificities, the so-called aptamers (Fig. 3A). However, as stated above, only a handful of such small molecule-binding aptamers have made it into the toolbox of RNA engineers. This raises the questions (i) what makes a riboswitch out of an aptamer; and (ii) how can we identify further aptamers which can be used for riboswitch engineering. A detailed analysis of neomycin-binding aptamers could shed some light on the properties such aptamers need in order to become regulatory molecules that are active in vivo. The neomycin-binding aptamer R23 was isolated by in vitro selection [111] and shows both high affinity and specificity for its ligand [111]. However, this aptamer – like many other small molecule-binding aptamers, too – showed no ligand-dependent regulation when tested for riboswitch activity in an assay for translational regulation (like in Fig. 2A) [53]. The subsequent analysis of the enriched aptamer pool (after seven rounds of selection) for sequences with regulatory activity using an in vivo screen, such as the one displayed in Fig. 3A, resulted in the isolation of a completely different aptamer, designated N1 and shown in Fig. 3B [53]. This new neomycin aptamer is active as riboswitch when inserted into the 5’-UTR of a reporter gene (Fig. 2A) [53], but also as sensing domain in the context of an allosteric ribozyme (Fig. 2C) [66]. Interestingly, N1 was not among the sequenced aptamers isolated after in vitro selection indicating that this aptamer was underrepresented in the pool enriched for binding aptamers [111]. A detailed genetic, biochemical and structural analysis of the aptamer N1 unravelled the molecular basis of this regulation. While the aptamer shows tight ligand binding with a Kd in the low nanomolar range, this affinity was still very similar to that of the originally identified in vivo inactive aptamer R23 [112]. However, substantial 251 Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com www.biotecvisions.com Figure 3. (A) Finding regulating aptamers. In the first step, several rounds of in vitro selection (SELEX) result in an enriched pool of aptamers that bind to the ligand. This pool is then cloned into the 5’-UTR of a GFP reporter gene. Yeast cells transformed with this library are screened for ligand-dependent changes in GFP expression. (B) The regulating neomycin-binding aptamer N1. The secondary structures of the ground- and ligand-bound states of the neomycin-binding aptamer N1 are shown. Nucleotides which form new base pairs in the bound state resulting in the stabilization of the RNA structure are marked in red. conformational changes concomitant to ligand binding were detected in N1, which were completely absent in the inactive aptamer R23 [113]. A short intervening sequence separating two short helical elements was identified as being important for regulation and therefore called “switching element”. It allows the ground state of the aptamer to retain an open, less structured conformation. Ligand binding then stabilizes a highly structured conformation, in which additional base pairs are formed. These elongate the two short helices and promote stacking of the upper onto the lower helix resulting in a long α-helical element (Fig. 3B). The ligand neomycin then acts like a clamp holding both the upper and the lower helix together. This conformation then is able to efficiently interfere with ribosomal scanning in yeast. It seems that the interplay between an open ground state and a highly structured ligand-bound state is important for the regulatory activity of the aptamer. 252 Further aptamers with confirmed riboswitch activity, like the malachite green or the theophylline aptamers, show a similar open, less structured ground state and a highly structured ligand-bound conformation [114, 115]. The conclusion drawn from this analysis is that regulating aptamers have to possess two properties: first, high affinity ligand binding, and second, significant conformational changes and stabilization upon ligand binding. The first point can be addressed in any classical SELEX experiment. By increasing the stringency of the selection process, high affinity binding aptamers can be obtained with binding constants even in the picomolar range [116]. Tight binding, however, is necessary but not sufficient for regulation. The second requirement – the conformational changes within the aptamer upon ligand binding – cannot be addressed by in vitro selection. An additional cellular screening step, preferably directly in the organism of © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotecvisions.com www.biotechnology-journal.com choice, is necessary to identify aptamers with regulatory activity. Several approaches have been undertaken which not only support the success, but also reinforce the importance of cellular screening for the identification of in vivo active regulatory devices. Gallivan and colleagues applied different screens to identify in vivo active riboswitches [83, 117], which proved to be extremely powerful in bacteria (see also Section 2.2). In vivo screening was also applied to identify aptamer-controlled ribozymes. Here, the communication module connecting the sensing domain (aptamer) with the catalytic domain (ribozyme) was randomized [118, 119]. Together with the neomycin screen described above, these examples underline the importance of this approach. 