Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-28T02:27:16.733Z Has data issue: false hasContentIssue false

Building a pipeline to identify and engineer constitutive and repressible promoters

Published online by Cambridge University Press:  19 October 2023

Eric J.Y. Yang
Affiliation:
Department of Biology, University of Washington, Seattle, WA, USA
Jennifer L. Nemhauser*
Affiliation:
Department of Biology, University of Washington, Seattle, WA, USA
*
Corresponding author: Jennifer L. Nemhauser; Email: jn7@uw.edu

Abstract

To support the increasingly complex circuits needed for plant synthetic biology applications, additional constitutive promoters are essential. Reusing promoter parts can lead to difficulty in cloning, increased heterogeneity between transformants, transgene silencing and trait instability. We have developed a pipeline to identify genes that have stable expression across a wide range of Arabidopsis tissues at different developmental stages and have identified a number of promoters that are well expressed in both transient (Nicotiana benthamiana) and stable (Arabidopsis) transformation assays. We have also introduced two genome-orthogonal gRNA target sites in a subset of the screened promoters, converting them into NOR logic gates. The work here establishes a pipeline to screen for additional constitutive promoters and can form the basis of constructing more complex information processing circuits in the future.

Type
Original Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press in association with The John Innes Centre

1. Introduction

Plant synthetic biology aims to provide greater control over plant form and function, a goal that is beginning to be realised. Several projects have produced measurable gains in photosynthetic efficiency (Batista-Silva et al., Reference Batista-Silva, da Fonseca-Pereira, Martins, Zsögön, Nunes-Nesi and Araújo2020; Orr et al., Reference Orr, Pereira, da Fonseca Pereira, Pereira-Lima, Zsögön and Araújo2017), and others have intervened in hormone response pathways to change plant architecture (Khakhar et al., Reference Khakhar, Leydon, Lemmex, Klavins and Nemhauser2018) or environmental response (Lim et al., Reference Lim, Mayer, Yim and Cushman2020; Park et al., Reference Park, Peterson, Mosquna, Yao, Volkman and Cutler2015). These advances rely on well-characterised promoters to ensure the expression of transgene in desired tissues.

Promoters can be broadly broken down into three categories based on expression pattern: constitutive, spatiotemporally restricted, and inducible (Peremarti et al., Reference Peremarti, Twyman, Gómez-Galera, Naqvi, Farré, Sabalza, Miralpeix, Dashevskaya, Yuan, Ramessar, Christou, Zhu, Bassie and Capell2010). Constitutive promoters are defined here as promoters expressed in all tissues at all times. These promoters regulate the transcription of what are commonly referred to as ‘housekeeping genes’. While each category of promoter is useful in plant engineering, constitutive promoters are often used to confer novel traits such as herbicide tolerance, to drive synthetic circuits, and used in metabolic engineering projects due to their broad tissue coverage (Bak & Emerson, Reference Bak and Emerson2020; Brophy et al., Reference Brophy, Magallon, Duan, Zhong, Ramachandran, Kniazev and Dinneny2022; South et al., Reference South, Cavanagh, Liu and Ort2019). Some of the most widely used plant constitutive promoters include variants from the Cauliflower Mosaic Virus 35S (35S) promoter, and promoters from members of the ubiquitin and actin families (Jiang et al., Reference Jiang, Zhang, Ding, He, Li, Zhu, Cheng, Zhang and Li2018; Peremarti et al., Reference Peremarti, Twyman, Gómez-Galera, Naqvi, Farré, Sabalza, Miralpeix, Dashevskaya, Yuan, Ramessar, Christou, Zhu, Bassie and Capell2010). However, the list of available plant constitutive promoters is short, and this lack of parts poses many challenges (Peremarti et al., Reference Peremarti, Twyman, Gómez-Galera, Naqvi, Farré, Sabalza, Miralpeix, Dashevskaya, Yuan, Ramessar, Christou, Zhu, Bassie and Capell2010). Having to reuse the limited number of promoters in increasingly complex plant gene circuits or metabolic engineering projects can quickly lead to instability of the transformed construct due to repeated elements rearranging and homology-dependent gene silencing, which is heritable (De Wilde et al., Reference De Wilde, Van Houdt, De Buck, Angenon, De Jaeger and Depicker2000; Peremarti et al., Reference Peremarti, Twyman, Gómez-Galera, Naqvi, Farré, Sabalza, Miralpeix, Dashevskaya, Yuan, Ramessar, Christou, Zhu, Bassie and Capell2010; Rajeev Kumar et al., Reference Rajeev Kumar, Anunanthini and Ramalingam2015).

To expand the number of promoters available, several groups have recently used distinct strategies to engineer both constitutively and conditionally expressed promoters. One approach builds synthetic promoters by adding cis-elements to a ‘minimal promoter region’, which is often 35S-derived. By varying the number and type of cis-elements, researchers were able to generate promoters with a wide range of expression levels and expression patterns (Ali & Kim, Reference Ali and Kim2019; Belcher et al., Reference Belcher, Vuu, Zhou, Mansoori, Agosto Ramos, Thompson, Scheller, Loqué and Shih2020; Brophy et al., Reference Brophy, Magallon, Duan, Zhong, Ramachandran, Kniazev and Dinneny2022; Liu & Stewart, Reference Liu and Stewart2016). Another approach uses sequences upstream of the minimal promoter region as a landing dock for synthetic activators guided by zinc-finger, TALE, or dCas9 to promote expression (Liu & Stewart, Reference Liu and Stewart2016). The expression strength of these promoters can be tuned by varying the number of target sites for the synthetic activators (Cai et al., Reference Cai, Kallam, Tidd, Gendarini, Salzman and Patron2020; Moreno-Giménez et al., Reference Moreno-Giménez, Selma, Calvache and Orzáez2022). These approaches, while quite powerful, are limited by the small number of characterised minimal promoters available to build upon and may still lead to repeated units in large constructs if the same minimal promoters were used.

Here, we employed an alternative approach for finding constitutive promoters. Instead of building and testing synthetic promoters, we looked to natural promoters found in the Arabidopsis genome. This approach has a few advantages. Synthetic promoters require extensive characterisation to determine their expression pattern and, because of practical constraints, are often only tested in a few selected tissues. In contrast, the wealth of RNA-seq data available for Arabidopsis provides highly detailed information about a given promoter’s likely expression potential, including the expression level of the gene throughout many developmental stages, tissue types, and even various growth and/or stress conditions. The expression of a native promoter has already been subject to selective pressures, and so is potentially more likely to remain stable across generations. By introducing a set of unique sequences, natural promoters also have the potential to minimise the likelihood of gene silencing or unwanted recombination through repeated units in multigenic constructs. Lastly, by employing some of the techniques in generating synthetic promoters described above, these native promoters could potentially form the foundation for suites of derived promoters with even more refined expression levels. A similar approach successfully expanded the range of promoter expressions available in B. subtilis (Guiziou et al., Reference Guiziou, Sauveplane, Chang, Clerté, Declerck, Jules and Bonnet2016). Since the argument for the need of additional promoter parts can be directly extended to the need for additional terminators, and terminators are known regulators of gene expression (Andreou et al., Reference Andreou, Nirkko, Ochoa-Villarreal and Nakayama2021; Wang et al., Reference Wang, Kumar, Zeng, McEwan, Wright and Gupta2020), we screened promoter-terminator pairs together. To further extend the utility of the new promoter/terminator pairs, we also introduced dCas9 target sites with sequences not found elsewhere in the Arabidopsis genome, thereby enabling specific repression by synthetic transcription factors without interfering with the cognate native genes.

2. Results

To identify the most stably expressing promoters available in the Arabidopsis genome, we analysed publicly available RNAseq datasets. The majority of the RNAseq dataset came from the Klepikova transcriptome profile which included multiple tissues from different development stages (Klepikova et al., Reference Klepikova, Kasianov, Gerasimov, Logacheva and Penin2016). We supplemented this dataset with an RNAseq dataset for pollen (Loraine et al., Reference Loraine, McCormick, Estrada, Patel and Qin2013), as this cell type was not represented in the Klepikova dataset. After processing the RNAseq datasets, there were 10,096 genes that were expressed in all the datasets (i.e. have none zero read counts) (Figure 1a). The coefficient of variation (CV) of expression across different tissues is often used as a metric for identifying stably expressed genes (Czechowski et al., Reference Czechowski, Stitt, Altmann, Udvardi and Scheible2005; Huang et al., Reference Huang, Li and Zhan2019; Wang et al., Reference Wang, Lyu, Pan, Zeng and Randhawa2019). Within the lowest 3% CV, there were 303 genes, which corresponds to a CV cutoff of 0.26 (Figure 1a,d). To facilitate dissemination of the parts quantified in this study, we adopted the Golden Gate MoClo system and cloned the promoter + 5UTR together as a standard MoClo part and similarly with 3UTR + terminator (Figure 1c; Engler et al., Reference Engler, Youles, Gruetzner, Ehnert, Werner, Jones, Patron and Marillonnet2014; Weber et al., Reference Weber, Engler, Gruetzner, Werner and Marillonnet2011). Since MoClo uses BsaI and BbsI type-II restriction enzymes for cloning, we removed any candidates with these restriction sites within the cloned regions. These cloning constraints left us with 61 candidate genes.

Figure 1. (a) Pipeline to identify constitutive promoters. The number of genes that pass each filter is indicated, along with the software used to implement the analysis. SRA is the ‘Sequence Read Archive’. Detailed methods, including parameters for each filter, are described in Section 4. (b) Schematic of filters used to select candidate promoters to engineer with synthetic gRNA target sites. (c) Schematic describing how we defined ‘promoter’ and ‘terminator’. The ‘promoter’ was defined here as starting from the transcription start site and going upstream to a maximum of 2,000 bp or to the next annotated neighbouring gene, whichever is shorter. Similarly, a ‘terminator’ was defined as starting from the transcription end site and going downstream a maximum of 250 bp or to the next annotated neighbouring gene, whichever is shorter. Promoters and terminators were cloned, along with their respective UTRs, following the Golden Gate MoClo system. (d) Plot showing values for the 10,096 genes expressed in all tissues. The geometric mean of expression across samples is plotted on the x-axis with the coefficient of variation (CV) on the y-axis. Both axes are on a base-10 log scale. The lowest 3% CV corresponds to a 0.26 CV cutoff, and the 303 genes with CV lower than 0.26 are highlighted in yellow. The final 33 candidates that fulfilled all criteria are highlighted in red. Several common promoters used in plant synthetic biology are annotated for reference.

To selectively activate or repress promoters in the context of a synthetic circuit, we wanted to modify segments of the promoter region to allow genome-orthogonal dCas9 targeting. While there are no specific guidelines on optimal placement for gRNA target sites in plants (Pan et al., Reference Pan, Sretenovic and Qi2021), studies in other eukaryotes have pointed to −50 to +300 bp from TSS in mammalian cells for CRISPRi, and within −200 bp from TSS in yeast (Jensen, Reference Jensen2018). Using the ‘Binding Site Prediction’ function from PlantRegMap we screened for predicted motifs within 500 bp of the promoter region from the TSS (Tian et al., Reference Tian, Yang, Meng, Jin and Gao2020). We retained promoters that could accommodate two 23 bp gRNA target sites (20 bp target sequence and 3 bp PAM site) without interrupting any predicted motifs and were at least 67 bp apart, following the spacing used in Gander et al. (Figure 1b). We were left with 33 candidate genes. Compared to the commonly used native Arabidopsis constitutive promoters, the candidates identified here were more stably expressed but have mostly weaker mean expression (Figure 1d). Detailed information of the 33 candidates can be found in Supplementary Table S2.

