Skip to main content
PLOS Biology logoLink to PLOS Biology
. 2024 Mar 1;22(3):e3002528. doi: 10.1371/journal.pbio.3002528

Structural and functional characterization of AfsR, an SARP family transcriptional activator of antibiotic biosynthesis in Streptomyces

Yiqun Wang 1, Xu Yang 1, Feng Yu 1, Zixin Deng 1, Shuangjun Lin 1, Jianting Zheng 1,2,*
Editor: Yunsun Nam3
PMCID: PMC10936776  PMID: 38427710

Abstract

Streptomyces antibiotic regulatory proteins (SARPs) are widely distributed activators of antibiotic biosynthesis. Streptomyces coelicolor AfsR is an SARP regulator with an additional nucleotide-binding oligomerization domain (NOD) and a tetratricopeptide repeat (TPR) domain. Here, we present cryo-electron microscopy (cryo-EM) structures and in vitro assays to demonstrate how the SARP domain activates transcription and how it is modulated by NOD and TPR domains. The structures of transcription initiation complexes (TICs) show that the SARP domain forms a side-by-side dimer to simultaneously engage the afs box overlapping the −35 element and the σHrdB region 4 (R4), resembling a sigma adaptation mechanism. The SARP extensively interacts with the subunits of the RNA polymerase (RNAP) core enzyme including the β-flap tip helix (FTH), the β′ zinc-binding domain (ZBD), and the highly flexible C-terminal domain of the α subunit (αCTD). Transcription assays of full-length AfsR and truncated proteins reveal the inhibitory effect of NOD and TPR on SARP transcription activation, which can be eliminated by ATP binding. In vitro phosphorylation hardly affects transcription activation of AfsR, but counteracts the disinhibition of ATP binding. Overall, our results present a detailed molecular view of how AfsR serves to activate transcription.


Streptomyces antibiotic regulatory proteins (SARPs) are widely distributed activators of antibiotic biosynthesis. This study combines cryo-EM structures with biochemical assays to explain how the Streptomyces coelicolor SARP AfsR activates transcription.

1. Introduction

Streptomycetes are multicellular bacteria with a complex developmental cycle [1]. They produce numerous bioactive natural products, including many antibiotics with important applications in medicine and agriculture. This capability makes streptomycetes one of the most important industrial microbial genera. Complex regulatory systems tightly govern the natural product biosynthesis of streptomycetes, involving both global and pathway-specific regulators. Streptomyces antibiotic regulatory protein (SARP) family regulators include many of the pathway-specific regulators and some global regulators. They are prominent initiators and activators of natural product biosynthesis and have mainly been found in actinomycetes [2]. Recently reported members of the SARP family include ChlF2 from Streptomyces antibioticus and MilR3 from Streptomyces bingchenggensis. ChlF2 activates the production of chlorothricin, known for its anti-inflammatory properties, and MilR3 stimulates the synthesis of an excellent insecticide milbemycin [3,4]. SARPs feature an N-terminal OmpR-type DNA-binding (ODB) domain and a C-terminal bacterial transcriptional activation (BTA) domain, collectively referred to as the SARP domain. The ODB domain is characterized by a winged helix-turn-helix (HTH) comprising 3 α-helices and 2 antiparallel β-sheets [5], whereas the BTA domain is all-helical [6]. In contrast to “small” SARPs that only contain an SARP domain, “large” SARPs possess extra domains at their C-terminals that are hypothesized to modulate the activity of the SARP domain (Fig 1A). “Small” SARPs are more frequently observed and extensively studied among actinomycetes, compared to their “large” counterparts [7]. SARPs bind to the target promoters having common features. They contain direct repeats, the 3′ repeat of which is located 8 bp from the −10 element, and each repeat is separated from the adjacent repeat by 11 bp or 22 bp, corresponding to 1 or 2 complete turns of the DNA helix, respectively [8].

Fig 1. The TIC of the SARP, RNAP, and afsS promoter DNA.

Fig 1

(A) Domain structures of SARPs. The N-terminal ODB together with the following BTA is referred as an SARP domain. S. coelicolor AfsR is a large SARP with additional NOD and TPR domains. (B) Hypothetical scheme for the regulation of S. coelicolor AfsR. (C) Fluorescence polarization assays of the SARP with afs box. Error bars represent mean ± SEM of n = 3 experiments. (D) Transcription assays with increasing concentrations (62.5 nM, 125 nM, 250 nM, 500 nM, 750 nM, 1,000 nM) of the SARP. CK represents the control group without the addition of the SARP. Data are presented as mean ± SEM from 3 independent assays. (E) Assembly of the SARP-TIC. The protein compositions in the dotted line boxed fractions are shown in the SDS-PAGE. The original gel image can be found in S1 Raw Images. (F) The afsS promoter fragment used for the SARP-TIC assembly. The −35 element, −10 element, the TSS, and the 6-bp noncomplementary bubble are denoted. The afs box is colored orange and contains DRup (the upstream direct repeat) and DRdown (the downstream direct repeat). The top (non-template, NT) strand and bottom (template, T) strand are colored light green and dark green, respectively. (G) Two views of cryo-EM map. The map was generated by merging the consensus map of the full SARP-TIC and the focused map of the SARP region in Chimera X. (H) Cartoon representation of the SARP-TIC structure. The subunits are colored as in the color scheme. The data underlying C, D, and E are provided in S1 Data. BTA, bacterial transcriptional activation; cryo-EM, cryo-electron microscopy; NOD, nucleotide-binding oligomerization domain; ODB, OmpR-type DNA-binding; RNAP, RNA polymerase; SARP, Streptomyces antibiotic regulatory protein; TIC, transcription initiation complex; TPR, tetratricopeptide repeat; TSS, transcription start site.