4 How can we further improve riboswitches? Chris Berens studied biology in Erlangen where he received his Ph. D. in microbiology. After postdoctoral work in the laboratory of Prof. Renée Schroeder on tetracycline-binding aptamers and metal ion binding sites in RNA, he went back to Erlangen to work on the development of inducible regulatory systems for application in both pro- and eukaryotes with a focus on studying cell death. He is currently at the Institute of Molecular Pathogenesis which is located at the Jena site of the Friedrich-Loeffler-Institut. Beatrix Suess studied biology in Greifswald and Erlangen where she received Despite of the success engineered riboswitches are enjoying as regulatory devices, there are still two important problems that have to be overcome to allow us to build even better riboswitches. One is the lack of a large number of alternative aptamers as ligand-binding modules and the second is the large discrepancy between the in vitro affinities of aptamers for their ligands and the high concentrations of ligand needed in vivo to flip the corresponding switches. Besides introducing an additional cellular screening step which allows the identification of regulating aptamers as specified above, reengineering the aptamer domain scaffolds of existing functional riboswitches to recognize novel ligands is also a promising approach to extend the ligand toolbox. The Micklefield group mutated a purine riboswitch to not recognize its natural ligand anymore and then screened this mutant pool for binding to synthetic, heterocyclic molecules. In the end, they obtained a ligand, PPDA, which controlled riboswitch activity efficiently [43, 108]. It will be interesting to see if this approach can be extended successfully to other natural riboswitches and, how different any potential novel ligands can be structurally from the respective native ligands. The many approaches to identify inhibitors of riboswitch activity for therapeutic applications might serve as starting points for finding such alternative ligands [120–122]. More extensive mutagenesis might contribute to this process, because natural sequence variants of riboswitches already display functional diversity, which is thought to reflect the adaptation to the idiosyncratic regulatory requirements of a specific gene [123]. Correspondingly, mutagenesis of residues that are not directly involved in ligand binding affect the regulatory properties of riboswitches [124, 125]. The ligand-binding site of the tetracycline aptamer is formed by a three-way-junction that is stabilized by longrange interactions and almost completely envelops the © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim her Ph.D.. After postdoctoral work in the lab of Prof. Dr. Wolfgang Hillen she started working on RNA aptamers and their use as synthetic control devices. In 2007, she was appointed as Aventis foundation-endowed professor for Chemical Biology at the Goethe-University Frankfurt, Main. Since 2012, she is full professor for Synthetic Biology at the Technical University Darmstadt. ligand [126]. This is reminiscent of natural riboswitches and distinguishes the tetracycline aptamer from the ligand-binding pockets of many aptamers which are much more open and solvent-exposed [127]. It might suggest that alternative aptamer selection protocols, like capillary electrophoresis, that do not require coupling of the ligand to a matrix and, thus, do not force surface exposure might be more advantageous for isolating in vivo functional aptamer domains [128–130]. A discrepancy between high affinity binding and efficient regulation has not only been observed for engineered riboswitches, but also for the tetrahydrofolate riboswitch. Although several ligand analogs and adenine derivatives were found to bind with high affinity to the riboswitch, they were not able to regulate its activity [131]. The mechanism(s) responsible for this effect are not clear at the moment, although for synthetic riboswitches, the fusion of an aptamer domain to an expression platform might result in a suboptimal interface between the two elements. Here, some optimization by mutagenesis and selection might be needed. A recent publication by the Gallivan lab might aid in this optimization process [132]. Mishler & Gallivan were able to recapitulate the approximately 1000-fold discrepancy between the concentrations needed for efficient binding of theophylline to its aptamer and the concentration needed for activating the 253 Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com riboswitch in E. coli in vitro in an S30 cell extract. This might finally allow in vitro screening and selection for more efficient switches requiring less ligand [132]. 