While one of the main goals of this study is to identify the best available natural stable genes through analysis of RNAseq data, the ‘stability’ of the candidate genes we screened for in this article is constrained by the choice of RNAseq dataset used. The Klepikova dataset included stress-treated leaf samples with heat, cold, and wounding treatments, but they were not included in the CV calculation since the samples were only collected from mature third leaves and no other tissue types. Instead, we normalised the stress data with untreated ‘mature whole third leaf’ and calculated their CV and included the result in Supplementary Table S2 for reference. Similarly, while the datasets capture coarse temporal resolutions throughout development, they cannot identify the fluctuation of circadian genes and therefore we supplemented the final table with identified circadian genes from CGDB for reference (Li et al., Reference Li, Shui, Zhang, Lv, Deng, Ullah, Zhang and Xue2017).

Of the 33 stably expressed genes identified from the bioinformatics pipeline, we successfully cloned 22 promoter-terminator pairs. We tested the promoters in Nicotiana benthamiana (tobacco) transient agroinfiltration assays and identified 16 promoters that had expressions that were significantly different from the negative control (Figure 2).

Figure 2. We identified 16 promoters that expressed in N. benthamiana. Six promoters were modified to introduce gRNA target sites. These sites are designated by brackets following the gene name. Three different constructs were injected per leaf, each containing a promoter to be tested driving NLS_YFP and an internal control of pUBQ10:NLS_mTURQ. Each leaf also has a negative control injection that only contains pUBQ10:NLS_mTURQ. Normalisation is performed using the formula: $\frac{{\mathrm{YFP}}_{\mathrm{promoter}}-\mathrm{median}\left({\mathrm{YFP}}_{\mathrm{Neg}.\mathrm{control}}\right)}{{\mathrm{mTURQ}}_{\mathrm{promoter}}}$ . For each construct, the three replicates with median fluorescence levels closest to the median of the group were selected for visualisation and statistical analysis. Each biological replicate is represented by a beeswarm plot of 24 datapoints (12 per leaf disc, 2 disc per injection) collected from the plate reader as well as a single summarising datapoint representing the median. The boxplots represent all biological replicates. Significance test was performed using Dunnett’s test for comparing multiple treatments with control at 95% family wise confidence level. Non-significant constructs are marked as NS. For a given construct, the colours signify datapoints derived from the same biological replicate.

To determine whether the promoters showed constitutive expression in Arabidopsis, 12 of the promoters were selected to drive expression of the RUBY reporter (He et al., Reference He, Zhang, Sun, Zhan and Zhao2020) in stable transformants. Since RUBY is a pigment that allows for simple visual readout, we were hoping it would be an effective way of evaluating the expression of the promoters in all the tissues throughout development. Three representative T1 lines were selected for each construct and six T2s per T1 line were observed at the seedling stage (12 days) and as mature plants (day 34). Eleven of the 12 promoters transformed showed expression in N. benthamiana, yet we only identified three promoters that displayed RUBY expression in Arabidopsis. Representative individuals are shown in Figure 3a (Supplementary Figure S3) with the intensity of RUBY colouring quantified in Fiji (Supplementary Figure S4).

Figure 3. (a) Three promoters showed expression of RUBY in Arabidopsis T2 plants. The flowers, siliques, and leaves were imaged on day 34, while the seedling images were imaged on day 12. The inset boxes are the same images at higher magnification. Red arrows indicate areas where RUBY expression is visible by the eye. (b) qPCR data on T2 whole seedlings in three biological replicates for each line and (c) qPCR data on tissues collected from T3 plants, each with three biological replicates and two technical replicates. RUBY expression was normalised against the reference gene PP2AA3, and the bars represent the mean expression of the RUBY reporter and the standard error of the mean (SEM).

Given that expression in N. benthamiana doesnot perfectly predict expression in Arabidopsis, we included two promoters (AT1G54080, AT1G71860) that did not show expression in tobacco infiltration in our Arabidopsis stable transformation experiment. Interestingly, AT1G54080 displayed RUBY expression in roots and pollen. AT5G37830 had visible expression in pollen, siliques, stems, and roots. AT3G08530 had the most ubiquitous expression and had visible expression in the flowers, pollen, siliques, stems, and roots. A visual summary of the Arabidopsis and N. benthamiana experiments can be found in Supplementary Figure S5.

Given that the majority of the promoters had no visible expression of RUBY by eye, we performed qPCR on whole seedlings from four promoter lines: two with visible RUBY expression in roots (AT1G54080 and AT5G37830) and two without (AT1G64550 and AT2G29080). The two lines without visible RUBY expression both had qPCR expression level between the two lines with visible RUBY expression in the roots. This result suggests that RUBY was not a reliable reporter for these low-expressing promoters and that the promoters were indeed functional in Arabidopsis (Figure 3b). To further confirm whether the promoters were truly constitutively expressed, we chose our strongest expressing line (Figure 3a) and one of the lines that did not appear red but had detectable expression by qPCR (Figure 3b) for more careful qPCR analysis on seedlings, adult roots, flowers, and leaves (Figure 3c). The result confirms that RUBY expression was detected in all the tissues analysed, even when the tissues might not appear red by visual inspection. An interesting observation from the qPCR experiments was that the expression level of RUBY mRNA detected through qPCR is weaker than expected from the RNAseq dataset. While the predicted expression level of all four of the genes in the qPCR experiment are higher than the reference gene PP2AA3, the measured result showed the opposite (Supplementary Table S7). This discrepancy could be attributed to the RUBY reporter or potential limitations in identifying additional transcriptional regulators (see Section 3).

To make the promoters screened in this experiment more versatile, we next introduced two gRNA target sites into six of the promoters screened with target site sequences not found in the Arabidopsis genome. A constitutive promoter with two unique gRNA target sites can function as a NOR gate (a two-input logic gate where the output is only ON when neither of the inputs is present) in the presence of a dCas9-guided repressor. The inputs for such a gate are the gRNAs. When either or both of the gRNAs are present, the dCas9-guided repressor should be able to keep the promoter OFF (Figure 4b). Only when neither of the guides is present can the promoter be turned ON. Nine different functional gRNA sequences (A–I) were selected from the literature (Supplementary Table S8).

Figure 4. (a) The four constructs co-injected for each injection. The injection always contains the mPromoter, dCas9-guided repressor, and the two self-cleaving input gRNAs. The gRNAs are denoted with X and Y representing a variable input. (b) Schematic of the NOR gate when both input gRNAs are present. (c) Pattern of injection for the four possible input combinations and the gRNAs used for each injection. (1,1) represents both guides are present while (0,0) represents neither is present. When a guide is not present, a non-matching gRNA is injected in its place, denoted here as gRNA_M or N. (d) Five of the six mPromoters functioned as NOR gates. All guides apart from gRNA_F are independently repressible. Each biological replicate is represented by a beeswarm plot of 24 datapoints (12 per leaf disc, 2 disc per injection) collected from the plate reader as well as a single summarising datapoint representing the median. The boxplots represent all biological replicates. The signal is measured as the YFP fluorescence (driven by the promoters being tested) divided by the mTURQ fluorescence (driven by pUBQ10). In each set of NOR gate injections, the (0,0) injection serves as the unrepressed control, and the dataset is normalised by dividing all values by the median of the unrepressed control on a per-leaf basis. The y-axis represents fold changes from the unrepressed control and each biological replicate of the control is centred on 1. Each colour represents a unique leaf. Letters above each boxplot are Compact Letter Display (CLD) for all pairwise comparisons within each set of injections using ANOVA followed by Tukey’s Honest Significant Difference Test. Numbers above the boxplot represent fold repression between the (0,0) and (1,1) injections.

We first confirmed that the introduction of the target sites did not abolish promoter expression (Figure 2). While in most cases there was little difference in expression between modified and native promoters, in one case (AT1G64550(E,A)), the expression level increased dramatically, possibly due to the introduction of new TF binding sites at the junction of the introduced gRNA target-site (Supplementary Figure S6). The repressibility of the modified promoters was tested in N. benthamiana with each infiltration containing all constructs required for repression (Figure 4a). Each set of experiment contains four possible input combinations for each repressible promoter and the extent of repression was evaluated against the non-repressed control using two non-matching gRNAs (Figure 4c,d). Five of the modified promoters (mPromoters) functioned as NOR gates while AT3G18480(F,G) acted as a NOT gate with input2 (gRNA_G) (Figure 4d). Of the NOR gates, AT2G26780(A,B) repressed to similar extents with either or both inputs. AT1G64550(E,A), AT2G29080(D,C) and AT3G08530(H,I) all displayed additive effects where having both inputs gave stronger repression than just having one alone. AT5G37830(B,E) had a well-repressed target site with input2 (gRNA_E) while input1 (gRNA_B) alone resulted in a weaker repressed state. The strongest repression was observed for AT5G37830(B,E) and AT1G64550(E,A) with about a twofold repression between input(1,1) and input(0,0), while the rest of the promoters had around a 1.2-fold repression. The repression strength observed in the assay is quite modest, and it is likely because the promoters are quite weak to begin with, making strong repression more difficult. The result displayed as normalised fluorescence and not as fold repression can be found in Supplementary Figure S9.

3. Discussion

Constitutive promoters are essential staples in stocking the synthetic biology toolbox. They are versatile due to their wide expression coverage, and form the foundation from which many synthetic promoters are built. Here, we report on the establishment of a pipeline to find the most stably expressing promoters in Arabidopsis. We successfully used this approach to identify 16 promoters that are predicted to be more stably expressed than some of the most widely used native plant constitutive promoters, and showed they can drive expression in transient transformations of N. benthamiana. We attempted to capture the expression pattern of these promoters in stably transformed Arabidopsis using the visual RUBY reporter and uncovered limitations in its utility, but we identified at least two promoters that showed expression in all the tissues tested throughout development via qPCR. Lastly, we engineered repressible versions of six promoters and showed that five of these can function as NOR logic gates.

One of the biggest challenges in having a small selection of promoters to choose from is the need to reuse promoters in larger constructs, which could pose challenges to long-term stability. The promoters identified in this article were selected from some of the most stably expressed genes available in the Arabidopsis genome and all have distinct sequences. A lack of promoter parts also means a lack of flexibility when it comes to the range of expression strength. Most of the promoters used in plant synthetic biology are quite strong and that is not ideal for every application. The availability of weaker, broadly expressed promoters like those characterised here allows more flexibility in promoter choices when excess production of target proteins can be a problem. For example, they can be beneficial in avoiding toxic intermediates or optimising flux in metabolic engineering projects (Brückner et al., Reference Brückner, Schäfer, Weber, Grützner, Marillonnet and Tissier2015; Patron, Reference Patron2020). If a minimal promoter sequence can be identified from these native promoters, they can also serve as the foundation of additional synthetic promoters where the expression pattern and strength can be freely modified by adding cis-elements or synthetic transcription factor binding sites. The pipeline employed in this article to arrive at new native constitutive promoters should be readily adaptable to other organisms if there is sufficiently broad sampling of transcriptomes and a reference genome. The pipeline could also be modified to identify native promoters with particular expression patterns. One caveat is that the promoters that can be extracted in this way are, by definition, limited by what is naturally available in the organism. On the other hand, they have the advantage of already being assayed in a whole range of tissue types and developmental stages – a breadth of information that can be logistically challenging to collect for synthetic promoters. It will be interesting to see if synthetic devices made with these modified native promoters prove more resilient to mutation than those using fully engineered promoters, as these sequences have presumably maintained stable expression in the face of mutation and selection.