Streptomyces coelicolor AfsR is a pleiotropic, global regulator of both primary and secondary metabolism belonging to the “large” SARP group, containing an SARP (Met1-Ala270) domain, a conserved approximately 35 kDa dubbed nucleotide-binding oligomerization domain (NOD) (Ala271 to Glu618), an arm domain (Arg619-Glu777) as well as a tentative tetratricopeptide repeats (TPRs) sensor domain (Asp778-Arg993) [9] (Fig 1A). AfsR is found for its vital role in antibiotic biosynthesis. AfsR null mutants show defects in afsS transcription, as well as production deficiencies in actinorhodin and undecylprodigiosin [10]. It is presumed that upon binding of AfsR to a 22-base pair (bp) binding box (afs box), RNA polymerase (RNAP) holoenzyme is recruited to the afsS promoter, allowing for transcriptional initiation [9]. AfsS serves as a master regulator of secondary metabolism and nutritional stress response [11]. AfsR-AfsS system is widely distributed among Streptomyces, and expressing AfsR in the heterologous hosts can awaken silent antibiotic production genes [12]. Additionally, AfsR and the master regulator PhoP bind to overlapping sequences within PhoR-PhoP regulon promoters, such as pstS, phoRP, and glnR, exerting crosstalking regulatory control over the response to phosphate and nitrogen scarcity [13,14] (Fig 1B). AfsR can be phosphorylated by several serine/threonine kinases including AfsK, AfsL, and PkaG, and achieves better DNA binding affinity [12,15]. The full-length AfsR, as well as truncations containing only the SARP domain or lacking the TPR domain, can bind to afs box and activate afsS transcription in vitro [9]. However, the NOD is essential for the functionality of AfsR in vivo [10].

The RNAP holoenzyme consists of a dissociable σ subunit that establishes specific interactions with the −35 and −10 elements of the promoter to initiate transcription. The positioning of RNAP is primarily influenced by the contacts between σ region 4 (R4) and the −35 element, while the formation of an open complex is driven by the contacts between σ region 2 (R2) and the −10 element [16,17]. The afsS promoter comprises a suboptimal −10 element (CACTGT) and a poor −35 element (TTCAGC). This characteristic is commonly observed in streptomycete promoters [18,19], which complicates the identification of the promoters. The 22-bp afs box, centered at position −29.5 relative to the transcription start site (TSS) in the afsS promoter, overlaps with the spacer between the −10 and −35 elements. It is composed of two 11-bp direct repeats (DRs), with the upstream DR (−40 to −30) overlapping with the −35 element (−38 to −33) [20,21]. The catabolite activator protein (CAP), also known as the cAMP receptor protein (CRP), is an extensively studied transcription factor. The 22-bp CAP binding box (cap box) is centered at –61.5 in class I promoters and at –41.5 in class II promoters [2224]. The binding position of AfsR suggests that it may possess a distinct activation mechanism.

Due to the presence of the NOD domain, AfsR is also classified as the signal transduction ATPases with numerous domains (STAND) family [25]. Representative examples of STAND include APAF1/CED4 that regulates apoptosis [26] and plant resistance proteins that respond to pathogens and stress [27], as well as bacterial transcriptional activator MalT in Escherichia coli [12,28,29]. STAND proteins are presumed to keep in a monomeric inactive state by arm-based NOD–sensor autoinhibitory interactions or inhibitor binding in the absence of cognate inducer molecules [26,3032]. Activation involves a multistep process of inducer binding, autoinhibition release, nucleotide exchange, and conformational changes that switch the protein to an active state [33]. However, the mechanisms by which AfsR regulates itself and activates transcription are still obscure.

Here, we report cryo-electron microscopy (cryo-EM) structures of the SARP-transcription initiation complex (TIC), elucidating a transcription-activating mechanism that could be generally used by SARP family activators. We find AfsR uses its SARP domain to engage the afs box and make extensive interactions with all RNAP subunits except ω subunit. We confirm the inhibitory effect of NOD and TPR on SARP transcription activation by in vitro assays. The inhibitory effect can be eliminated by ATP binding, but phosphorylation of AfsR counteracts the disinhibition by ATP. Overall, we present a detailed molecular view of how AfsR activates transcription.

2. Results

SARP engages DNA and RNAP to activate transcription

Despite abundant genetical and biochemical studies on SARPs, the molecular basis of SARP-based transcription activation remains unknown. The isolated SARP domain of AfsR is able to bind afs box as evidenced by fluorescence polarization assays (Fig 1C), and mutating the upstream DR (M1) or the downstream DR (M2) attenuates this specific interaction (S1 Fig), in agreement with the previous report [9]. In vitro MangoIII-based transcription assays demonstrated the SARP domain could activate the transcription of afsS promoter (Fig 1D), while it has no significant effect on a control actII-4 promoter comprising a consensus −10 and −35 element but lacking the afs box (S1C Fig). Consistent with the DNA binding results, when the DRs were mutated, the SARP domain lost much of its ability to activate transcription, especially with mutation at the downstream DR (S1C Fig). To investigate the molecular basis of SARP-based transcription activation, we combined the isolated SARP domain, a DNA scaffold, RbpA, CarD, and RNAP σHrdB-holoenzyme to assembly the initiation complex (Fig 1E). The DNA scaffold is engineered from the afsS promoter, consisting of a 44-bp (−50 to −7) upstream promoter double-stranded DNA (dsDNA) which contains the 22-bp afs box and the consensus –10 element, a 6-bp (−6 to −1) non-complementary transcription bubble, and a 15-bp (+1 to +15) downstream promoter dsDNA (Fig 1F). RbpA and CarD, 2 RNAP-binding proteins discovered in actinobacteria, can stabilize the TIC and have been found in many TIC structures of Mycobacterium tuberculosis [3436]. Through in vitro transcription assays, we observed that SARP maintained its role as a transcriptional activator in the presence of RbpA and CarD (S2 Fig).