5 Conclusions Riboswitch engineering has come a long way since the first applications were introduced approximately fifteen years ago. This technology has finally reached a stage in which it can not only complement other established mechanisms and strategies to regulate gene expression, but in which there are applications where we believe that it can even effectively replace these. The riboswitch engineer’s toolbox now provides a broad platform which should be able to offer flexible solutions to most synthetic biology applications requiring regulation of gene expression. Advances in selection technology and further improvement of our understanding of riboswitch structure and function will guide the design of better riboswitches to harness the full regulatory, diagnostic and therapeutic potential of aptamer-controlled regulatory devices. We would like to thank the Deutsche Forschungsgemeinschaft (SFB902/A2) and Loewe CGT for financial support. EU FP7-KBBE-2013-7 no. 613745, Promys The authors declare no financial or commercial conflict of interest. 6 References [1] Crick, F., Central dogma of molecular biology. Nature 1970, 227, 561–563. [2] Breaker, R. R., Joyce, G. F., The Expanding View of RNA and DNA Function. Chem. Biol. 2014, 21, 1059–1065. [3] Jeck, W. R., Sharpless, N. E., Detecting and characterizing circular RNAs. Nat. Biotechnol. 2014, 32, 453–461. [4] Li, L., Chang, H. Y., Physiological roles of long noncoding RNAs: insight from knockout mice. Trends Cell Biol. 2014, 24, 594–602. [5] van der Oost, J., Westra, E. R., Jackson, R. N., Wiedenheft, B., Unravelling the structural and mechanistic basis of CRISPR-Cas systems. Nat. Rev. Microbiol. 2014, 12, 479–492. [6] Hannon, G. J., RNA interference. Nature 2002, 418, 244–251. [7] Fellmann, C., Lowe, S. W., Stable RNA interference rules for silencing. Nat. Cell. Biol. 2014, 16, 10–18. [8] Callegari, E., Gramantieri, L., Domenicali, M., D’Abundo, L., et al., MicroRNAs in liver cancer: a model for investigating pathogenesis and novel therapeutic approaches. Cell Death Differ. 2015, 22, 46–57. [9] Harrison, M. M., Jenkins, B. V., O’Connor-Giles, K. M., Wildonger, J., A CRISPR view of development. Genes Dev. 2014, 28, 1859–1872. [10] Hsu, P. D., Lander, E. S., Zhang, F., Development and applications of CRISPR-Cas9 for genome engineering. Cell 2014, 157, 1262–1278. [11] Tucker, B. J., Breaker, R. R., Riboswitches as versatile gene control elements. Curr. Opin. Struct. Biol. 2005, 15, 342–348. [12] Tang, J., Breaker, R. R., Rational design of allosteric ribozymes. Chem. Biol. 1997, 4, 453–459. 254 www.biotecvisions.com [13] Soukup, G. A., Breaker, R. R., Engineering precision RNA molecular switches. Proc. Natl. Acad. Sci. USA 1999, 96, 3584–3589. [14] Werstuck, G., Green, M. R., Controlling gene expression in living cells through small molecule-RNA interactions. Science 1998, 282, 296–298. [15] Nahvi, A., Sudarsan, N., Ebert, M. S., Zou, X., et al., Genetic control by a metabolite binding mRNA. Chem. Biol. 2002, 9, 1043. [16] Mironov, A. S., Gusarov, I., Rafikov, R., Lopez, L. E., et al., Sensing small molecules by nascent RNA: a mechanism to control transcription in bacteria. Cell 2002, 111, 747–756. [17] Winkler, W., Nahvi, A., Breaker, R. R., Thiamine derivatives bind messenger RNAs directly to regulate bacterial gene expression. Nature 2002, 419, 952–956. [18] Breaker, R. R., Riboswitches and the RNA world. Cold Spring Harb. Perspect. Biol. 2012, 4, a03566. [19] Serganov, A., Nudler, E., A decade of riboswitches. Cell 2013, 152, 17–24. [20] Barrick, J. E., Corbino, K. A., Winkler, W. C., Nahvi, A., et al., New RNA motifs suggest an expanded scope for riboswitches in bacterial genetic control. Proc. Natl. Acad. Sci. USA 2004, 101, 6421–6426. [21] Suess, B., Hanson, S., Berens, C., Fink, B., et al., Conditional gene expression by controlling translation with tetracycline-binding aptamers. Nucleic Acids Res. 2003, 31, 1853–1858. [22] Topp, S., Gallivan, J. P., Emerging applications of riboswitches in chemical biology. ACS Chem. Biol. 2010, 5, 139–148. [23] Chang, A. L., Wolf, J. J., Smolke, C. D., Synthetic RNA switches as a tool for temporal and spatial control over gene expression. Curr. Opin. Biotechnol. 