Evaluating the promoters using RUBY revealed that the novel reporter had limited sensitivity when driven by weaker promoters. We were able to detect RUBY expression in seedlings and adult tissues without visible colouration using qPCR, a more sensitive assay. However, it is important to note that detecting transcripts doesnot always imply comparable levels of protein production due to post-transcriptional and post-translational regulation. In our design, we attempted to capture the effects of any post-transcriptional regulation by including the UTRs, but other potential transcriptional regulators could be missed. The lower-than-expected RUBY mRNA levels detected could be due to such regulators. Promoter-proximal introns after the translation start codon, for example, would not be captured in the cloning pipeline though it is known to contribute to gene expression (Rose, Reference Rose2019; Rose et al., Reference Rose, Elfersi, Parra and Korf2008). Distally located regulatory regions would also not be captured, but they should be rare in the compact genome of Arabidopsis (Galli et al., Reference Galli, Feng and Gallavotti2020; Lu et al., Reference Lu, Marand, Ricci, Ethridge, Zhang and Schmitz2019).

Working with native promoters also provided an opportunity to learn more about the biology of promoters themselves. Yamamoto and colleagues suggested that plant promoters can be grouped into a few core promoter categories based on the presence or absence of certain location-sensitive motifs (Yamamoto et al., Reference Yamamoto, Yoshioka, Hyakumachi and Obokata2011). Interestingly, they reported that TATA-box containing promoters tend to be regulated promoters while Coreless promoters (promoters that donot have any characteristic location-sensitive motifs) tend to be constitutively expressed. The vast majority of the constitutive promoters used today in plant synthetic biology are from the TATA promoter class, and we also have a much better understanding of how their expression is regulated (Cai et al., Reference Cai, Kallam, Tidd, Gendarini, Salzman and Patron2020). If the goal is to find constitutive promoters, however, the analysis by Yamamoto and colleagues would suggest that we should look to Coreless promoters instead. Indeed, only 9% (3/33) of the candidate genes identified in this study contain TATA boxes, while 45% (15/33) are Coreless (Supplementary Table S2).

The ability to selectively activate and repress genes provides the tools necessary to perform Boolean logic, which would allow more complex computations (Kassaw et al., Reference Kassaw, Donayre-Torres, Antunes, Morey and Medford2018). Plants naturally perform complex computations to determine when and where a gene should be expressed by integrating internal and external signals, and genetic logic gates provide a modular way to synthetically construct these input–output relationships by using simple genetic parts. There are many ways to achieve the different logic operations using molecular biology (Patron, Reference Patron2020). A NOR gate is powerful in that it can be used to construct any logic gate by just stringing together multiple NOR gates, and its efficacy has been demonstrated in yeast (Gander et al., Reference Gander, Vrana, Voje, Carothers and Klavins2017). To date, the feasibility of building more complex logic circuits in plants has been hindered by the lack of unique and strongly repressible promoter parts. With just our design constraints and no additional refinement, five of the six NOR gates built showed the correct behaviour, suggesting the pipeline used holds promise in identifying additional promoter candidates for engineering. The repressible promoters evaluated in this work can be further improved through additional design-build-test cycles to optimise the individual gRNA target sites by varying their position and sequence. The repressor design can also be potentially improved upon to lower the overall OFF state. While N. benthamiana serves as a great prototyping platform, the performance of the gates would also need to be evaluated in stable Arabidopsis lines to validate their viability. The pipeline and the repressible promoter screened in this work contributes to the construction of more complex, synthetic plant logic operations in the future.

4. Methods

4.1. Downloading and processing RNA-seq datasets

We used a custom UseGalaxy pipeline to process the RNA-seq datasets (Afgan et al., Reference Afgan, Baker, van den Beek, Blankenberg, Bouvier, Čech, Chilton, Clements, Coraor, Eberhard, Grüning, Guerler, Hillman-Jackson, Von Kuster, Rasche, Soranzo, Turaga, Taylor, Nekrutenko and Goecks2016). SRR accession codes from BioProject IDs PRJNA314076 (138 samples; Klepikova et al., Reference Klepikova, Kasianov, Gerasimov, Logacheva and Penin2016), PRJNA268115 (20 samples; Klepikova et al., Reference Klepikova, Logacheva, Dmitriev and Penin2015), PRJNA324514 (32 samples), PRJNA194429 (2 samples; Loraine et al., Reference Loraine, McCormick, Estrada, Patel and Qin2013) were input into ‘Faster Download and Extract Reads in FASTQ (Galaxy Version 2.10.8+galaxy0)’ with default settings. The FASTQ files were pipped into ‘FastQC (Galaxy Version 0.73+galaxy0)’ and ‘Trimmomatic (Galaxy Version 0.38.0)’ with sliding window trimming averaging across 4 bases with required average quality 20, and a minimum read length of 36. The trimmed files were input into ‘HISAT2 (Galaxy Version 2.1.0+galaxy5)’ with reference genome assembly TAIR10 and Araport11 genome annotation from The Arabidopsis Information Resource (TAIR). Minimum intron length was set to 60, and maximum intron length was set to 6000 (Marquez et al., Reference Marquez, Brown, Simpson, Barta and Kalyna2012). Features from the Araport11 annotation were counted with ‘htseq-count (Galaxy Version 0.9.1)’ set to Union Mode and counting only reads within regions defined as ‘exons’ in the Araport11 annotation while not counting non-unique/ambiguous reads (Klepikova et al., Reference Klepikova, Kasianov, Gerasimov, Logacheva and Penin2016). The counted features were downloaded, and subsequent analysis was done in R (R Core Team, 2022).

4.2. Identifying stable promoters

All samples excluding the stress dataset (PRJNA324514) were normalised using the Median Ratios method from the DESeq2 package in R (Love et al., Reference Love, Huber and Anders2014). The coefficient of variation (CV) for each gene was calculated from the normalised data. Genes with the lowest 3% CV were kept for further analysis. The stress dataset from PRJNA324514 was normalised with ‘mature whole third leaf’ from PRJNA314076 for CV calculation, separate from the rest of the data.

4.3. Extracting promoter and terminator sequences

Promoter+5UTR region (from before the start codon and extending upstream till the first annotated neighbouring gene or to a maximum of 2 kb from the transcription start site, whichever is shorter) and 3UTR+terminators (from after the stop codon and extending downstream till the first annotated neighbouring gene or to a maximum of 250 bp past the transcription end site, whichever is shorter) of the remaining genes were extracted using the Araport11 genome annotation and the ‘3,000 bp upstream and downstream’ sequence files from the TAIR website. The extracted sequences were screened for BbsI and BsaI restriction enzyme cut sites and only those without were kept. Any genes with their promoter+5UTR and 3UTR+terminator overlapping annotations from neighbouring genes in the Araport11 annotation were also removed.

4.4. Transcription factor binding site prediction

The promoter sequences of the remaining genes were uploaded onto PlantRegMap using the ‘Binding Site Prediction’ function (Tian et al., Reference Tian, Yang, Meng, Jin and Gao2020) and the predicted motifs for each promoter sequence were downloaded. Only genes that can fit two 23 bp gRNA target sites at least 67 bp apart without interrupting any of the predicted motifs while being within 500 bp of the TSS were kept.

4.5. Annotating candidate genes

For the final 33 candidate genes, CV for the Stress Dataset (StressCV) and promoter and terminator sequences were extracted as described above. The promoter core type was annotated by Tokizawa et al. (Reference Tokizawa, Kusunoki, Koyama, Kurotani, Sakurai, Suzuki, Sakamoto, Kurata and Yamamoto2017). A list of experimentally determined circadian genes in Arabidopsis was downloaded from CGDB (Li et al., Reference Li, Shui, Zhang, Lv, Deng, Ullah, Zhang and Xue2017), and any UniprotKB identifiers were converted to ATG identifiers with the Uniprot Retrieve/ID mapping tool. Gene Descriptions (Representative Gene Model Name, Gene Description, Gene Model Type, Primary Gene Symbol, and All Gene Symbols) were retrieved from TAIR.

4.6. Construction of plasmids

Promoter+5UTR and 3UTR+terminator for each candidate genes as defined above were cloned with PCR from extracted genomic Col-0 DNA into their respective MoClo level zero acceptors (pICH41295 and pICH41276, respectively) (Engler et al., Reference Engler, Youles, Gruetzner, Ehnert, Werner, Jones, Patron and Marillonnet2014). The promoter and terminator pair of the candidate genes was paired with nuclear-localised Venus to make level one constructs in ‘position one’. Venus level one constructs were paired with pUBQ10 promoter driving nuclear localised mTURQ with an Act2 terminator from the MoClo Plant Parts Kit (pICH44300) in ‘position two’ to form ratio-metric lvl2s with a binary Ti vector backbone (pAGM4673 or pAGM4723). RUBY from (He et al., Reference He, Zhang, Sun, Zhan and Zhao2020) was cloned into level zero constructs and then cloned directly into level-2 Ti vector backbone with the promoter and terminator pairs (pICH86966). List of primers and plasmid maps can be found in Supplementary Tables S10 and S11 and Genbank files can be found in Supplementary Data S13. Plasmids used in this manuscripts are available at https://www.addgene.org/Jennifer_Nemhauser/with IDs 205359 - 205408.

4.7. gRNA target-site introduction

gRNA target sites were cloned into regions that do not disrupt any predicted TF binding sites (as described above) through Gibson assembly by replacing the original sequence (Gibson et al., Reference Gibson, Young, Chuang, Venter, Hutchison and Smith2009). Primers can be found in Supplementary Table S10.

4.8. Agrobacterium infiltration

In total, 5 mL cultures of Agrobacterium containing constructs to be injected along with a separate 25 mL culture of P19 (Win & Kamoun, Reference Win and Kamoun2004) were grown overnight at 30C with the appropriate antibiotics. On the following day, the overnights were centrifuged at 3,000 × g for 10 minutes. The pellets were resuspended with 1 mL MMA (10 mM MgCl2, 10 mM MES (pH 5.6), 100 uM acetosyringone). The OD of the cultures were measured and about 1~2 mL volume mixture with 5.0 OD for construct to be tested and 5.0 OD for P19 were prepared. The infiltration mix was rotated to mix at room temperature for 3 hours before injecting into fully emerged N. benthamiana leaves with a 1mL syringe. The injections were always injected as triplicates on three separate leaves on three separate tobacco plants. Each leaf is also always injected with a pUBQ10:mTURQ control.

4.9. Fluorescence quantification in N. benthamiana

At 3 days post infiltration, the leaves were clipped off and visualised in the Azure C600 Western Blot imaging system with exposure times Cy5 = 0 sec, Cy3 = 15 sec, Cy2 = 5 sec. Two hole punches were taken out of representative regions of each injection, and damaged regions with high background fluorescence were avoided. The leaf discs were placed in a 96-well plate on top of 200 uL of water, and the plates were read with a TECAN SPARK plate reader with YFP: excitation 506(15) and emission 541(15), Gain 100. mTurq: excitation 430(15) and emission 480(15), Gain 50. mScarlet: excitation 565(15) and emission 600(15), Gain 100. Settings: Multiple Reads Per Well; Circle (Filled) 4 × 4 with border 800 uM. Each leaf disc was read 12 times giving a total of 24 datapoints per injection per leaf. The output data was read into a custom R file for annotation, clean-up and visualisation. Multiple biological replicates were assayed for each promoter being tested, and the three replicates closest to the median of all replicates were kept for visualisation and statistical analysis. Each injection’s YFP value is subtracted by the median YFP value of the pUBQ10:mTURQ negative control on the same leaf. The YFP value is then divided by the mTURQ value for each injection to normalise across results. A Dunnett Test, a post hoc pairwise multiple comparison test from the DescTools package, was used to determine whether the injections were significantly different from the negative control (Signorell et al., Reference Signorell, Aho, Alfons, Anderegg and Aragon2022).