The final cryo-EM map of the SARP-TIC was reconstructed using a total of 95,223 single particles and refined to a nominal resolution of 3.35 Å, with approximately 3 Å at the center of RNAP and approximately 6 Å at the peripheral SARP (S3 Fig and S1 Table). Local refinement focused on the SARP region generated a 3.77-Å-resolution map. In the cryo-EM maps, the densities allowed unambiguous docking of 2 SARP protomers, a 61-bp promoter DNA (−46 to +15), a σHrdB subunit except disordered region 1.1, 5 subunits of RNAP core (α2ββ′ω), RbpA, and CarD. In addition, the cryo-EM density map shows a clear signal for an αCTD (Fig 1G). The 2 SARP protomers simultaneously engage the afs box and RNAP σHrdB-holoenzyme. SARP-TIC is an open complex containing a transcription bubble of 13 nt. The overall structure of S. coelicolor RNAP in SARP-TIC closely resembles that in our recently reported S. coelicolor RNAPσHrdB-Zur-DNA structure (PDB ID: 7X75, rmsd of 0.334 Å for 3,002 aligned Cα atoms) [37], and other actinobacteria RNAP structures including M. tuberculosis RNAP-promoter open complex (PDB ID: 6VVY, rmsd of 1.133 Å for 2,745 aligned Cα atoms) [38] and Mycobacterium smegmatis RNAP TIC (PDB ID: 5VI5, rmsd of 1.048 Å for 2,621 aligned Cα atoms) [39] (S4 Fig). As shown in Fig 1H, σHrdB R4 is positioned on top of 2 SARP protomers, instead of binding in the major groove of the −35 element as observed in the reported structures [3739], suggesting a sigma adaptation mechanism like M. smegmatis PafBC [40]. CarD binds the β-subunit and interacts with the upstream double-stranded/single-stranded (ds/ss) junction of the transcription bubble [41], while RbpA enhances interactions between β′, σHrdB, and promoter DNA [42].

A side-by-side SARP dimer interacts with afs box

The SARP-TIC structure elucidates the binding interactions between SARP protomers and 2 major grooves of the afs box, with each protomer making contacts with 1 DR (Fig 2A). The contacts span the first 8 conserved base pairs. In contrast, the last 3 base pairs do not make any contact with SARP (Fig 2B). Overall conformations of the 2 protomers are essentially the same with an overall rmsd of 0.8 Å (S5 Fig). A dimer interface of approximately 1,120 Å2 is formed between the 2 SARP protomers, revealing a unique side-by-side arrangement (Fig 2C). Since SARP is monomeric in solution, the dimerization of SARP occurs upon DNA binding. As a representative example of SARP family transcription activators, the SARP folds into an N-terminal ODB domain (residues 25–111) and a C-terminal BTA domain (residues 120–270) (Fig 2D). The ODB domain consists of 3 α helices packed against 2 antiparallel β sheets. Two helices, α2 and α3, and the seven-residue connecting loop constitute an HTH DNA-binding motif, followed by a β hairpin that contacts the strand between the helices α1 and α2. The BTA domain comprises 7 α-helices, of which the first 3 ones stand on the first β sheet and the first 2 helices of the N-terminal ODB domain, burying an interfacial area of approximately 860 Å2. The overall structure of SARP protomers closely resembles EmbR (PDB ID: 2FEZ, rmsd of 2.8 Å for 247 aligned Cα atoms) [6], a transcriptional regulator of M. tuberculosis, except that EmbR comprises an additional C-terminal forkhead-associated (FHA) domain (S6 Fig).

Fig 2. Interactions of SARP with afs box.

Fig 2

(A) Two SARP protomers bind to the afs box (−19 to −40) side-by-side with each protomer contacting 1 DR. The density map is shown as transparent surface. T-30 (nt) is the last nucleotide of DRup. Upstream and downstream SARP are colored blue and orange, respectively. (B) Conserved sequences corresponding to the 11-nt DR of the afs box generated by MEME. (C) The dimer interface between 2 SARP protomers. (D) The domain organization indicated in the downstream SARP protomer. (E) Detailed interactions of the upstream protomer with the DNA backbone. Hydrogen bonds and salt bridges are shown as yellow and red dashed lines, respectively. (F) Contacts of SARP with specific nucleotides. The residues R92 and T88 make hydrogen bonds (shown as yellow dashed lines) with O4 of T-38(nt) and N7 of G-36(t), respectively. (G) Sequence alignments of SARP regulators from different Streptomyces strains, highlighting the residues interacting with DNA (green). Only ODB domains were compared. These proteins include AfsR (P25941), ActII-4 (P46106), RedD (P16922) from S. coelicolor, PimR (Q70DY8) from S. natalensis, PolY (ABX24502.1), PolR (ABX24503.1) from S. asoensis, SanG (Q5IW77) from S. ansochromogenes, ChlF2 (Q0R4N4) from S. antibioticus, MilR3 (D7BZQ7) from S. bingchenggensis, CcaR (P97060) from S. clavuligerus, DnrI (P25047) from S. peucetius, MtmR (Q194R8) from S. argillaceus, and TlyS (M4ML56) from S. fradiae. The black boxes highlight the positions conserved. DR, direct repeat; ODB, OmpR-type DNA-binding; SARP, Streptomyces antibiotic regulatory protein.

The 2 SARP protomers establish a total contact surface of approximately 1,300 Å2 with the dsDNA, resulting in a 20° bend of the helical axis of the upstream dsDNA at T-30(nt) (-30T on the non-template strand) (S7 Fig). The DNA bending at the midpoint between the 2 binding sites suggests that both protomers equally contribute to curve the DNA. The DNA contacts made by the upstream SARP protomer were described in the following section (Fig 2E). The N-terminal end of helix α1, the C-terminal end of helix α2, and the HTH loop make extensive contacts with the backbone phosphate groups from C-40 to T-37 of the nt-strand, involving both hydrogen bonds (S46 and Q48) and van der Waals interactions (W74, P79, S80, and Q81). The recognition helix α3 penetrates the DNA major groove and is almost perpendicular to the DNA axis. R87, R94, and K95 form salt bridges with the backbone phosphate groups from T−35 to G-33 of the t-strand. The side chain of S91 makes a hydrogen bond with the phosphate group of T−35(t). In addition to these nonspecific contacts with backbone phosphate groups, the recognition helix α3 also establishes specific DNA contacts (Fig 2F). The R92 makes hydrogen bonds with the O4 atom of T-38(nt) via its guanidinium group. The side chain of T88 makes a hydrogen bond with N7 of G-36(t). Consistent with the structural observations, previous EMSAs show that changing the T-38(nt) of the upstream DR or the corresponding T-27(nt) of the downstream DR to adenosine prevents the binding of AfsR [21]. However, changing the G-36(t) to adenosine does not impair the binding of AfsR since A-25(t) is observed at the corresponding position of the downstream DR [21]. The 2 ODB domains make similar contacts with DNA, inserting the helix α3 of the HTH into the major groove of the DR (S8 Fig). Most residues involved in DNA interactions are highly conserved in both “small” and “large” SARP homologues across various Streptomyces strains (Fig 2G), including Q48, W74, P79, T88, R94, and K95.