2012, 23, 679–688. [24] Wittmann, A., Suess, B., Engineered riboswitches: Expanding researchers’ toolbox with synthetic RNA regulators. FEBS Lett. 2012, 586, 2076–2083. [25] Vazquez-Anderson, J., Contreras, L. M., Regulatory RNAs: charming gene management styles for synthetic biology applications. RNA Biol. 2013, 10, 1778–1797. [26] Groher, F., Suess, B., Synthetic riboswitches – A tool comes of age. Biochim. Biophys. Acta 2014, 1839, 964–973. [27] Ahmadvand, D., Rahbarizadeh, F., Moghimi, S. M., Biological targeting and innovative therapeutic interventions with phage-displayed peptides and structured nucleic acids (aptamers). Curr. Opin. Biotechnol. 2011, 22, 832–838. [28] Burnett, J. C., Rossi, J. J., RNA-based therapeutics: current progress and future prospects. Chem. Biol. 2012, 19, 60–71. [29] Meyer, M., Scheper, T., Walter, J. G., Aptamers: versatile probes for flow cytometry. Appl. Microbiol. Biotechnol. 2013, 97, 7097–7109. [30] Smuc, T., Ahn, I. Y., Ulrich, H., Nucleic acid aptamers as high affinity ligands in biotechnology and biosensorics. J. Pharm. Biomed. Anal. 2013, 81-82, 210–217. [31] Winkler, W. C., Nahvi, A., Roth, A., Collins, J. A., Breaker, R. R., Control of gene expression by a natural metabolite-responsive ribozyme. Nature 2004, 428, 281–286. [32] Lee, E. R., Baker, J. L., Weinberg, Z., Sudarsan, N., Breaker, R. R., An allosteric self-splicing ribozyme triggered by a bacterial second messenger. Science 2010, 329, 845–848. [33] Loh, E., Dussurget, O., Gripenland, J., Vaitkevicius, K., et al., A transacting riboswitch controls expression of the virulence regulator PrfA in Listeria monocytogenes. Cell 2009, 139, 770–779. [34] DebRoy, S., Gebbie, M., Ramesh, A., Goodson, J. R., et al., A riboswitch-containing sRNA controls gene expression by sequestration of a response regulator. Science 2014, 345, 937–940. [35] Mellin, J. R., Koutero, M., Dar, D., Nahori, M. A., et al., Sequestration of a two-component response regulator by a riboswitch-regulated noncoding RNA. Science 2014, 345, 940–943. © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com [36] Mellin, J. R., Tiensuu, T., Becavin, C., Gouin, E., et al., A riboswitchregulated antisense RNA in Listeria monocytogenes. Proc. Natl. Acad. Sci. USA 2013, 110, 13132–13137. [37] Suess, B., Fink, B., Berens, C., Stentz, R., Hillen, W., A theophylline responsive riboswitch based on helix slipping controls gene expression in vivo. Nucleic Acids Res. 2004, 32, 1610–1614. [38] Topp, S., Reynoso, C. M., Seeliger, J. C., Goldlust, I. S., et al., Synthetic riboswitches that induce gene expression in diverse bacterial species. Appl. Environ. Microbiol. 2010, 76, 7881–7884. [39] Topp, S., Gallivan, J. P., Guiding bacteria with small molecules and RNA. J. Am. Chem. Soc. 2007, 129, 6807–6811. [40] Wachsmuth, M., Findeiß, S., Weissheimer, N., Stadler, P. F., Mörl, M., De novo design of a synthetic riboswitch that regulates transcription termination. Nucleic Acids Res. 2013, 41, 2541–2551. [41] Ceres, P., Garst, A. D., Marcano-Velazquez, J. G., Batey, R. T., Modularity of select riboswitch expression platforms enables facile engineering of novel genetic regulatory devices. ACS Synth. Biol. 2013, 2, 463–472. [42] Ceres, P., Trausch, J. J., Batey, R. T., Engineering modular ‘ON’ RNA switches using biological components. Nucleic Acids Res. 2013, 41, 10449–10461. [43] Robinson, C. J., Vincent, H. A., Wu, M. C., Lowe, P. T., et al., Modular riboswitch toolsets for synthetic genetic control in diverse bacterial species. J. Am. Chem. Soc. 2014, 136, 10615–10624. [44] Wieland, M., Hartig, J. S., Artificial riboswitches: synthetic mRNAbased regulators of gene expression. ChemBioChem 2008, 9, 1873– 1878. [45] Yen, L., Svendsen, J., Lee, J. S., Gray, J. T., et al., Exogenous control of mammalian gene expression through modulation of RNA selfcleavage. Nature 2004, 431, 471–476. [46] Win, M. N., Smolke, C. D., A modular and extensible RNA-based gene-regulatory platform for engineering cellular function. Proc. Natl. Acad. Sci. USA 2007, 104, 14283–14288. [47] Wieland, M., Hartig, J. S., Improved aptazyme design and in vivo screening enable riboswitching in bacteria. Angew. Chem. Int. Ed. Engl. 2008, 47, 2604–2607. [48] Wittmann, A., Suess, B., Selection of tetracycline inducible selfcleaving ribozymes as synthetic devices for gene regulation in yeast. Mol. Biosyst. 