4.10. Repression assays

Repression assays were performed using the modified promoters (mPromoters) with gRNA target-sites driving NLS-YFP and pUBQ10:NLS-mTURQ internal control as the reporter. The mPromoters (5OD) were co-injected with P19 (1OD), TPL repressor (1OD), self-cleaving gRNA_1 (1OD), and self-cleaving gRNA_2 (1OD). To test the mPromoters’ NOR gate functionality, two gRNA inputs were required. In cases where only one input is present, the other self-cleaving gRNA will be a non-matching guide to the mPromoter. When neither inputs are present, sometimes two non-matching gRNAs (1OD each) were co-injected and sometimes only one (2OD), but the final total OD were always consistent. The exact injection combinations can be found in the R script. The TPL repressor construct contains pUBQ1:tdTomato-pUBQ10:dCas9_TPL(N188) and is modified from Khakhar et al. (Reference Khakhar, Leydon, Lemmex, Klavins and Nemhauser2018). Self-cleaving gRNAs were designed in accordance with Zhang et al. (Reference Zhang, Gao, Wang and Zhao2017), and the modifications (gRNA and complementary sequences) were introduced in one step using Q5 mutagenesis (NEB) and were placed in the MoClo pICH86988 acceptor with a 35S promoter. The four possible input combinations for the NOR gate for each promoter were always injected on the same leaf, and the result was read with a plate reader as described above. The YFP value of each injection was divided by the mTURQ value to normalise the data, and the value of each injection was divided by the median of the no-input control. An ANOVA followed by a Tukey’s Honest Significant Difference Test was used to determine significant differences in expression between samples. List of plasmid maps used can be found in Supplementary Table S11 and Genbank files in Supplementary Data S13.

4.11. RUBY expression in Arabidopsis

Constructs with the candidate promoters driving RUBY were transformed into Col-0 through floral dipping method (Clough & Bent, Reference Clough and Bent1998). T1 seeds were selected on 0.5× LS + 50 ug/mL Kanamycin+0.8% bactoagar. Plates were stratified for 2 days, light pulsed for 6 hours then kept in the dark for 3 days. Resistant seedlings were transplanted to soil to collect T2 seeds. For each promoter lines, three representative T1 lines were chosen to have their T2 seedlings phenotyped, and for each line, 19 T2 seeds were plated on 120 × 120 × 17 square petri dishes with 0.5× LS + 0.8% bactoagar without selection. The plates were imaged on day 4, 8, and 12 post-germination, and six representative seedlings were transplanted to soil. The plants were imaged with a digital camera on day 34. The flowers were imaged under a Leica S8AP0 dissecting scope. A representative leaf, a segment of the inflorescence, and silique were placed between two clear projector sheets and scanned with a flatbed scanner.

4.12. RUBY redness quantification

Images of were loaded into Fiji (Schindelin et al., Reference Schindelin, Arganda-Carreras, Frise, Kaynig, Longair, Pietzsch, Preibisch, Rueden, Saalfeld, Schmid, Tinevez, White, Hartenstein, Eliceiri, Tomancak and Cardona2012), and then converted to Lab stack to isolate the a* stack. The default colour of the extracted stack was green, so the image was further converted to an RGB stack so that the Green-channel could be used for region of interest (ROI) quantification.

4.13. qPCR

T2 seedlings were grown vertically on 0.5 × LS + 0.8% Phytoagar and without selection. The plates were stratified at 4°C for 2 days. On day 12, approximately five seedlings per replicate were frozen in liquid nitrogen. Three biological replicates were prepared for each genotype. Col-0 seedlings were also collected on day 12 as a single biological replicate. For T3 tissues, seedlings were grown on vertical 0.5 × LS + 0.8% Phytoagar plates without selection. On day 10, three sets of four seedlings were collected for each line and frozen in liquid nitrogen. Four seedlings per line were transplanted to fresh LS + Phytoagar plates to collect older roots from, and the rest of the seedlings were transplanted to soil. On day 22, three whole roots were collected from plants on plates and the tissues were frozen. The entire inflorescence and one leaf from three plants on soil were collected for each line between days 27 and 31. The tissues were collected in 2 mL tubes and powdered with a metal bead using a Retsch MM400 shaker after freezing the samples in liquid nitrogen. RNA was purified using an Illustra RNAspin Mini Kit (GE Healthcare). 1 μg of extracted RNA was then used with the iScript cDNA synthesis kit (BIO-RAD). qPCR was performed using the iQ SYBR Green Supermix (BIO-RAD). PP2AA3 was used as a reference gene and the primers for PP2AA3 and RUBY can be found in Supplementary Table S1. The standard curves were established using a pool of all the cDNAs. T2 RUBY seedlings were run with three biological replicates per line while Col-0 seedlings were run as four technical replicates. T3 tissues were run with three biological replicates per tissue and two technical replicates per line. The qPCR was performed on a C1000 Thermal Cycler (BIO-RAD) and the result was read using the Bio-Rad CFX Maestro software and analysed using standard methods (Pfaffl, Reference Pfaffl2001).

Acknowledgements

We thank Wesley George, Cassandra Maranas, Dr. Román Ramos Báez, Dr. Sarah Guiziou and Dr. Alexander Leydon for careful reading of the manuscript, as well as other members of the Nemhauser, Imaizumi and Steinbrenner labs for their feedback on this project. We thank Dr. Nicholas J. Provart for his help with the RNA-seq datasets.

Funding statement

This work was supported by the National Institute of Health (R01-GM107084), the National Science Foundation (IOS-1546873) and a Faculty Scholar Award from the Howard Hughes Medical Institute.

Competing interest

The authors declare no competing interests.

Author contribution

Experimental design and analysis: E.J.Y.Y., J.L.N. Research: E.J.Y.Y. Writing: E.J.Y.Y., J.L.N.

Data availability statement

The codes and datasets used in this study can be found on GitHub at https://github.com/Nemhauserlab/StablePromoters, and on Zenodo with DOI: 10.5281/zenodo.8170303. The repositories contain all the raw data as well as scripts to annotate, normalise and generate the figures used in the article. Datasets before annotation and annotated data before normalisation are both available. To minimise the supplemental file size, scripts without the datasets can be found in Supplementary Data S12.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/qpb.2023.10.