SARP establishes extensive interactions with RNAP

Both SARP protomers make contacts with σHrdB R4, burying a total interface of 800 Å2 (Fig 3A). The upstream protomer contacts σHrdB R4 by its ODB domain, of which the HTH forms a concave surface at the N-terminal of the helix α3 to enfold the N-terminal of the helix α4 of σHrdB R4. The negatively charged residues E76 and E77 in the loop connecting helices α2 and α3, and the residues E68 and D71 in helix α2 of the ODB domain form salt bridges with R485 and R487 in σHrdB R4, respectively (Fig 3B). The interface where the shape and charge complement each other buries an area of approximately 320 Å2. The downstream protomer mainly uses its BTA domain to contact σHrdB R4, with a buried surface area of approximately 480 Å2 (Fig 3C). The helices α3 to α7 of the downstream BTA domain form a “chair” to let σHrdB R4 sit on it. The C-terminal of helix α4, the N-terminal of helix α5, and the connecting turn of σHrdB R4 fit in a groove formed by helices α3 to α7 of the BTA domain. The R498 at the C-terminal of helix α4 of σHrdB R4 makes contacts with the E176 and V172 on the “chair back.” The H499 in the connecting turn of σHrdB R4 is enfolded in an amphiphilic pocket formed by residues L211, L243, L247, V249, E213, and R252 of the BTA domain. The K496 at the C-terminal of helix α4 of σHrdB R4 makes a salt bridge with the E246 at the C-terminal of helix α6 of the BTA domain. In addition, the HTH loop of the ODB domain forms interactions with the loop connecting helices α2 and α3 of σHrdB R4.

Fig 3. SARP interacts with σHrdB R4, β FTH, β′ ZBD.

Fig 3

(A) The SARP protomers interact with σHrdB, β, and β′. (B) The upstream SARP protomer contacts σHrdB R4 by its ODB domain. Salt bridges are shown as red dashed lines. (C) The H499 of σHrdBR4 is enfolded in an amphiphilic pocket of the BTA domain. The K496 of σHrdBR4 makes a salt bridge with the E246 of the BTA domain. (D) The downstream SARP protomer make extensive interactions with the β FTH, the preceding loop (TPL) and the following loop (TFL) of the β flap. SARP is colored orange and β flap is colored blue. Hydrogen bonds, salt-bridges, and van der Waals interactions are shown as yellow, red, and gray dashed lines, respectively. (E) Interactions between the β′ ZBD and the ODB of downstream SARP. The positively charged R67 and R69 of β′ ZBD contact the negatively charged E76 and E77 of the HTH loop of the ODB domain. (F) Mutating interfacial residues of SARP impaired transcription activation. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3; *P < 0.05; **P < 0.01 in comparison with the wild-type SARP. (G) Sequence alignments of SARP regulators from different Streptomyces strains, highlighting the residues interacting with β (blue), σ R4 (cyan), and αCTD (purple). The black boxes highlight the positions conserved. BTA, bacterial transcriptional activation; FTH, flap tip helix; ODB, OmpR-type DNA-binding; SARP, Streptomyces antibiotic regulatory protein; ZBD, zinc-binding domain.

The downstream protomer also interacts with the RNAP β and β′ subunits. The ODB domain contacts the β-flap tip helix (FTH), the preceding loop (TPL), and the following loop (TFL) of the β flap by the helices α1 and β2, and the second antiparallel β sheet (Fig 3D). The residues D71 and E76 of the ODB make salt bridges with the R805 of the β FTH. The R59 at the C-terminal of the ODB α1 forms a salt bridge with the E817 in TFL. The R62 and T63 of the strand β4 connecting the helices α1 and α2 contact the R816 in TFL and the K795 in TPL, respectively. The S106 in the loop connecting the strands β5 and β6 forms a hydrogen bond with the main chain oxygen atom of the K749 in the loop following the first strand of the β flap. The BTA domain forms interactions with the C-terminal of the β FTH via the N-terminal of the helix α3, of which the R173 side chain is parallel to TFL. The positively charged R67 and R69 region of β′ zinc-binding domain (ZBD) complements the negatively charged E76 and E77 of the HTH loop of the ODB domain (Fig 3E).

To confirm the importance of the residues contacting RNAP subunits in transcription activation, they were replaced by alanine through site-directed mutagenesis. The mutants were purified and assayed by MangoIII-based transcription assays as the wild-type protein (Fig 3F). Mutating residues R59, D71, E76, and R173 diminished the capability of SARP to active transcription, whereas no obvious difference was observed between the T63A, E77A, E246A, L247A, and V249A mutants and the wild-type protein (S9 Fig). Mutating residues E176, L211, and L243 resulted in insoluble proteins when they were expressed under the same conditions as the wild-type protein. Some residues involved in RNAP interactions are highly conserved in SARP homologues across various Streptomyces strains (Fig 3G), including T63 and D71 for interaction with the β subunit, L243, E246, L247, and V249 for interaction with σ R4.