2011, 7, 2419–2427. [49] Reining, A., Nozinovic, S., Schlepckow, K., Buhr, F., et al., Threestate mechanism couples ligand and temperature sensing in riboswitches. Nature 2013, 499, 355–359. [50] Saragliadis, A., Krajewski, S. S., Rehm, C., Narberhaus, F., Hartig, J. S., Thermozymes: Synthetic RNA thermometers based on ribozyme activity. RNA Biol. 2013, 10, 1010–1016. [51] Grate, D., Wilson, C., Inducible regulation of the S. cerevisiae cell cycle mediated by an RNA aptamer-ligand complex. Bioorg. Med. Chem 2001, 9, 2565–2570. [52] Hanson, S., Berthelot, K., Fink, B., McCarthy, J. E., Suess, B., Tetracycline-aptamer-mediated translational regulation in yeast. Mol. Microbiol. 2003, 49, 1627–1637. [53] Weigand, J. E., Sanchez, M., Gunnesch, E. B., Zeiher, S., et al., Screening for engineered neomycin riboswitches that control translation initiation. RNA 2008, 14, 89–97. [54] Harvey, I., Garneau, P., Pelletier, J., Inhibition of translation by RNAsmall molecule interactions. RNA 2002, 8, 452–463. [55] Bayer, T. S., Smolke, C. D., Programmable ligand-controlled riboregulators of eukaryotic gene expression. Nat. Biotechnol. 2005, 23, 337–343. [56] Ogawa, A., Rational design of artificial riboswitches based on ligand-dependent modulation of internal ribosome entry in wheat germ extract and their applications as label-free biosensors. RNA 2011, 17, 478–488. © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.biotecvisions.com [57] Bocobza, S., Adato, A., Mandel, T., Shapira, M., et al., Riboswitchdependent gene regulation and its evolution in the plant kingdom. Genes Dev. 2007, 21, 2874–2879. [58] Cheah, M. T., Wachter, A., Sudarsan, N., Breaker, R. R., Control of alternative RNA splicing and gene expression by eukaryotic riboswitches. Nature 2007, 447, 497–500. [59] Croft, M. T., Moulin, M., Webb, M. E., Smith, A. G., Thiamine biosynthesis in algae is regulated by riboswitches. Proc. Natl. Acad. Sci. USA 2007, 104, 20770–20775. [60] Wachter, A., Tunc-Ozdemir, M., Grove, B. C., Green, P. J., et al., Riboswitch control of gene expression in plants by splicing and alternative 3’ end processing of mRNAs. Plant Cell 2007, 19, 3437– 3450. [61] Kim, D. S., Gusti, V., Pillai, S. G., Gaur, R. K., An artificial riboswitch for controlling pre-mRNA splicing. RNA 2005, 11, 1667–1677. [62] Kim, D. S., Gusti, V., Dery, K. J., Gaur, R. K., Ligand-induced sequestering of branchpoint sequence allows conditional control of splicing. BMC Mol. Biol. 2008, 9, 23. [63] Weigand, J. E., Suess, B., Tetracycline aptamer-controlled regulation of pre-mRNA splicing in yeast. Nucleic Acids Res. 2007, 35, 4179– 4185. [64] Ausländer, S., Ketzer, P., Hartig, J. S., A ligand-dependent hammerhead ribozyme switch for controlling mammalian gene expression. Mol. Biosyst. 2010, 6, 807–814. [65] Beilstein, K., Wittmann, A., Grez, M., Suess, B., Conditional control of mammalian gene expression by tetracycline-dependent hammerhead ribozymes. ACS Synth. Biol. 2014, DOI: 10.1021/sb500270h. [66] Klauser, B., Atanasov, J., Siewert, L. K., Hartig, J. S., Ribozyme-based aminoglycoside switches of gene expression engineered by genetic selection in S. cerevisiae. ACS Synth. Biol. 2014, DOI: 10.1021/ sb500062p. [67] Nomura, Y., Zhou, L., Miu, A., Yokobayashi, Y., Controlling mammalian gene expression by allosteric hepatitis delta virus ribozymes. ACS Synth. Biol. 2013, 2, 684–689. [68] Krol, J., Loedige, I., Filipowicz, W., The widespread regulation of microRNA biogenesis, function and decay. Nat. Rev. Genet. 2010, 11, 597–610. [69] Djuranovic, S., Nahvi, A., Green, R., A parsimonious model for gene regulation by miRNAs. Science 2011, 331, 550–553. [70] Gurtan, A. M., Sharp, P. A., The role of miRNAs in regulating gene expression networks. J. Mol. Biol. 2013, 425, 3582–3600. [71] Kurreck, J., RNA interference: from basic research to therapeutic applications. Angew. Chem. Int. Ed. Engl. 2009, 48, 1378–1398. [72] Beisel, C. L., Bayer, T. S., Hoff, K. G., Smolke, C. D., Model-guided design of ligand-regulated RNAi for programmable control of gene expression. Mol. Syst. Biol. 2008, 4, 224. [73] Tuleuova, N., An, C. I., Ramanculov, E., Revzin, A., Yokobayashi, Y., Modulating endogenous gene expression of mammalian cells via RNA-small molecule interaction. Biochem. Biophys. Res. Commun. 2008, 376, 169–173. [74] Kumar, D., An, C. I., Yokobayashi, Y., Conditional RNA interference mediated by allosteric ribozyme. J. Am. Chem. Soc. 2009, 131, 13906–13907. [75] Garai, P., Gnanadhas, D. P., Chakravortty, D., Salmonella enterica serovars Typhimurium and Typhi as model organisms: revealing paradigm of host-pathogen interactions. Virulence 2012, 3, 377–388. [76] Kirkpatrick, C. L., Viollier, P. H., Decoding Caulobacter development. FEMS Microbiol Rev 2012, 36, 193–205. [77] McCormick, J. R., Flärdh, K., Signals and regulators that govern Streptomyces development. FEMS Microbiol. Rev. 2012, 36, 206– 231. [78] McConnell, M. J., Actis, L., Pachon, J., Acinetobacter baumannii: human infections, factors contributing to pathogenesis and animal models. FEMS Microbiol. Rev. 2013, 37, 130–155. 