References

Afgan, E., Baker, D., van den Beek, M., Blankenberg, D., Bouvier, D., Čech, M., Chilton, J., Clements, D., Coraor, N., Eberhard, C., Grüning, B., Guerler, A., Hillman-Jackson, J., Von Kuster, G., Rasche, E., Soranzo, N., Turaga, N., Taylor, J., Nekrutenko, A., & Goecks, J. (2016). The galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2016 update. Nucleic Acids Research, 44(Web Server issue), W3W10. https://doi.org/10.1093/nar/gkw343 CrossRefGoogle ScholarPubMed
Ali, S., & Kim, W.-C. (2019). A fruitful decade using synthetic promoters in the improvement of transgenic plants. Frontiers in Plant Science, 10, 1433. https://doi.org/10.3389/fpls.2019.01433 CrossRefGoogle ScholarPubMed
Andreou, A. I., Nirkko, J., Ochoa-Villarreal, M., & Nakayama, N. (2021). Mobius assembly for plant systems highlights promoter–terminator interaction in gene regulation (p. 2021.03.31.437819). bioRxiv. https://doi.org/10.1101/2021.03.31.437819 CrossRefGoogle Scholar
Bak, A., & Emerson, J. B. (2020). Cauliflower mosaic virus (CaMV) biology, management, and relevance to GM plant detection for sustainable organic agriculture. Frontiers in Sustainable Food Systems, 4, 21. https://www.frontiersin.org/articles/10.3389/fsufs.2020.00021 CrossRefGoogle Scholar
Batista-Silva, W., da Fonseca-Pereira, P., Martins, A. O., Zsögön, A., Nunes-Nesi, A., & Araújo, W. L. (2020). Engineering improved photosynthesis in the era of synthetic biology. Plant Communications, 1(2), 100032. https://doi.org/10.1016/j.xplc.2020.100032 CrossRefGoogle ScholarPubMed
Belcher, M. S., Vuu, K. M., Zhou, A., Mansoori, N., Agosto Ramos, A., Thompson, M. G., Scheller, H. V., Loqué, D., & Shih, P. M. (2020). Design of orthogonal regulatory systems for modulating gene expression in plants. Nature Chemical Biology, 16(8), 857865. https://doi.org/10.1038/s41589-020-0547-4 CrossRefGoogle ScholarPubMed
Brophy, J. A. N., Magallon, K. J., Duan, L., Zhong, V., Ramachandran, P., Kniazev, K., & Dinneny, J. R. (2022). Synthetic genetic circuits as a means of reprogramming plant roots. Science, 377(6607), 747751. https://doi.org/10.1126/science.abo4326 CrossRefGoogle ScholarPubMed
Brückner, K., Schäfer, P., Weber, E., Grützner, R., Marillonnet, S., & Tissier, A. (2015). A library of synthetic transcription activator-like effector-activated promoters for coordinated orthogonal gene expression in plants. The Plant Journal, 82(4), 707716. https://doi.org/10.1111/tpj.12843 CrossRefGoogle ScholarPubMed
Cai, Y.-M., Kallam, K., Tidd, H., Gendarini, G., Salzman, A., & Patron, N. J. (2020). Rational design of minimal synthetic promoters for plants. Nucleic Acids Research, 48(21), 1184511856. https://doi.org/10.1093/nar/gkaa682 CrossRefGoogle ScholarPubMed
Clough, S. J., & Bent, A. F. (1998). Floral dip: A simplified method for agrobacterium-mediated transformation of Arabidopsis thaliana . The Plant Journal, 16(6), 735743. https://doi.org/10.1046/j.1365-313x.1998.00343.x CrossRefGoogle Scholar
Czechowski, T., Stitt, M., Altmann, T., Udvardi, M. K., & Scheible, W.-R. (2005). Genome-wide identification and testing of superior reference genes for transcript normalization in Arabidopsis. Plant Physiology, 139(1), 517. https://doi.org/10.1104/pp.105.063743 CrossRefGoogle ScholarPubMed
De Wilde, C., Van Houdt, H., De Buck, S., Angenon, G., De Jaeger, G., & Depicker, A. (2000). Plants as bioreactors for protein production: Avoiding the problem of transgene silencing. Plant Molecular Biology, 43(2), 347359. https://doi.org/10.1023/A:1006464304199 CrossRefGoogle ScholarPubMed
Engler, C., Youles, M., Gruetzner, R., Ehnert, T.-M., Werner, S., Jones, J. D. G., Patron, N. J., & Marillonnet, S. (2014). A Golden Gate modular cloning toolbox for plants. ACS Synthetic Biology, 3(11), 839843. https://doi.org/10.1021/sb4001504 CrossRefGoogle ScholarPubMed
Galli, M., Feng, F., & Gallavotti, A. (2020). Mapping regulatory determinants in plants. Frontiers in Genetics, 11, 591194. https://doi.org/10.3389/fgene.2020.591194 CrossRefGoogle ScholarPubMed
Gander, M. W., Vrana, J. D., Voje, W. E., Carothers, J. M., & Klavins, E. (2017). Digital logic circuits in yeast with CRISPR-dCas9 NOR gates. Nature Communications, 8(1), 1. https://doi.org/10.1038/ncomms15459 CrossRefGoogle ScholarPubMed
Gibson, D. G., Young, L., Chuang, R.-Y., Venter, J. C., Hutchison, C. A., & Smith, H. O. (2009). Enzymatic assembly of DNA molecules up to several hundred kilobases. Nature Methods, 6(5), 5. https://doi.org/10.1038/nmeth.1318 CrossRefGoogle ScholarPubMed
Guiziou, S., Sauveplane, V., Chang, H.-J., Clerté, C., Declerck, N., Jules, M., & Bonnet, J. (2016). A part toolbox to tune genetic expression in Bacillus subtilis . Nucleic Acids Research, 44(15), 74957508. https://doi.org/10.1093/nar/gkw624 Google ScholarPubMed
He, Y., Zhang, T., Sun, H., Zhan, H., & Zhao, Y. (2020). A reporter for noninvasively monitoring gene expression and plant transformation. Horticulture Research, 7(1), 1. https://doi.org/10.1038/s41438-020-00390-1 CrossRefGoogle ScholarPubMed
Huang, X., Li, S., & Zhan, A. (2019). Genome-wide identification and evaluation of new reference genes for gene expression analysis under temperature and salinity stresses in Ciona savignyi. Frontiers in Genetics, 10, 71. https://doi.org/10.3389/fgene.2019.00071 CrossRefGoogle ScholarPubMed
Jensen, M. K. (2018). Design principles for nuclease-deficient CRISPR-based transcriptional regulators. FEMS Yeast Research, 18(4), foy039. https://doi.org/10.1093/femsyr/foy039 CrossRefGoogle ScholarPubMed
Jiang, P., Zhang, K., Ding, Z., He, Q., Li, W., Zhu, S., Cheng, W., Zhang, K., & Li, K. (2018). Characterization of a strong and constitutive promoter from the Arabidopsis serine carboxypeptidase-like gene AtSCPL30 as a potential tool for crop transgenic breeding. BMC Biotechnology, 18(1), 59. https://doi.org/10.1186/s12896-018-0470-x CrossRefGoogle Scholar
Kassaw, T. K., Donayre-Torres, A. J., Antunes, M. S., Morey, K. J., & Medford, J. I. (2018). Engineering synthetic regulatory circuits in plants. Plant Science, 273, 1322. https://doi.org/10.1016/j.plantsci.2018.04.005 CrossRefGoogle ScholarPubMed
Khakhar, A., Leydon, A. R., Lemmex, A. C., Klavins, E., & Nemhauser, J. L. (2018). Synthetic hormone-responsive transcription factors can monitor and re-program plant development. eLife, 7, e34702. https://doi.org/10.7554/eLife.34702 CrossRefGoogle ScholarPubMed
Klepikova, A. V., Logacheva, M. D., Dmitriev, S. E., & Penin, A. A., (2015). RNA-seq analysis of an apical meristem time series reveals a critical point in Arabidopsis thaliana flower initiation. BMC Genomics, 16(1), 466. https://doi.org/10.1186/s12864-015-1688-9 CrossRefGoogle ScholarPubMed
Klepikova, A. V., Kasianov, A. S., Gerasimov, E. S., Logacheva, M. D., & Penin, A. A. (2016). A high resolution map of the Arabidopsis thaliana developmental transcriptome based on RNA-seq profiling. The Plant Journal, 88(6), 10581070. https://doi.org/10.1111/tpj.13312 CrossRefGoogle ScholarPubMed
Li, S., Shui, K., Zhang, Y., Lv, Y., Deng, W., Ullah, S., Zhang, L., & Xue, Y. (2017). CGDB: A database of circadian genes in eukaryotes. Nucleic Acids Research, 45(Database issue), D397D403. https://doi.org/10.1093/nar/gkw1028 Google ScholarPubMed
Lim, S. D., Mayer, J. A., Yim, W. C., & Cushman, J. C. (2020). Plant tissue succulence engineering improves water-use efficiency, water-deficit stress attenuation and salinity tolerance in Arabidopsis. The Plant Journal, 103(3), 10491072. https://doi.org/10.1111/tpj.14783 CrossRefGoogle ScholarPubMed
Liu, W., & Stewart, C. N. (2016). Plant synthetic promoters and transcription factors. Current Opinion in Biotechnology, 37, 3644. https://doi.org/10.1016/j.copbio.2015.10.001 CrossRefGoogle ScholarPubMed
Loraine, A. E., McCormick, S., Estrada, A., Patel, K., & Qin, P. (2013). RNA-Seq of Arabidopsis pollen uncovers novel transcription and alternative Splicing1[C][W][OA]. Plant Physiology, 162(2), 10921109. https://doi.org/10.1104/pp.112.211441 CrossRefGoogle Scholar
Love, M. I., Huber, W., & Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology, 15(12), 550. https://doi.org/10.1186/s13059-014-0550-8 CrossRefGoogle ScholarPubMed
Lu, Z., Marand, A. P., Ricci, W. A., Ethridge, C. L., Zhang, X., & Schmitz, R. J. (2019). The prevalence, evolution and chromatin signatures of plant regulatory elements. Nature Plants, 5(12), 12. https://doi.org/10.1038/s41477-019-0548-z CrossRefGoogle ScholarPubMed
Marquez, Y., Brown, J. W. S., Simpson, C., Barta, A., & Kalyna, M. (2012). Transcriptome survey reveals increased complexity of the alternative splicing landscape in Arabidopsis. Genome Research, 22(6), 11841195. https://doi.org/10.1101/gr.134106.111 CrossRefGoogle ScholarPubMed
Moreno-Giménez, E., Selma, S., Calvache, C., & Orzáez, D. (2022). GB_SynP: A modular dCas9-regulated synthetic promoter collection for fine-tuned recombinant gene expression in plants (p. 2022.04.28.489949). bioRxiv. https://doi.org/10.1101/2022.04.28.489949 CrossRefGoogle Scholar
Orr, D. J., Pereira, A. M., da Fonseca Pereira, P., Pereira-Lima, Í. A., Zsögön, A., & Araújo, W. L. (2017). Engineering photosynthesis: Progress and perspectives. F1000Research, 6, 1891. https://doi.org/10.12688/f1000research.12181.1 CrossRefGoogle Scholar
Pan, C., Sretenovic, S., & Qi, Y. (2021). CRISPR/dCas-mediated transcriptional and epigenetic regulation in plants. Current Opinion in Plant Biology, 60, 101980. https://doi.org/10.1016/j.pbi.2020.101980 CrossRefGoogle ScholarPubMed
Park, S.-Y., Peterson, F. C., Mosquna, A., Yao, J., Volkman, B. F., & Cutler, S. R. (2015). Agrochemical control of plant water use using engineered abscisic acid receptors. Nature, 520(7548), 7548. https://doi.org/10.1038/nature14123 CrossRefGoogle ScholarPubMed
Patron, N. J. (2020). Beyond natural: Synthetic expansions of botanical form and function. New Phytologist, 227(2), 295310. https://doi.org/10.1111/nph.16562 CrossRefGoogle ScholarPubMed
Peremarti, A., Twyman, R. M., Gómez-Galera, S., Naqvi, S., Farré, G., Sabalza, M., Miralpeix, B., Dashevskaya, S., Yuan, D., Ramessar, K., Christou, P., Zhu, C., Bassie, L., & Capell, T. (2010). Promoter diversity in multigene transformation. Plant Molecular Biology, 73(4), 363378. https://doi.org/10.1007/s11103-010-9628-1 CrossRefGoogle ScholarPubMed
Pfaffl, M. W. (2001). A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Research, 29(9), e45.CrossRefGoogle ScholarPubMed
R Core Team. (2022). R: A language and environment for statistical computing. R Foundation for Statistical Computing. https://www.R-project.org/ Google Scholar
Rajeev Kumar, S., Anunanthini, P., & Ramalingam, S. (2015). Epigenetic silencing in transgenic plants. Frontiers in Plant Science, 6, 693. https://doi.org/10.3389/fpls.2015.00693 Google Scholar
Rose, A. B. (2019). Introns as gene regulators: A brick on the accelerator. Frontiers in Genetics, 9, 672. https://www.frontiersin.org/articles/10.3389/fgene.2018.00672 CrossRefGoogle ScholarPubMed
Rose, A. B., Elfersi, T., Parra, G., & Korf, I. (2008). Promoter-proximal introns in Arabidopsis thaliana are enriched in dispersed signals that elevate gene expression. The Plant Cell, 20(3), 543551. https://doi.org/10.1105/tpc.107.057190 CrossRefGoogle ScholarPubMed
Schindelin, J., Arganda-Carreras, I., Frise, E., Kaynig, V., Longair, M., Pietzsch, T., Preibisch, S., Rueden, C., Saalfeld, S., Schmid, B., Tinevez, J.-Y., White, D. J., Hartenstein, V., Eliceiri, K., Tomancak, P., & Cardona, A. (2012). Fiji: An open-source platform for biological-image analysis. Nature Methods, 9(7), 7. https://doi.org/10.1038/nmeth.2019 CrossRefGoogle ScholarPubMed
Signorell, A., Aho, K., Alfons, A., Anderegg, N., & Aragon, T. (2022). DescTools: Tools for descriptive statistics. https://cran.r-project.org/package=DescTools Google Scholar
South, P. F., Cavanagh, A. P., Liu, H. W., & Ort, D. R. (2019). Synthetic glycolate metabolism pathways stimulate crop growth and productivity in the field. Science, 363(6422), eaat9077. https://doi.org/10.1126/science.aat9077 CrossRefGoogle ScholarPubMed
Tian, F., Yang, D.-C., Meng, Y.-Q., Jin, J., & Gao, G. (2020). PlantRegMap: Charting functional regulatory maps in plants. Nucleic Acids Research, 48(D1), D1104D1113. https://doi.org/10.1093/nar/gkz1020 Google ScholarPubMed
Tokizawa, M., Kusunoki, K., Koyama, H., Kurotani, A., Sakurai, T., Suzuki, Y., Sakamoto, T., Kurata, T., & Yamamoto, Y. Y. (2017). Identification of Arabidopsis genic and non-genic promoters by paired-end sequencing of TSS tags. The Plant Journal: For Cell and Molecular Biology, 90(3), 587605. https://doi.org/10.1111/tpj.13511 CrossRefGoogle ScholarPubMed
Wang, P.-H., Kumar, S., Zeng, J., McEwan, R., Wright, T. R., & Gupta, M. (2020). Transcription terminator-mediated enhancement in transgene expression in maize: Preponderance of the AUGAAU motif overlapping with poly(a) signals. Frontiers in Plant Science, 11, 570778. https://doi.org/10.3389/fpls.2020.570778 CrossRefGoogle ScholarPubMed
Wang, Z., Lyu, Z., Pan, L., Zeng, G., & Randhawa, P. (2019). Defining housekeeping genes suitable for RNA-seq analysis of the human allograft kidney biopsy tissue. BMC Medical Genomics, 12(1), 86. https://doi.org/10.1186/s12920-019-0538-z CrossRefGoogle ScholarPubMed
Weber, E., Engler, C., Gruetzner, R., Werner, S., & Marillonnet, S. (2011). A modular cloning system for standardized assembly of multigene constructs. PLoS One, 6(2), e16765. https://doi.org/10.1371/journal.pone.0016765 CrossRefGoogle ScholarPubMed
Win, J., & Kamoun, S. (2004). pCB301-p19: A Binary Plasmid Vector to Enhance Transient Expression of Transgenes by Agroinfiltration. http://www.KamounLab.net.Google Scholar
Yamamoto, Y. Y., Yoshioka, Y., Hyakumachi, M., & Obokata, J. (2011). Characteristics of core promoter types with respect to gene structure and expression in Arabidopsis thaliana . DNA Research: An International Journal for Rapid Publication of Reports on Genes and Genomes, 18(5), 333342. https://doi.org/10.1093/dnares/dsr020 CrossRefGoogle ScholarPubMed
Zhang, T., Gao, Y., Wang, R., & Zhao, Y. (2017). Production of guide RNAs in vitro and in vivo for CRISPR using ribozymes and RNA polymerase II promoters. Bio-Protocol, 7(4), e2148. https://doi.org/10.21769/BioProtoc.2148 CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. (a) Pipeline to identify constitutive promoters. The number of genes that pass each filter is indicated, along with the software used to implement the analysis. SRA is the ‘Sequence Read Archive’. Detailed methods, including parameters for each filter, are described in Section 4. (b) Schematic of filters used to select candidate promoters to engineer with synthetic gRNA target sites. (c) Schematic describing how we defined ‘promoter’ and ‘terminator’. The ‘promoter’ was defined here as starting from the transcription start site and going upstream to a maximum of 2,000 bp or to the next annotated neighbouring gene, whichever is shorter. Similarly, a ‘terminator’ was defined as starting from the transcription end site and going downstream a maximum of 250 bp or to the next annotated neighbouring gene, whichever is shorter. Promoters and terminators were cloned, along with their respective UTRs, following the Golden Gate MoClo system. (d) Plot showing values for the 10,096 genes expressed in all tissues. The geometric mean of expression across samples is plotted on the x-axis with the coefficient of variation (CV) on the y-axis. Both axes are on a base-10 log scale. The lowest 3% CV corresponds to a 0.26 CV cutoff, and the 303 genes with CV lower than 0.26 are highlighted in yellow. The final 33 candidates that fulfilled all criteria are highlighted in red. Several common promoters used in plant synthetic biology are annotated for reference.