αCTD has versatility and important regulatory roles in modulating RNAP performance. In E. coli RNAP, it recognizes specific sequences called UP elements in some promoters [43]. Many regulators have been reported to activate transcription by interacting with αCTD. The SARP-TIC structure demonstrates that both BTA domains interact with the αCTD, burying an interface area of approximately 660 Å2 (Fig 4A). The αCTD is positioned on top of the helix bundle of the downstream BTA domain. The helix α2 of αCTD interacts with the helix bundles of both upstream and downstream BTA domains and is almost perpendicular to them. The loops preceding and following helix α2, and the loop connecting helices α4 and α5 of αCTD make contacts with the helix bundle of the downstream and upstream BTA domain, respectively. The αCTD interface is predominantly positively charged (R259, R266, R287, K292), complementing the negatively charged interface of BTA domains (D245, E246, and D250 of the upstream BTA domain) (Fig 4B). The residues R240 and L263 of the downstream BTA domain interact with αCTD by electrostatic and hydrophobic interactions, respectively. As shown in Fig 4C, the αCTD is required for SARP-dependent activation. Site-directed mutational studies further confirmed the importance of interfacial residues for SARP-dependent transcriptional activation. Moreover, most residues mentioned above for contacting the αCTD are highly conserved in SARP homologues, including R240, D245, E246, and L263 (Fig 3G).

Fig 4. Interactions of SARP with RNAP αCTD.

Fig 4

(A) BTA domains of both SARP protomers interact with the αCTD. Detailed interactions are shown in the gray box. (B) The electrostatic potential interface of BTA domains composed of R240, D245, E246, D250, and L263, and that of RNAP αCTD composed of E252, R259, K265, R266, R287, and K292. (C) Removing αCTD or mutating SARP residues involved in interactions with αCTD impaired transcription activation compared with wild-type SARP. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3; *P < 0.05; **P < 0.01; ***P < 0.001 in comparison with the wild-type SARP. (D) Proposed model for SARP-dependent transcription activation. BTA, bacterial transcriptional activation; RNAP, RNA polymerase; SARP, Streptomyces antibiotic regulatory protein.

Transcription activation of SARP is regulated by C-terminal domains

AfsR is a large SARP protein that contains additional NOD and TPR domains at its C-terminus, belonging to STAND family. STAND proteins are presumed to keep in an inactive state in the absence of stimulus signals and the activation involves a multistep of conformational changes. Here, we aimed to investigate whether the SARP effector domain of AfsR is similarly regulated. We then expressed and purified full-length AfsR and a truncation lacking the TPR domain (ΔTPR). Consistent with the isolated SARP, AfsR and ΔTPR migrate primarily as monomers in the SEC column [9] (S10 Fig). Notably, when assessing the transcription activation of full-length AfsR and ΔTPR, we observed that they exhibited lower levels compared to the isolated SARP domain (Fig 5A), indicating that the NOD and TPR domains exert an inhibitory effect on the transcription activation of the SARP domain.

Fig 5. ATP binding activates unphosphorylated AfsR.

Fig 5

(A) Transcription assays with increasing concentrations of AfsR as well as its truncations ΔTPR and SARP. (B) Sequence alignment of Streptomyces STAND family members highlighting the consensuses sequences of Walker A motif. The T337 phosphorylated by E. coli kinases is colored orange. (C) Transcription assays of 500 nM dephosphorylated AfsR preincubated with different additional nucleotides. Data are presented as mean ± SEM from 3 independent assays. (D) Representative SDS-PAGE analysis of proteolysis resistance of dephosphorylated AfsR, phosphorylated AfsR by AfsKΔC, AfsRT337A, and AfsR4E in the absence or presence of 1 mM ATP. Arrows and triangles indicate the primary AfsR bands and major degradation bands, respectively. Increasing concentrations of trypsin (0–100 μg/ml) were used. The original gel images can be found in S1 Raw Images. (E) Transcription assays of 500 nM AfsR (untreated), dephosphorylated AfsR (dephos), phosphorylated AfsR (phos), AfsRT337A and AfsR4E with or without preincubation with 1 mM ATP. CK represents the control group without the addition of AfsR. Data are presented as mean ± SEM from 3 independent assays. (F) ATPase activity assay of the AfsR4E, phosphorylated AfsR, and dephosphorylated AfsR. The AfsR4E showed Km of 0.995 ± 0.152 mM and kcat of 0.147 ± 0.00841 s−1. The phosphorylated AfsR showed Km of 0.767 ± 0.163 mM and kcat of 0.070 ± 0.00515 s−1. The dephosphorylated AfsR showed Km of 1.16 ± 0.231 mM and kcat of 0.0453 ± 0.00356 s−1. Data shown are the mean ± SEM for n = 3 experiments. (G) The cryo-EM map of AfsR-TIC. The map was generated by merging the consensus map of the full AfsR-TIC and the focused maps of the AfsR in Chimera X. Detailed interactions in the focused map focused on SARP are shown in the box. The data underlying A, C, E, and F are provided in S1 Data. cryo-EM, cryo-electron microscopy; SARP, Streptomyces antibiotic regulatory protein; STAND, signal transduction ATPases with numerous domains; TIC, transcription initiation complex.

In bacterial two-component systems, phosphorylation of the response regulator usually induces conformational changes, leading to its activation. Previous reports have indicated that AfsR can undergo phosphorylation by multiple serine/threonine kinases. Considering the possibility of phosphorylation by endogenous kinases in E. coli, we conducted mass spectrometry (MS) analysis on AfsR purified from E. coli. The MS results revealed phosphorylation at T337, the last residue of the Walker-A motif of the nucleotide-binding domain (NBD) (Fig 5B and S2 Table). We treated AfsR with λ phosphatase. The resulting dephosphorylated AfsR exhibited unaltered transcriptional activity compared to the untreated counterpart (Fig 5A and 5C). To imitate the phosphorylated AfsR in S. coelicolor, we purified AfsKΔC, a kinase encoded by a gene located in the same cluster as the afsR. We utilized AfsKΔC to phosphorylate AfsR (pre-dephosphorylated) and subsequently performed MS assays. The results indicated phosphorylation at residues S22, T337, S391, T506, and S953 (S11A Fig and S2 Table). Among these phosphorylated residues, T337, S391, T506, and S953 in NOD and TPR domains are conserved (S11B Fig). Furthermore, phosphorylated AfsR showed slightly higher transcriptional activation compared to the unmodified form (S11C Fig). It has been demonstrated that in vitro phosphorylation of AfsR by AfsK yields a very small population of the phosphorylated AfsR due to the low phosphorylation efficiency [10]. To further investigate the impact of phosphorylation, we mimicked the phosphorylation by introducing a glutamate substitution at each site. Four mutations (S391E, T337E, T506E, and S953E) marginally contribute to transcriptional activation (S11D Fig). Subsequently, glutamate substitutions were introduced at all 4 sites (4E). The 4E mutant formed more oligomers (S11E Fig) and demonstrated stronger binding to the afs box (S11F Fig) compared to the dephosphorylated AfsR, which is consistent with previous experiments conducted with phosphorylated AfsR [10]. However, the transcription activation of the 4E mutant was still only slightly higher than that of dephosphorylated AfsR (S11D Fig), suggesting that phosphorylation mediated by AfsK likely does not trigger activation of AfsR.