255 Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com [79] Spadiut, O., Capone, S., Krainer, F., Glieder, A., Herwig, C., Microbials for the production of monoclonal antibodies and antibody fragments. Trends Biotechnol. 2014, 32, 54–60. [80] Wendisch, V. F., Microbial production of amino acids and derived chemicals: Synthetic biology approaches to strain development. Curr. Opin. Biotechnol. 2014, 30C, 51–58. [81] Barrick, J. E., Breaker, R. R., The distributions, mechanisms, and structures of metabolite-binding riboswitches. Genome Biol. 2007, 8, R239. [82] Lynch, S. A., Desai, S. K., Sajja, H. K., Gallivan, J. P., A high-throughput screen for synthetic riboswitches reveals mechanistic insights into their function. Chem. Biol. 2007, 14, 173–184. [83] Topp, S., Gallivan, J. P., Random walks to synthetic riboswitches – a high-throughput selection based on cell motility. ChemBioChem 2008, 9, 210–213. [84] Rudolph, M. M., Vockenhuber, M. P., Suess, B., Synthetic riboswitches for the conditional control of gene expression in Streptomyces coelicolor. Microbiology 2013, 159, 1416–1422. [85] Nakahira, Y., Ogawa, A., Asano, H., Oyama, T., Tozawa, Y., Theophylline-dependent riboswitch as a novel genetic tool for strict regulation of protein expression in Cyanobacterium Synechococcus elongatus PCC 7942. Plant Cell Physiol. 2013, 54, 1724–1735. [86] Ma, A. T., Schmidt, C. M., Golden, J. W., Regulation of gene expression in diverse cyanobacterial species by using theophylline-responsive riboswitches. Appl. Environ. Microbiol. 2014, 80, 6704–6713. [87] Verhounig, A., Karcher, D., Bock, R., Inducible gene expression from the plastid genome by a synthetic riboswitch. Proc. Natl. Acad. Sci. USA 2010, 107, 6204–6209. [88] Seeliger, J. C., Topp, S., Sogi, K. M., Previti, M. L., et al., A riboswitchbased inducible gene expression system for mycobacteria. PLoS One 2012, 7, e29266. [89] Reynoso, C. M. K., Miller, M. A., Bina, J. E., Gallivan, J. P., Weiss, D. S., Riboswitches for intracellular study of genes involved in Francisella pathogenesis. MBio. 2012, 3, e00253-12. [90] Demolli, S., Geist, M. M., Weigand, J. E., Matschiavelli, N., et al., Development of β-lactamase as a tool for monitoring conditional gene expression by a tetracycline-riboswitch in Methanosarcina acetivorans. Archaea 2014, 2014, 725610. [91] Berens, C., Thain, A., Schroeder, R., A tetracycline-binding RNA aptamer. Bioorg. Med. Chem. 2001, 9, 2549–2556. [92] Ketzer, P., Haas, S. F., Engelhardt, S., Hartig, J. S., Nettelbeck, D. M., Synthetic riboswitches for external regulation of genes transferred by replication-deficient and oncolytic adenoviruses. Nucleic Acids Res. 2012, 40, e167. [93] Ketzer, P., Kaufmann, J. K., Engelhardt, S., Bossow, S., et al., Artificial riboswitches for gene expression and replication control of DNA and RNA viruses. Proc. Natl. Acad. Sci. USA 2014, 111, E554–562. [94] Gardner, T. S., Cantor, C. R., Collins, J. J., Construction of a genetic toggle switch in Escherichia coli. Nature 2000, 403, 339–342. [95] Elowitz, M. B., Leibler, S., A synthetic oscillatory network of transcriptional regulators. Nature 2000, 403, 335–338. [96] Silva-Rocha, R., de Lorenzo, V., Mining logic gates in prokaryotic transcriptional regulation networks. FEBS Lett. 2008, 582, 1237– 1244. [97] Rossi, F. M., Kringstein, A. M., Spicher, A., Guicherit, O. M., Blau, H. M., Transcriptional control: rheostat converted to on/off switch. Mol. Cell 2000, 6, 723–728. [98] Becskei, A., Seraphin, B., Serrano, L., Positive feedback in eukaryotic gene networks: cell differentiation by graded to binary response conversion. EMBO J 2001, 20, 2528–2535. [99] Becskei, A., Serrano, L., Engineering stability in gene networks by autoregulation. Nature 2000, 405, 590–593. 256 www.biotecvisions.com [100] Nevozhay, D., Adams, R. M., Balazsi, G., Linearizer gene circuits with negative feedback regulation. Methods Mol. Biol. 2011, 734, 81–100. [101] Watson, P. Y., Fedor, M. J., The glmS riboswitch integrates signals from activating and inhibitory metabolites in vivo. Nat. Struct. Mol. Biol. 2011, 18, 359–363. [102] Sudarsan, N., Hammond, M. C., Block, K. F., Welz, R., et al., Tandem riboswitch architectures exhibit complex gene control functions. Science 2006, 314, 300–304. [103] Sharma, V., Nomura, Y., Yokobayashi, Y., Engineering complex riboswitch regulation by dual genetic selection. J. Am. Chem. Soc. 2008, 130, 16310–16315. [104] Muranaka, N., Yokobayashi, Y., A synthetic riboswitch with chemical band-pass response. Chem. Commun. 