Figure 1

Figure 2. We identified 16 promoters that expressed in N. benthamiana. Six promoters were modified to introduce gRNA target sites. These sites are designated by brackets following the gene name. Three different constructs were injected per leaf, each containing a promoter to be tested driving NLS_YFP and an internal control of pUBQ10:NLS_mTURQ. Each leaf also has a negative control injection that only contains pUBQ10:NLS_mTURQ. Normalisation is performed using the formula: $\frac{{\mathrm{YFP}}_{\mathrm{promoter}}-\mathrm{median}\left({\mathrm{YFP}}_{\mathrm{Neg}.\mathrm{control}}\right)}{{\mathrm{mTURQ}}_{\mathrm{promoter}}}$. For each construct, the three replicates with median fluorescence levels closest to the median of the group were selected for visualisation and statistical analysis. Each biological replicate is represented by a beeswarm plot of 24 datapoints (12 per leaf disc, 2 disc per injection) collected from the plate reader as well as a single summarising datapoint representing the median. The boxplots represent all biological replicates. Significance test was performed using Dunnett’s test for comparing multiple treatments with control at 95% family wise confidence level. Non-significant constructs are marked as NS. For a given construct, the colours signify datapoints derived from the same biological replicate.

Figure 2

Figure 3. (a) Three promoters showed expression of RUBY in Arabidopsis T2 plants. The flowers, siliques, and leaves were imaged on day 34, while the seedling images were imaged on day 12. The inset boxes are the same images at higher magnification. Red arrows indicate areas where RUBY expression is visible by the eye. (b) qPCR data on T2 whole seedlings in three biological replicates for each line and (c) qPCR data on tissues collected from T3 plants, each with three biological replicates and two technical replicates. RUBY expression was normalised against the reference gene PP2AA3, and the bars represent the mean expression of the RUBY reporter and the standard error of the mean (SEM).

Figure 3

Figure 4. (a) The four constructs co-injected for each injection. The injection always contains the mPromoter, dCas9-guided repressor, and the two self-cleaving input gRNAs. The gRNAs are denoted with X and Y representing a variable input. (b) Schematic of the NOR gate when both input gRNAs are present. (c) Pattern of injection for the four possible input combinations and the gRNAs used for each injection. (1,1) represents both guides are present while (0,0) represents neither is present. When a guide is not present, a non-matching gRNA is injected in its place, denoted here as gRNA_M or N. (d) Five of the six mPromoters functioned as NOR gates. All guides apart from gRNA_F are independently repressible. Each biological replicate is represented by a beeswarm plot of 24 datapoints (12 per leaf disc, 2 disc per injection) collected from the plate reader as well as a single summarising datapoint representing the median. The boxplots represent all biological replicates. The signal is measured as the YFP fluorescence (driven by the promoters being tested) divided by the mTURQ fluorescence (driven by pUBQ10). In each set of NOR gate injections, the (0,0) injection serves as the unrepressed control, and the dataset is normalised by dividing all values by the median of the unrepressed control on a per-leaf basis. The y-axis represents fold changes from the unrepressed control and each biological replicate of the control is centred on 1. Each colour represents a unique leaf. Letters above each boxplot are Compact Letter Display (CLD) for all pairwise comparisons within each set of injections using ANOVA followed by Tukey’s Honest Significant Difference Test. Numbers above the boxplot represent fold repression between the (0,0) and (1,1) injections.

Supplementary material: File

Yang and Nemhauser supplementary material 1
Download undefined(File)
File 3.4 MB
Supplementary material: File

Yang and Nemhauser supplementary material 2
Download undefined(File)
File 67.9 KB
Supplementary material: File

Yang and Nemhauser supplementary material 3
Download undefined(File)
File 25.5 KB
Supplementary material: File

Yang and Nemhauser supplementary material 4
Download undefined(File)
File 29 KB
Supplementary material: File

Yang and Nemhauser supplementary material 5
Download undefined(File)
File 237.7 KB
Supplementary material: File

Yang and Nemhauser supplementary material 6
Download undefined(File)
File 690.5 KB

Author comment: Building a pipeline to identify and engineer constitutive and repressible promoters — R0/PR1

Comments

13 October 2022

Dr. Olivier Hamant

Editor-in-Chief

Quantitative Plant Biology

Dear Dr. Hamant:

Attached please find our manuscript entitled “Expanding the synthetic biology toolbox with a library of constitutive and repressible promoters”, which we are submitting for consideration as an article. None of the material has been published or is under consideration elsewhere. We have submitted a preprint: MS ID#: BIORXIV/2022/511673.

For the ambitious engineering projects currently under development in plants, one of the major bottlenecks is the lack of well-characterized, constitutive promoters. Reusing promoter parts in large genetic constructs makes cloning challenging and increases the likelihood of transgene silencing, among other concerns. In our manuscript, we describe the development and implementation of a pipeline to identify broadly expressed genes. We then cloned out presumptive promoters and terminators of these loci, and validated their expression in both transient (Nicotiana benthamiana) and stable (Arabidopsis thaliana) transformation assays. To further increase the functionality for building complex circuits, we engineered a subset of these promoters with unique gRNA target sites and successfully generated orthogonally repressible NOR gates. The constitutive promoters screened in this study can help meet the need for additional promoter parts, and by converting them into NOR gates, they form the basis of constructing more complex logic gates in the future.

Based on the reception of this work by our colleagues in workshops and meetings, we believe it will appeal to a diverse range of scientists interested in transcriptional regulation, networks and synthetic biology. It would be fantastic if the debut of this toolset could be in an open-access journal like Quantitative Plant Biology.

Some potential reviewers for this work include:

Jenn Brophy, Stanford, jbrophy@stanford.edu, expertise: plant synthetic biology, esp. logic gates

Kevin Cox, Danforth Center, KCox@danforthcenter.org, expertise: transcriptional regulation

David Ehrhardt, Carnegie Institute, dehrhardt@carnegiescience.edu, expertise: vector and tool development

Naomi Nakayama, Imperial College London, n.nakayama@imperial.ac.uk, expertise: plant synthetic biology; transcriptional regulation, esp. promoters and terminators

Diego Orsáez, IBMCP, CSIC Valencia, dorzaez@ibmcp.upv.es, expertise: plant synthetic biology; dCas9-based tool development

Nicola Patron, Earlham Institute, nicola.patron@earlham.ac.uk, expertise: plant synthetic biology; transcriptional regulation, esp. promoters

Ross Sozzani, NCSU, rsozzan@ncsu.edu, expertise: vector and tool development; bioinformatics

Yoshiharu Y. Yamamoto, Gifu University, yyy@gifu-u.ac.jp, expertise: transcriptional regulation, esp. promoters

Thank you for your time and consideration.

Sincerely,

Dr. Jennifer Nemhauser, jn7@uw.edu

Department of Biology

University of Washington

Review: Building a pipeline to identify and engineer constitutive and repressible promoters — R0/PR2

Conflict of interest statement

Reviewer declares none. I have no conflict of interest in the research field of the reviewed article.

Comments

In this report, the authors selected constitutively and stably expressed genes using public gene expression database of Arabidopsis, and their promoters and terminators, totally sixteen, were subjected to transient expression analysis in N. benthamiana. Among them, six sets were selected, gRNA tartet sites were introduced in order to repress their gene expression by expression of gRNA genes. Suppression by gRNA was analyzed in transient expression assays of N. benthamiana. Expression of four promoter sets were confirmed in transgenic Arabidopsis as well.

Because of the shown achievements are less than ones required for one article, I don’t recommend this report to be published on Quantitative Plant Biology.

Major points

1) Title is too big to express the contents. Development of an Expression and Repression System for Transgenic Plants, for example, would be more appropriate.

2) Function of gene repression would be the key of the report. However, the demonstrated repression in Fig. 4D is much less effective than expected. This is not like repression but small modulation. The authors should try to develop ten fold repression or more. Otherwise, I don’t think of beneficial applications of the system.

Minor points

1) Fig. 2 and 3. A positive control (e.g. 35S promoter - NOS terminator) should be included to see expression levels of the developed promoter-terminator pairs.

2) Fig. 4D. Degree of repression should be shown. This is important information, and some researchers are really interested in it.

Review: Building a pipeline to identify and engineer constitutive and repressible promoters — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

The manuscript by Yang and Nemhauser describes first a pipeline for identifying natural promoters with stable expression in plants using transcriptomic databases, followed by the modification of a subset of these promoters with gRNA target sequences that can be repressed by a Cas9-based repressor and used to build NOR logic gates.

The development of additional tools to control gene expression in plants fills an important need of plant synthetic biology. The authors took a different approach to develop these promoters by choosing endogenous plant promoters, rather than developing synthetic ones, and although natural promoters may not function in a completely orthogonal manner, they do have some advantages, as pointed out in the manuscript.

Sixteen selected promoters showed expression that is significantly different from the negative control in transient expression assays in N. benthamiana. However, the choice of reporter (RUBY) prevented a proper assessment on the constitutive activity of these promoters in stable transgenic Arabidopsis plants. Although these promoters appear to have constitutive activity based on their expression patterns from transcriptomics data, it is not clear from the data shown in this paper that the 2 kb of upstream sequence (promoter) is capable of conferring constitutive activity in transgenic plants. For this reason, I suggest that the authors remove any statements that these promoters are constitutive, unless experiments are done in transgenic plants where a different (perhaps more sensitive) reporter gene is used (e.g., GUS) and gene expression is assessed in different tissues throughout the plant, at different developmental stages. Alternatively, RT-qPCR could be done on mRNA from individual tissues from the transgenic plants, at different developmental stages, to show constitutive activity.

The modifications introduced into some of the promoters to make them repressible by a Cas9-based repressor, in response to two different gRNAs as inputs, is nicely demonstrated in transient assays in N. benthamiana. However, given the large variability observed in these transient assays, it is not clear that these NOR logic gates would function robustly in stable transgenic plants. I suggest this limitation be discussed.

The data are clearly presented, and the experiments have a reasonable number of replicates and proper controls.

Minor suggestions are included below:

Line 90: in the legend for Figure 1, define SRA as “Sequence Read Archive”.

Lines 144-146: Twelve promoters were selected to transform Arabidopsis, yet one of those did not show activity in N. benthamiana? Why was this one promoter tested in Arabidopsis? This is not clear.

Line 173: I could not find any instances of panels A and C from Figure 4 being mentioned in the text.

Line 190: provide full gene number – AT1G64550(E,A)

Line 199: fluorescence

Line 217: As in my comments above, the reference to these promoters being constitutive should be removed.

Lines 323-324: I don’t believe this sentence is referencing the correct Supplementary file numbers.