The NOD domain of AfsR is known to hydrolyze ATP and GTP [10]. Thus, AfsR was preincubated with ATP or GTP to test the effects on transcription activation. As depicted in Fig 5C, preincubating dephosphorylated AfsR with ATP or GTP significantly enhances its transcription activation. In the case of incubation with ATP, the transcription activation is comparable to that of the isolated SARP (Fig 5A and 5C). However, preincubation with ADP or GDP exerted no discernible influence. Besides, the activation of AfsR was likely triggered by ATP binding rather than ATP hydrolysis, as evidenced by the comparable effect of preincubation with a non-hydrolyzable ATP analogue, ATPγS. This behavior is reminiscent of MalT, another STAND regulator, where ATP binding instead of ATP hydrolysis is essential for transcription activation [44,45]. To gain further insights, limited protease digestion experiments were conducted, revealing that preincubation with ATP induces conformational changes in dephosphorylated AfsR and AfsRT337A mutant so that they are more resistant to trypsin digestion (Fig 5D). However, preincubation with ATP does not affect the oligomerization state or DNA-binding capacity of AfsR (S12 Fig). Interestingly, phosphorylated AfsR and the 4E mutant, which mimics phosphorylation, are no longer activated by preincubation with ATP (Fig 5E), likely due to the increased ATPase activity (Fig 5F), which converts ATP to ADP. Phosphorylated AfsR, along with the 4E mutant, exhibits partial activation in response to non-hydrolysable ATPγS (S13 Fig). In addition, the limited protease digestion profiles of phosphorylated AfsR and the 4E mutant remained unchanged in the presence of ATP. However, the major digestion bands detected around 40 kDa in dephosphorylated AfsR were absent in phosphorylated AfsR and the 4E mutant (Fig 5D).

We tried to find out the interplay between SARP and its C-terminal domains during transcription activation by assembling AfsR-TIC (S10 Fig). AfsRT337A was used to eliminate the interference of E. coli endogenous kinases and ATPγS was used to active AfsR. The final cryo-EM map was refined into a nominal resolution of 3.63 Å (S14 Fig). But the density extending from the C-terminal of SARP was very ambiguous and could not be improved by 3D classification and local refinement, presumably due to the high flexibility of the NOD and TPR domains relative to the SARP domain. The structures of SARP and RNAP in the AfsR-TIC resemble those in SARP-TIC (Figs 5G and S15), indicating that the SARP domain in the full-length AfsR and the isolated SARP activates transcription by similar mechanism.

3. Discussion

Streptomycetes exhibit a wide diversity in promoter sequences and transcription patterns to finetune the secondary metabolism to synthesize numerous natural products in response to environmental change. SARPs are well-known transcription activators of antibiotic biosynthesis. The majority of filamentous Actinobacteria genera are known to possess SARP-type regulatory genes, with examples including 98% of Streptomyces and 100% of Salinispora, while the occurrence in Mycobacterium is lower, at only 42% [46]. “Small” SARP regulators exhibit higher prevalence than “large” ones [7]. These “small” SARPs hold significant potential as tools for activating biosynthetic gene clusters, as evidenced by the manipulations of SARPs to enhance the yield of natural products in the native producers or the heterologous hosts [7,4749].

The diminished basal activity of SARP-regulated promoters may be attributed to deficiencies in their −10 region, −35 region, and the extended spacer between the −10 and −35 elements. Here, we report a strategy used by SARP to circumvent the recognition of a suboptimal promoter and activate transcription. Given the comparable protein architecture of SARPs and their binding to similar motifs, SARPs are likely to function in a similar mode (Fig 4D). The position of afs box (centered at −29.5) is different from the binding boxes of typical class II transcriptional activators (centered at −41.5). Consequently, the downstream SARP protomer is closer to RNAP and makes extensive contacts with β and β′ subunits (Fig 3A). Within the scope of our investigation, SARP demonstrates the most comprehensive interaction with RNAP among the studied transcription activators, encompassing all subunits of RNAP except for ω, namely α, β, β′, and σ subunits. Particularly, SARP provides an alternative docking point for σ R4. In reported class II activator-TIC structures [23], the σ R4 binds in the major groove of the −35 element and interacts with class II activators such as CAP [22], SoxS [50], and Rob [51] (S16 Fig). The transcriptional activation mode of SARP mostly resembles that of the mycobacterial transcriptional activator PafBC termed “sigma adaptation” [40] (S16F Fig). PafBC inserts between PafBC-specific −26 element and σ R4 to facilitate the recognition of promoters with suboptimal −35 elements. It has been reported that most promoters of S. coelicolor and M. tuberculosis have a conserved −10 element, but a highly variable −35 element [18,19,52]. Our study showed the “sigma adaption” principle compensating for the suboptimal −35 elements may be widely adopted in actinobacteria, as indicated by the ubiquitous presence of SARP regulators. In addition, both SARP protomers make contacts with the αCTD, enhancing the ability to activate transcription by recruiting RNAP. Notably, although many activators in reported TIC structures are dimer, they only use 1 protomer to contact αCTD. In class I CAP-TAC structures, the downstream protomer contacts the αCTD [22]. In contrast, the upstream protomer contacts the αCTD in class II TAP-TAC structures [23].