2010, 46, 6825–6827. [105] Klauser, B., Saragliadis, A., Ausländer, S., Wieland, M., et al., Posttranscriptional Boolean computation by combining aptazymes controlling mRNA translation initiation and tRNA activation. Mol Biosyst. 2012, 8, 2242–2248. [106] Win, M. N., Smolke, C. D., Higher-order cellular information processing with synthetic RNA devices. Science 2008, 322, 456–460. [107] Davidson, M. E., Harbaugh, S. V., Chushak, Y. G., Stone, M. O., Kelley-Loughnane, N., Development of a 2,4-dinitrotoluene-responsive synthetic riboswitch in E. coli cells. ACS Chem. Biol. 2013, 8, 234–241. [108] Dixon, N., Duncan, J. N., Geerlings, T., Dunstan, M. S., et al., Reengineering orthogonally selective riboswitches. Proc. Natl. Acad. Sci. USA 2010, 107, 2830–2835. [109] Ellington, A. D., Szostak, J. W., In vitro selection of RNA molecules that bind specific ligands. Nature 1990, 346, 818–822. [110] Tuerk, C., Gold, L., Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 1990, 249, 505–510. [111] Wallis, M. G., von Ahsen, U., Schroeder, R., Famulok, M., A novel RNA motif for neomycin recognition. Chem. Biol. 1995, 2, 543–552. [112] Weigand, J. E., Schmidtke, S. R., Will, T. J., Duchardt-Ferner, E., et al., Mechanistic insights into an engineered riboswitch: a switching element which confers riboswitch activity. Nucleic Acids Res. 2011, 39, 3363–3372. [113] Duchardt-Ferner, E., Weigand, J. E., Öhlenschlager, O., Schmidtke, S. R., et al., Highly modular structure and ligand binding by conformational capture in a minimalistic riboswitch. Angew. Chem. Int. Ed. Engl. 2010, 49, 6216–6219. [114] Flinders, J., DeFina, S. C., Brackett, D. M., Baugh, C., et al., Recognition of planar and nonplanar ligands in the malachite green-RNA aptamer complex. ChemBioChem 2004, 5, 62–72. [115] Latham, M. P., Zimmermann, G. R., Pardi, A., NMR chemical exchange as a probe for ligand-binding kinetics in a theophyllinebinding RNA aptamer. J. Am. Chem. Soc. 2009, 131, 5052–5053. [116] Müller, M., Weigand, J. E., Weichenrieder, O., Suess, B., Thermodynamic characterization of an engineered tetracycline-binding riboswitch. Nucleic Acids Res. 2006, 34, 2607–2617. [117] Lynch, S. A., Gallivan, J. P., A flow cytometry-based screen for synthetic riboswitches. Nucleic Acids Res. 2009, 37, 184–192. [118] Saragliadis, A., Klauser, B., Hartig, J. S., In vivo screening of liganddependent hammerhead ribozymes. Methods Mol. Biol. 2012, 848, 455–463. [119] Rehm, C., Hartig, J. S., In vivo screening for aptazyme-based bacterial riboswitches. Methods Mol. Biol. 2014, 1111, 237–249. [120] Deigan, K. E., Ferré-D’Amaré, A. R., Riboswitches: discovery of drugs that target bacterial gene-regulatory RNAs. Acc. Chem. Res. 2011, 44, 1329–1338. © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Biotechnology Journal Biotechnol. J. 2015, 10, 246–257 www.biotechnology-journal.com [121] Ham, Y. W., Humphreys, D. J., Choi, S., Dayton, D. L., Rational design of SAM analogues targeting SAM-II riboswitch aptamer. Bioorg. Med. Chem. Lett. 2011, 21, 5071–5074. [122] Lünse, C. E., Schmidt, M. S., Wittmann, V., Mayer, G., Carba-sugars activate the glmS-riboswitch of Staphylococcus aureus. ACS Chem. Biol. 2011, 6, 675–678. [123] Tomsic, J., McDaniel, B. A., Grundy, F. J., Henkin, T. M., Natural variability in S-adenosylmethionine (SAM)-dependent riboswitches: S-box elements in Bacillus subtilis exhibit differential sensitivity to SAM in vivo and in vitro. J. Bacteriol. 2008, 190, 823–833. [124] Stoddard, C. D., Widmann, J., Trausch, J. J., Marcano-Velazquez, J. G., et al., Nucleotides adjacent to the ligand-binding pocket are linked to activity tuning in the purine riboswitch. J. Mol. Biol. 2013, 425, 1596–1611. [125] Weigand, J. E., Gottstein-Schmidtke, S. R., Demolli, S., Groher, F., et al., Sequence elements distal to the ligand binding pocket modulate the efficiency of a synthetic riboswitch. Chembiochem 2014, 15, 1627–1637. [126] Xiao, H., Edwards, T. E., Ferré-D’Amaré, A. R., Structural basis for specific, high-affinity tetracycline binding by an in vitro evolved aptamer and artificial riboswitch. Chem. Biol. 2008, 15, 1125–1137. © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.biotecvisions.com [127] Hermann, T., Patel, D. J., Adaptive recognition by nucleic acid aptamers. Science 2000, 287, 820–825. [128] Berezovski, M. V., Musheev, M. U., Drabovich, A. P., Jitkova, J. V., Krylov, S. N., Non-SELEX: selection of aptamers without intermediate amplification of candidate oligonucleotides. Nat. Protoc. 