Line 329: …into regions that do not disrupt…

Line 345: Azure C600 Western Blot

Lines 387-388: Were these seeds plated on media without selection? T2 plants should still be segregating for the transgenes, so why were these grown without selection? How do you know that a plant not expressing RUBY is not because it doesn’t contain the transgene?

Review: Building a pipeline to identify and engineer constitutive and repressible promoters — R0/PR4

Conflict of interest statement

Reviewer declares none

Comments

This manuscript describes the study undertaken by Yang and Nemhauser to expand the repertoire of constitutive promoters to be used to drive transgene expression in plants. This is a valuable effort as more promoters are needed for synthetic biology/metabolic engineering approaches, as indicated by the authors. The strategy presented here relies on the identification of endogenous Arabidopsis thaliana promoters and their analysis in plants through transient and stable expression assays. The authors enrich the analysis/manipulation by the valuable addition of two crispR targets site to repress RNA transcription, and thus tune down transgene expression in future applications (according to a NOR logic gate).

Overall, the manuscript is well written and clear to follow by a broad readership. The most salient claims are justified by the experimental data. In some instances, however, alternative explanations could be proposed/explored. I list the following comments/suggestions for improvement:

Main considerations

The endogenous nature of the selected sequences prompts a question about their conservation, and thus transferability, to other species. The authors do address this when they test the promoters in transient tobacco agroinfiltration, however there is potential information being missed by such an approach. I wonder whether this could be addressed with an in silico approach, at least for the 5-6 promoters which are validated experimentally: for example can synteny of conserved transcrtiption binding sites be compared across Arabidopsis accessions and other brassica species (for which the genome sequence is available). Similarly, the position where to insert gRNA targeted-sequences could be inspected for variability (with the assumption that high variability implies negligible regulatory capacity).

Ruby usefulness is questioned, based on the results showed in fig 3. However, the authors do not consider potential post-transcriptional regulation that the UTRs included in their design might entail. A comparison between qPCR and protein levels (if any antibody is available to test it) would be very informative.

Figure 2 shows that the selected promoters/terminators regulate genes with a range of expression levels. How does this compare between the endogenous genes and transgenes (analysed in figure 3)?

Finally, the constitutive nature of the promoters is only evaluated at the spatial level (and to a lesser extent, at the developmental one), but not their resilience to alteration by environmental stimuli (e.g. light/temperature). This could be tested in the T2 lines.

Specific comments in the text:

Introduction

The second paragraph sets out the goal of this study and provides a definition for constitutive promoters. The term ubiquitous could be presented as well, as a tissue specific promoter could still be constitutively active at most/all developmental stages, but I appreciate that there is ambiguity in the literature about these terms.

The third paragraphs describes the previous approaches to generate constitutive promoters. By the authors own definition of ‘constitutive’, promoters with different expression patterns could not be defined as such. Similarly, the use of exogenous ZF, TALEN or Cas binding sites does not seem to address the problem, as their effectiveness would depend on the abundance of the trans-acting factor, moving the issue of regulation one level up.

Results

Can the authors justify why limiting the analysis of transcription factor binding sites to 500 bp? Is regulation potential expected to decrease with distance from TSS?

Figure 2 legend: from how many leaves does the beeswarm plot of datapoint collected from the platereader?

Figure 3: the text (l155) mentions expression in the pollen but the resolution of the photo does not allow to see that (at least, not in the version available for review). More in general Ruby production is assessed locally, but it is unclear which tissue was used for RNA extraction before the qPCR analysis. I believe it would be interesting to repeat these qPCR in (at least some) different tissues, separately, and to compare transgene expression with the Ruby phenotype. It would be interesting to compare this information with transgenic lines where Ruby is driven by traditional constitutive promoters, if such transgenic plants are available.

Figure 4. The legend states that colours represent datapoints from a single leaf (if I understand correctly). It seems to be a lot of datapoints from a single leaf, while in the analysis showed in figure 2 comes from leaves which were infiltrated in only 4 spots.

Discussion:

In the introduction, the authors correctly indicate the need to identify new minimal promoters, to design new full promoters that respond to specific stimuli. They could propose that such an analysis should be undertaken in the future on the promoters they identified in this study.

Materials and methods:

A description of how the gRNA target sequences where introduced in the cloned promoters is missing.

Recommendation: Building a pipeline to identify and engineer constitutive and repressible promoters — R0/PR5

Comments

Your manuscript has been now assessed by three reviewers, please apologize the delays. I share with them the opinion that your manuscript “Expanding the synthetic biology toolbox with a library of constitutive and repressible promoters“ is very interesting, timely and potentially relevant for the community. However, several valid concerns have been raised. Among other points, and in particular, the reviewers identified that the choice of RUBY as a reporter is not ideal, including the limitations shown in terms of sensitivity and that the dynamic range is relatively low. In addition, the experimental data doesn´t seem to support the claim of constitutive activation for the promoter. Further comparisons and in silico studies are suggested.

I would therefore be glad to consider publication of your article provided that it undergoes a revision tackling the issues raised by the reviewers

Decision: Building a pipeline to identify and engineer constitutive and repressible promoters — R0/PR6

Comments

No accompanying comment.

Author comment: Building a pipeline to identify and engineer constitutive and repressible promoters — R1/PR7

Comments

21 May 2023

Dr. Olivier Hamant

Editor-in-Chief

Quantitative Plant Biology

Dear Dr. Hamant:

Attached please find our revised manuscript entitled “Building a pipeline to identify and engineer constitutive and repressible promoters”. We thank the editor and reviewers for their careful reading of our work. We have made a number of changes described below in response to the reviewer comments (including modifying the title), as well as adding new experimental results. We believe that these changes improve the manuscript and increase the accessibility of our findings. We hope that you agree.

Thank you for your time and consideration.

Sincerely,

Dr. Jennifer Nemhauser

Department of Biology

University of Washington

jn7@uw.edu

Detailed response to reviews:

Reviewer 1 (R1)

Major Comments:

1) Title is too big to express the contents. Development of an Expression and Repression System for Transgenic Plants, for example, would be more appropriate.

Our response: Inspired by the reviewer’s suggestion, we changed the title to: Building a pipeline to identify and engineer constitutive and repressible promoters. We hope that this change captures the spirit of the suggested revision.

(R1) 2) Function of gene repression would be the key of the report. However, the demonstrated repression in Fig. 4D is much less effective than expected. This is not like repression but small modulation. The authors should try to develop ten fold repression or more. Otherwise, I don’t think of beneficial applications of the system.

Our response: While we understand (and share) the reviewer’s frustration with the relatively subtle impact of the repression we observe, it is worth noting that these parts have not been optimized (e.g., guide sites have not been moved around within the promoter), and that the overall low level of expression makes strong repression quite difficult to achieve. We have added text throughout the manuscript to emphasize the proof-of-concept nature of this study, and its shortcomings. Despite these issues, we strongly believe that these results will be of use to the plant synthetic biology community, as well as those interested in transcriptional regulation more generally.

(R1) Minor Comments:

1) Fig. 2 and 3. A positive control (e.g. 35S promoter - NOS terminator) should be included to see expression levels of the developed promoter-terminator pairs.

Our response: We can appreciate the reviewer’s desire to have a familiar promoter to serve as a point of comparison; however, we think that the levels of a 35S promoter are far too high to serve as a meaningful control for these endogenous, low-expressing promoters.

(R1) 2) Fig. 4D. Degree of repression should be shown. This is important information, and some researchers are really interested in it.

Our response: Thank you for this suggestion. We have modified Figure 4D accordingly.

(R2) Reviewer: 2

Major comments:

Although these promoters appear to have constitutive activity based on their expression patterns from transcriptomics data, it is not clear from the data shown in this paper that the 2 kb of upstream sequence (promoter) is capable of conferring constitutive activity in transgenic plants. For this reason, I suggest that the authors remove any statements that these promoters are constitutive, unless experiments are done in transgenic plants where a different (perhaps more sensitive) reporter gene is used (e.g., GUS) and gene expression is assessed in different tissues throughout the plant, at different developmental stages. Alternatively, RT-qPCR could be done on mRNA from individual tissues from the transgenic plants, at different developmental stages, to show constitutive activity.

Our response. We thank the reviewer for improving the precision of our language. We have now changed language throughout the text to differentiate between the objective of our screen versus what we have observed with our promoters. Additional qPCR had also been performed as per reviewer 2 and reviewer 3’s suggestion to confirm the expression of RUBY in the various tissues collected for two of the promoters. (Figure 3C had been added).

(R2) (Given the large variations in Tobacco data)... It is not clear that these NOR logic gates would function robustly in stable transgenic plants. I suggest this limitation be discussed.

Our response: We have added text to the Discussion to incorporate this suggestion as follows:

“While N. benthamiana serves as a great prototyping platform, the performance of the gates would also need to be evaluated in stable Arabidopsis lines to validate their viability.”

(R2) Minor suggestions:

(R2) Line 90: in the legend for Figure 1, define SRA as “Sequence Read Archive”.

Our response: We thank the reviewer for the suggestion, and have made this edit.

(R2) Lines 144-146: Twelve promoters were selected to transform Arabidopsis, yet one of those did not show activity in N. benthamiana? Why was this one promoter tested in Arabidopsis? This is not clear.

Our response: We have added a clarifying sentence regarding the promoter in question: “Given that expression in N. benthamiana doesn’t perfectly predict expression in Arabidopsis, we included two promoters (AT1G54080, AT1G71860) that did not show expression in tobacco infiltration in our Arabidopsis stable transformation experiment. Interestingly, AT1G54080 displayed RUBY expression in roots and pollen.”

(R2) Line 173: I could not find any instances of panels A and C from Figure 4 being mentioned in the text.

Our response: We thank the reviewer for the suggestion, and have added callouts for these panels.

(R2) Line 190: provide full gene number – AT1G64550(E,A)

Our response: We have made this correction.

(R2) Line 199: fluorescence

Our response: We have made this correction.

(R2) Line 217: As in my comments above, the reference to these promoters being constitutive should be removed.

Our response: We thank the reviewer for the suggestion, and have made this edit.

(R2) Lines 323-324: I don’t believe this sentence is referencing the correct Supplementary file numbers.

Our response: We have made this correction.

(R2) Line 329: …into regions that do not disrupt…

Our response: We have made this correction.

(R2) Line 345: Azure C600 Western Blot

Our response: We have made this correction.

(R2) Lines 387-388: Were these seeds plated on media without selection? T2 plants should still be segregating for the transgenes, so why were these grown without selection? How do you know that a plant not expressing RUBY is not because it doesn’t contain the transgene?

Our response: The seedlings were screened without selection to minimize stress on the seedlings. To ensure that at least some proportion of our seedlings did contain the transgene, we examined at least 19 T2 individuals per T1-line (independent insertion event).

Reviewer: 3 (R3)

Main considerations

(R3) The endogenous nature of the selected sequences prompts a question about their conservation, and thus transferability, to other species. The authors do address this when they test the promoters in transient tobacco agroinfiltration, however there is potential information being missed by such an approach. I wonder whether this could be addressed with an in silico approach, at least for the 5-6 promoters which are validated experimentally: for example can synteny of conserved transcription binding sites be compared across Arabidopsis accessions and other brassica species (for which the genome sequence is available). Similarly, the position where to insert gRNA targeted-sequences could be inspected for variability (with the assumption that high variability implies negligible regulatory capacity).

Our response: We thank the reviewer for this suggestion. We are also interested in exploring this cross-species promoter analysis. However, we believe the question is best addressed using a broader bioinformatic approach including many additional species, which would be a significant amount of additional work and is outside the scope of the current manuscript.

(R3) Ruby usefulness is questioned, based on the results showed in fig 3. However, the authors do not consider potential post-transcriptional regulation that the UTRs included in their design might entail. A comparison between qPCR and protein levels (if any antibody is available to test it) would be very informative.