Like other STAND regulators, the transcription activation of AfsR exerted by the SARP effector domain is inhibited by the NOD and TPR domains (S17 Fig). By in vitro transcriptional assays, we demonstrate that preincubation of AfsR with ATP enhances its capability of activating transcription comparable to that of the isolated SARP domain. Preincubation with ADP does not enhance transcription activation of AfsR, suggesting a role for the triphosphate molecule in modulating AfsR function. Limited protease digestions indicate ATP may induce conformational changes of AfsR. However, the precise mechanism still requires further investigation. The prototypical bacterial STAND regulators MalT and GutR are activated by the sequential binding of inducer (maltotriose and glucitol, respectively) and ATP [29,53]. AfsR contains a tentative TPR sensor domain; however, it remains elusive whether there is an inducer binding to the AfsR sensor domain. AfsR is phosphorylated by several kinases including AfsK, which is located upstream of AfsR and separated by 2 genes. We found 5 phosphorylation sites in AfsR. The 4 phosphorylation sites (T337, S391, T506, and S953) in NOD and TPR are highly conserved in AfsR homologs (S11B Fig). In contrast to the response regulators of two-component systems that are activated by phosphorylation, phosphorylation modifications have little effect on AfsR transcriptional activation. However, preincubation of phosphorylated AfsR with ATP does not enhance its transcription activation, suggesting that phosphorylation may counteract the effects of ATP binding. We observed that phosphorylated AfsR exhibits higher ATPase activity, but further investigations are required to decipher the detailed mechanism of how ATP binding and phosphorylation modulate AfsR transcription activation.

In summary, transcription activators of the SARP family have been widely found in actinomycetes and play important roles in regulating secondary metabolism. Our detailed structural and functional analysis provides a molecular basis for understanding SARP-mediated antibiotic regulation in streptomycetes and offers potential optimizations for antibiotic production.

4. Materials and methods

Plasmids

The genes encoding AfsR, ΔTPR (residues from 1 to 618 of AfsR), SARP (residues from 1 to 270 of AfsR), RbpA, and CarD were cloned from the genomic DNA of S. coelicolor M145, and inserted into the pET28a via NdeI and EcoRI restriction sites to obtain pET28a-SARP, pET28a-rbpA, and pET28a-carD, respectively. The gene encoding AfsKΔC (Met1 to Arg311) was cloned from the genomic DNA of S. coelicolor M145 and inserted into the pET32a vector. Plasmids carrying SARP mutants were constructed using site-directed mutagenesis. Primers used are listed in S3 Table.

Protein expression and purification

The vector pET28a-afsR, pET28a-ΔTPR, pET28a-SARP, pET32a-afsKΔC, pET28a-rbpA, and pET28a-carD were transformed into BL21(DE3), respectively. Induction of expression was achieved by adding 0.3 mM isopropyl-β-D-thiogalactopyranoside (IPTG) and incubated for 12 h at 16°C and 220 rpm. The cells were then harvested and resuspended in buffer A (50 mM Tris (pH 8.0), 500 mM NaCl, 10% glycerol, 1 mM β-mercaptoethanol, and 5 mM imidazole). After purification by nickel-NTA, the eluate was further loaded onto a size-exclusion chromatography (SEC) column (Superdex 200, Cytiva) equilibrated with buffer B (20 mM Tris (pH 8.0), 150 mM NaCl, 5% glycerol) and the purified proteins were stored at −80°C. Purification of S. coelicolor RNAP and σHrdB was carried out as described previously [37].

In vitro transcription

The transcription activities are evaluated by MangoIII-based transcription assay as previously described [37]. DNA fragments containing the afsS promoter (−50 to +15) or its mutants and the MangoIII sequence were used as transcription templates. Primers used are listed in S3 Table. A gradient concentration range of AfsR or its variants (ranging from 62.5 nM to 1,000 nM) was combined with 10 nM of promoter DNA at 30°C for 10 min, followed by supplementation of 100 nM of RNAP-σHrdB and 1 mM of NTPs (0.25 mM NTP each) in transcription buffer (20 mM Tris–HCl (pH 8.0), 100 mM KCl, 5 mM MgCl2, 1 mM DTT, 4 U RNaseIn, 1 μm TO1–PEG–biotin, and 5% glycerol). The reactions were incubated for 15 min at 30°C and stopped by 0.5 ng/μl (final concentration) heparin. In the investigation of the impact of RbpA and CarD on the functionality of SARP and the transcription of afsS promoter, as well as subsequent experiments involving AfsR or SARP mutants, RbpA and CarD were co-incubated within the transcriptional system as global transcription factors [54,55]. A total of 500 nM (final concentration) of AfsR, SARP or their variants, RbpA, and CarD were used. For investigating the impact of ATP, ATPγS, GTP, ADP, or GDP, 500 nM dephosphorylated AfsR was preincubated with 1 mM corresponding NTP or NDP for 10 min at 30°C and transferred into the reaction mixture (final concentration of additional nucleotides was less than 0.02 mM to avoid the influence on transcription). The reaction without NTP was used as blank. A multi-detection microplate reader (Tecan Spark) was used to measure fluorescence intensities. The emission wavelength was 535 nm and the excitation wavelength was 510 nm.

Fluorescence polarization (FP)

A 32-bp afs box (−45 to −14) or its mutants consisting of a 5′-conjugated FAM was synthesized, annealed, and used as the DNA probe. The reaction mixture contained 10 nM of labeled dsDNA and 0 to 2 μm purified AfsR, SARP, or their variants in binding buffer (20 mM HEPES (pH 7.5), 100 mM NaCl, 5 mM MgCl2, 2 mM DTT, 5% glycerol, and 0.1 μg/ml of poly(dI-dC)). After incubation at 30°C for 20 min, the fluorescence polarization of the reaction mixture was detected by a multi-detection microplate reader (Tecan Spark) with excitation wavelength of 485 nm and emission wavelength of 535 nm. Data from technical triplicates were fitted to a binding equation Y = Bmax*X/(kD-ave+X) to obtain the kD-ave, where Y is the ΔmP measured at a given protein concentration (X) and Bmax is the maximum ΔmP of completely bound DNA.