2006, 1, 1359–1369. [129] Berezovski, M., Musheev, M., Drabovich, A., Krylov, S. N., NonSELEX selection of aptamers. J. Am. Chem. Soc. 2006, 128, 1410– 1411. [130] Mosing, R. K., Bowser, M. T., Isolating aptamers using capillary electrophoresis-SELEX (CE-SELEX). Methods Mol. Biol. 2009, 535, 33-43. [131] Trausch, J. J., Batey, R. T., A disconnect between high-affinity binding and efficient regulation by antifolates and purines in the tetrahydrofolate riboswitch. Chem. Biol. 2014, 21, 205–216. [132] Mishler, D. M., Gallivan, J. P., A family of synthetic riboswitches adopts a kinetic trapping mechanism. Nucleic Acids Res. 2014, 42, 6753–6761. [133] Henkin, T. M., Preface – Riboswitches. Biochim. Biophys. Acta 2014, 1839, 899. 257 ISSN 1860-6768 · BJIOAM 10 (2) 227–338 (2015) · Vol. 10 · February 2015 Systems & Synthetic Biology · Nanobiotech · Medicine 2/2015 Minimal cells Artificial chromosomes Expression control Synthetic Biology www.biotechnology-journal.com Special issue: Synthetic Biology. Synthetic biology has become more application oriented, by designing and implementing synthetic pathways in industrial biotechnology. This Special issue, edited by Roland Eils (German Cancer Research Center, DKFZ, and University of Heidelberg), Julia Ritzerfeld (German Cancer Research Center, Heidelberg) and Wolfgang Wiechert (IBG-1: Biotechnology, Forschungszentrum Jülich), includes contributions from the Helmholtz Initiative on Synthetic Biology and focusses on applications of synthetic biology in biotechnology. Image: © Sergey Nivens Fotolia.com Biotechnology Journal – list of articles published in the February 2015 issue. Editorial: Synthetic biology – ready for application Roland Eils, Julia Ritzerfeld, Wolfgang Wiechert http://dx.doi.org/10.1002/biot.201400842 Forum Synthetic biology’s self-fulfilling prophecy – dangers of confinement from within and outside Daniel Frank, Reinhard Heil, Christopher Coenen, Harald König http://dx.doi.org/10.1002/biot.201400477 Forum Synthetic biology and intellectual property rights: Six recommendations Timo Minssen, Berthold Rutz, Esther van Zimmeren http://dx.doi.org/10.1002/biot.201400604 Commentary Fully glycerol-independent microbial production of 1,3-propanediol via non-natural pathway: Paving the way to success with synthetic tiles Ewelina CelinĚska http://dx.doi.org/10.1002/biot.201400360 Commentary Chassis organism from Corynebacterium glutamicum: The way towards biotechnological domestication of Corynebacteria Víctor de Lorenzo http://dx.doi.org/10.1002/biot.201400493 Review RNA aptamers as genetic control devices: The potential of riboswitches as synthetic elements for regulating gene expression Christian Berens, Florian Groher and Beatrix Suess http://dx.doi.org/10.1002/biot.201300498 Review CRISPR genome engineering and viral gene delivery: A case of mutual attraction Florian Schmidt and Dirk Grimm http://dx.doi.org/10.1002/biot.201400529 Review Optogenetic control of signaling in mammalian cells Rapid Communication Protein design and engineering of a de novo pathway for microbial production of 1,3-propanediol from glucose Zhen Chen, Feng Geng and An-Ping Zeng http://dx.doi.org/10.1002/biot.201400235 Research Article Chassis organism from Corynebacterium glutamicum – a top-down approach to identify and delete irrelevant gene clusters Simon Unthan, Meike Baumgart, Andreas Radek, Marius Herbst, Daniel Siebert, Natalie Brühl, Anna Bartsch, Michael Bott, Wolfgang Wiechert, Kay Marin, Stephan Hans, Reinhard Krämer, Gerd Seibold, Julia Frunzke, Jörn Kalinowski, Christian Rückert, Volker F. Wendisch and Stephan Noack http://dx.doi.org/10.1002/biot.201400041 Research Article Synthetic secondary chromosomes in Escherichia coli based on the replication origin of chromosome II in Vibrio cholerae Sonja J. Messerschmidt, Franziska S. Kemter, Daniel Schindler and Torsten Waldminghaus http://dx.doi.org/10.1002/biot.201400031 Research Article Bacterial XylRs and synthetic promoters function as genetically encoded xylose biosensors in Saccharomyces cerevisiae Wei Suong Teo and Matthew Wook Chang http://dx.doi.org/10.1002/biot.201400159 Biotech Method Single-cell analysis reveals heterogeneity in onset of transgene expression from synthetic tetracycline-dependent promoters Ulfert Rand, Jan Riedel, Upneet Hillebrand, Danim Shin, Steffi Willenberg, Sara Behme, Frank Klawonn, Mario Köster, Hansjörg Hauser and Dagmar Wirth http://dx.doi.org/10.1002/biot.201400076 Biotech Method Engineering connectivity by multiscale micropatterning of individual populations of neurons Jonas Albers, Koji Toma and Andreas Offenhäusser http://dx.doi.org/10.1002/biot.201400609 Hannes M. Beyer, Sebastian Naumann, Wilfried Weber and Gerald Radziwill http://dx.doi.org/10.1002/biot.201400077 © 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.biotechnology-journal.com