Our response: We thank the reviewer for this suggestion and have added the aspect of post-transcriptional regulation in the discussion section:

“Evaluating the promoters using RUBY revealed that the novel reporter had limited sensitivity when driven by weaker promoters. We were able to detect RUBY expression in seedlings and adult tissues without visible coloration using qPCR, a more sensitive assay. However, it is important to note that detecting transcripts doesn’t always imply comparable levels of protein production due to post-transcriptional and post-translational regulation. In our design, we attempted to capture the effects of any post-transcriptional regulation by including the UTRs, but other potential transcriptional regulators could be missed. The lower-than-expected RUBY mRNA levels detected could be due to such regulators. Promoter-proximal introns after the translation start codon, for example, would not be captured in the cloning pipeline though it is known to contribute to gene expression (Rose, 2019; Rose et al., 2008). Distally located regulatory regions would also not be captured, but they should be rare in the compact genome of Arabidopsis (Galli et al., 2020; Lu et al., 2019).”

In regards to protein level analysis, we appreciate the suggestion though we are unaware of an antibody for RUBY.

(R3) Figure 2 shows that the selected promoters/terminators regulate genes with a range of expression levels. How does this compare between the endogenous genes and transgenes (analysed in figure 3)?

Our response: We thank the reviewer for suggesting this analysis. We have added a brief discussion of this as well as an additional supplementary table addressing the comparison between endogenous gene and transgene levels.

“The expression level of RUBY mRNA detected through qPCR is weaker than expected from the RNAseq dataset. While the predicted expression level of all four of the genes in the qPCR experiment are higher than the reference gene PP2AA3, the measured result showed the opposite (Supplementary Table S6). This discrepancy could be attributed to the RUBY reporter or potential limitations in identifying additional transcriptional regulators (see discussion).”

(R3) Finally, the constitutive nature of the promoters is only evaluated at the spatial level (and to a lesser extent, at the developmental one), but not their resilience to alteration by environmental stimuli (e.g. light/temperature). This could be tested in the T2 lines.

Our response: We thank the reviewer for their suggestion. Our main design specification was to identify broadly-expressed promoters, and our selection criteria do not include stress-resistance, therefore we have only evaluated these promoters on their expression pattern. We did, however, include a metric for stress-resistance in our Supplementary_Table_S2 that was determined from RNAseq data.

Minor considerations:

(R3) The second paragraph sets out the goal of this study and provides a definition for constitutive promoters. The term ubiquitous could be presented as well, as a tissue specific promoter could still be constitutively active at most/all developmental stages, but I appreciate that there is ambiguity in the literature about these terms.

Our response: We appreciate the reviewer noting the ambiguity in the use of the term “constitutive.” We have modified our text to state explicitly our definition: “Constitutive promoters are defined here as expressed in all tissues at all times”

(R3) The third paragraphs describes the previous approaches to generate constitutive promoters. By the authors own definition of ‘constitutive’, promoters with different expression patterns could not be defined as such. Similarly, the use of exogenous ZF, TALEN or Cas binding sites does not seem to address the problem, as their effectiveness would depend on the abundance of the trans-acting factor, moving the issue of regulation one level up.

Our response: We have modified the text to reflect this suggestion: “To expand the number of promoters available, several groups have recently used distinct strategies to engineer both constitutively and conditionally expressed promoters”

Results

(R3) Can the authors justify why limiting the analysis of transcription factor binding sites to 500 bp? Is regulation potential expected to decrease with distance from TSS?

Our response: We have added a few references to clarify this point: “While there are no specific guidelines on optimal placement for gRNA target-sites in plants (Pan et al., 2021), studies in other eukaryotes have pointed to -50 to +300bp from TSS in mammalian cells for CRISPRi, and within -200bp from TSS in yeast (Jensen, 2018)”

(R3) Figure 2 legend: from how many leaves does the beeswarm plot of datapoint collected from the platereader?

Our response: We have added text to the figure legend and methods section to clarify this point.

“For each construct, the three replicates with median fluorescence levels closest to the median of the group were selected for visualization and statistical analysis. Each biological replicate is represented by a beeswarm plot of 24 datapoints (12 per leaf disc, 2 disc per injection) collected from the plate reader as well as a single summarizing datapoint representing the median. The boxplots represent all biological replicates.”

(R3) Figure 3: the text (l155) mentions expression in the pollen but the resolution of the photo does not allow to see that (at least, not in the version available for review). More in general Ruby production is assessed locally, but it is unclear which tissue was used for RNA extraction before the qPCR analysis. I believe it would be interesting to repeat these qPCR in (at least some) different tissues, separately, and to compare transgene expression with the Ruby phenotype. It would be interesting to compare this information with transgenic lines where Ruby is driven by traditional constitutive promoters, if such transgenic plants are available.

Our response: We have made a clarification in the figure legends that the qPCR data was collected from T2 whole seedlings. Additional qPCR had also been performed as per the reviewer 2 and reviewer 3’s suggestion to confirm the expression of RUBY in the various tissues collected for two of the promoters. (Figure 3C had been added).

(R3) Figure 4. The legend states that colours represent datapoints from a single leaf (if I understand correctly). It seems to be a lot of datapoints from a single leaf, while in the analysis showed in figure 2 comes from leaves which were infiltrated in only 4 spots.

Our response: We have added text to the figure legend and methods section to clarify this point.

Discussion:

(R3) In the introduction, the authors correctly indicate the need to identify new minimal promoters, to design new full promoters that respond to specific stimuli. They could propose that such an analysis should be undertaken in the future on the promoters they identified in this study.

Our response: We thank the reviewer for their suggestion and have modified our discussion section accordingly

“If a minimal promoter sequence can be identified from these native promoters, they can also serve as the foundation of additional synthetic promoters where the expression pattern and strength can be freely modified by adding cis-elements or synthetic transcription factor binding sites.”

Materials and methods:

(R3) A description of how the gRNA target sequences where introduced in the cloned promoters is missing.

Our response: We have added text to the methods section to clarify this point. “gRNA target-sites were cloned into regions that does not disrupt any predicted TF binding sites through Gibson assembly by replacing the original sequence (Gibson et al., 2009)”

Review: Building a pipeline to identify and engineer constitutive and repressible promoters — R1/PR8

Conflict of interest statement

None

Comments

This article reports a pipeline for identification of natural promoters with constitutive and ubiquitous expression and construction of artificial constructs using them. In addition, an additional repression system for the artificial constructs is reported.

Because presented data is limited and success of the development is not fully shown, I don’t recommend this article to be published at Quantitative Plant Biology.

Major issues

*I could not understand why the authors want to show “a pipeline” instead of successfully developed tools, which are functional promoter cassettes. After publication of the article, do the authors expect for other researchers to follow this scheme to get additional constitutive promoters ???

*Fig. 2 and 3: It is better to show fold change over negative control levels. Presented figures are difficult to guess strength of the signals obtained from positive samples.

*Fig. 3A: Signal of RUBY is not easy to see. Pictures need improvements.

*Fig. 4: Degree of suppression are all around 0.5 fold or more, which, for me, indicates failure of the suppression system. These examples do not look like a practical system.

*I could not find proof of constitutive function of the developed promoters in the figures. This information is necessary.

Review: Building a pipeline to identify and engineer constitutive and repressible promoters — R1/PR9

Conflict of interest statement

Reviewer declares none.

Comments

In the new version of the manuscript, the authors addressed all the points I previously raised.

Recommendation: Building a pipeline to identify and engineer constitutive and repressible promoters — R1/PR10

Comments

Your revised manuscript has now been evaluated by the two of the three previously involved reviewers. I apologise for the delays. Thanks for your efforts preparing a new version which is indeed improved. You have tackled most of the scientific issues. I am pleased to accept the manuscript for publication provided that you tackle the minor comments of Rev#1 on Figs 2 and 3. I´m glad to be able to bring you encouraging news and I look forward to receiving the revised manuscript.

Decision: Building a pipeline to identify and engineer constitutive and repressible promoters — R1/PR11

Comments

No accompanying comment.

Author comment: Building a pipeline to identify and engineer constitutive and repressible promoters — R2/PR12

Comments

23 July 2023

Dr. Olivier Hamant, Editor-in-Chief

Dr. Matias Zurbriggen, Associate Editor:

Quantitative Plant Biology

Dear Dr. Hamant and Dr. Zurbriggen:

Thank you for the wonderful news about the acceptance of our manuscript. We have made addressed the remaining reviewer concerns to the best of our ability, described in detail below.

Detailed response to reviews:

Reviewer 1 (R1)

1) I could not understand why the authors want to show “a pipeline” instead of successfully developed tools, which are functional promoter cassettes. After publication of the article, do the authors expect for other researchers to follow this scheme to get additional constitutive promoters ???

Our response: We hope the reviewer will agree that we have successfully used our pipeline to identify promoters that function in tobacco and Arabidopsis. The engineering constraints we applied in this manuscript are almost certainly not universally applicable to all potential uses. We expect that others may find that they can make use of our pipeline, likely with their own set of constraints, for identifying promoters with particular patterns of expression (e.g., expressed in a subset of tissues rather than constitutively). In addition, the pipeline can be readily adapted to other organisms, an approach we ourselves have taken in another publication currently under review.

2) Fig. 2 and 3: It is better to show fold change over negative control levels. Presented figures are difficult to guess strength of the signals obtained from positive samples.

Our response: We thank the reviewer for this suggestion and acknowledge that there are other ways to process and represent the data. We maintain that our choice is the clearest for readers, but understand that there is room for disagreement. To facilitate alternative analyses and provide full transparency, all data and scripts (including those for annotating and analyzing the data, as well as for generating figures) are now linked to Zenodo DOI: 10.5281/zenodo.8170303. The data files include individual outputs generated by the plate reader that are consolidated and annotated using R-script before normalized for plotting, and this pre-normalized consolidated dataset is also available for download.

3) Fig. 3A: Signal of RUBY is not easy to see. Pictures need improvements.

Our response: We put a significant amount of effort into photographing the subtle RUBY expression found in our transgenic lines, and, in fact, did a number of ‘blinded’ tests to verify that one plant part was indeed redder than the control. To provide more quantitative measurements of the RUBY signal, we have now added an unbiased analysis of the photographs in the manuscript, where the level of “red” within a region of interest was determined by using the a* axis in the CIELAB color space (Supplementary Figure S4).

4) Fig. 4: Degree of suppression are all around 0.5 fold or more, which, for me, indicates failure of the suppression system. These examples do not look like a practical system.

Our response: We agree that the repression system can be further optimized, and the practical application of the system would depend on the dynamic range desired.

5) I could not find proof of constitutive function of the developed promoters in the figures. This information is necessary.

Our response: For two of the transgenic Arabidopsis lines, we performed qPCR on flower, leaf, root, and seedlings. In all cases, we were able to detect reporter mRNA.

Thank you for again your time and consideration of our work.

Sincerely,

Dr. Jennifer Nemhauser

Department of Biology

University of Washington

jn7@uw.edu

Review: Building a pipeline to identify and engineer constitutive and repressible promoters — R2/PR13

Conflict of interest statement

I decleare no competing interests

Comments

The new revised version of this manuscript addresses points raised by another reviewer. While I share some of their doubts/criticism, I also believe the amount of work and the rigour with which this has been conducted and presented provides valuable information for future similar efforts and thus is of great interest for QPB’s readership.

Recommendation: Building a pipeline to identify and engineer constitutive and repressible promoters — R2/PR14

Comments

Dear authors, thanks for having tackled the last remaining issues and delivered such a good quality article. Please apologise for the long time needed. I’m glad to inform you that your manuscript has been accepted for publication.

Decision: Building a pipeline to identify and engineer constitutive and repressible promoters — R2/PR15

Comments

No accompanying comment.