ATPase activity

ATPase activities were evaluated by using Malachite Green Phosphate Detection Kit (Beyotime). Each reaction consisted of 70 μl volumes containing 5 μm of protein in a reaction mixture of 20 mM Tris-HCl (pH 7.5), 150 mM NaCl, 5 mM MgCl2, and ATP at varying concentrations. Thirty minutes after ATP addition, a multi-detection microplate reader (Tecan Spark) was used to measure the absorbance at 620 nm. At least 3 technical replicate data were fitted to the Michaelis–Menten equation to generate kinetic parameters.

Dephosphorylation and phosphorylation of AfsR proteins

AfsR protein (5 μm) purified from E. coli was treated with λ protein phosphatase (100U, Beyotime) in the reaction buffer (50 mM HEPES (pH 7.5), 10 mM NaCl, 1 mM MnCl2, 2 mM DTT, 0.01% Brij35) for 30 min at 30°C, and then loaded onto the Superdex 200 SEC column equilibrated with 20 mM Tris-HCl (pH 8.0), 150 mM NaCl, and 5% glycerol to remove the phosphatase and the aggregation and to obtain dephosphorylated AfsR. In vitro phosphorylation of the pre-dephosphorylated AfsR was achieved by mixing the protein with AfsKΔC in a reaction buffer (40 mM Tris-Cl (pH 8.0), 5 mM MgCl2, 5 mM MnCl2, and 5% glycerol) for 30 min at 30°C, followed by separation by the Superdex 200 SEC column equilibrated with the 20 mM Tris (pH 8.0), 150 mM NaCl, 5% glycerol to obtain the phosphorylated AfsR.

LC-MS/MS analysis

AfsR and phosphorylated AfsR were digested with trypsin (Promega) and analyzed with Nano-LC-MS/MS system (Easy nLC1200/Q Exactive Plus, Thermo Scientific, Bremen, Germany), respectively. The mobile phase buffer A was 0.1% formic acid in water and the mobile phase buffer B was 80% acetonitrile with 0.1% formic acid. The flow rate was 300 nL/min, and the procedure was 2% to 20% buffer B in 42 min, 20% to 35% in 5 min, 35% to 100% in 1 min, and then 100% for 12 min. For MS, the scan ranged from 350 to 1,800 m/z. For MS/MS, the scan ranged from 200 to 2,000 m/z. Sequence analysis was performed by proteome Discoverer (version 2.5) software using the sequence of AfsR, AfsKΔC, trypsin as well as the database of expression strain E. coli BL21(DE3) and up to 2 missed cleavages were permitted. The maximum false discovery rate (FDR) for both peptides and proteins was set at 1%.

Assembly of SARP-TIC and AfsR-TIC complex

To construct SARP-TIC and AfsR-TIC, we used a DNA scaffold engineered from afsS promoter carrying 6-bp pre-melted transcription bubble. The DNA was prepared by annealing nt-strand (100 μm final) and t-strand (110 μm). For assembly of SARP-TIC, 32 μm SARP was preincubated with 5 μm afsS promoter for 10 min at 25°C. At the same time, 4 μm RNAP core, 16 μm σHrdB, 16 μm CarD, and 32 μm RbpA were incubated for 10 min at 25°C, then mixed with preincubated SARP-DNA and incubated for another 10 min at 25°C in assembling buffer (20 mM Tris-HCl (pH 8.0), 50 mM KCl, 5 mM MgCl2, and 3 mM DTT). For assembly of AfsRT337A-TIC, 32 μm AfsRT337A was preincubated with 5 μm afsS promoter as well as 1 mM ATPγS for 10 min at 25°C, and 4 μm RNAP core, 16 μm σHrdB, 16 μm CarD, and 32 μm RbpA were incubated for 10 min at 25°C, then mixed with preincubated AfsRT337A-DNA and incubated for another 10 min at 25°C in assembling buffer. We then loaded the complexes onto a Superose 6 10/300 GL column (GE Healthcare), eluted with assembling buffer, and used them directly for cryo-EM grid preparation.

Cryo-EM data acquisition and processing

Approximately 0.5 mM ATPγS was supplemented to AfsRT337A-TIC, and 3.5 μl of the SARP-TIC or AfsRT337A-TIC samples were added to freshly glow-discharged Quantifoil R1.2/1.3 Au 300 mesh grids. In a Vitrobot (FEI), grids were plunge-frozen in liquid ethane after being blotted for 2 s at 16°C with 100% chamber humidity. The grids were imaged using SerialEM on Titan Krios 300 kV microscopes with a K2 detector. The defocus ranged from −1.0 to −2.0 μm, and the magnification was ×130,000 in counting mode; 32 frames per movie were collected with a total dose of 40 e2. For SARP-TIC, 3,741 movies were collected, while 3,078 movies were collected for AfsRT337A-TIC.

The SARP-TIC or AfsRT337A-TIC data was processed using CryoSPARC suite v4.2.1. After motion correction, patch CTF estimation, manual exposure curation, and template picker using selected 2D classes from blob picker, 2 rounds of 2D classification were conducted [56]. For SARP-TIC, 603,659 particles were picked and used for ab initio reconstruction (4 classes). The selected particles (193,644 particles) were used for the second round of ab initio reconstruction (3 classes) and heterogeneous refinement. The second 3D class (95,223 particles, 49.2%) was selected and refined. Then, local resolution estimation and local filtering were conducted to give final maps. By the process of particle subtraction and masked local refinements to reserve only the signal around the SARP region, a local map with improved quality was obtained (S3 Fig and S1 Table). For AfsRT337A-TIC, 287,126 particles were picked and used for ab initio reconstruction (3 classes). The selected particles (131,066 particles) were used for the second round of ab initio reconstruction (3 classes) and heterogeneous refinement. The first 3D class (45,208 particles, 34.5%) was selected and refined by local resolution estimation as well as local filtering to give final map. By the processes of particle subtraction and masked local refinements to reserve only the signal around the SARP region or AfsRT337A region, 2 local maps were obtained, respectively (S14 Fig and S1 Table).

Cryo-EM model building and refinement

The initial models of S. coelicolor RNAPσHrdB, RbpA, and CarD were generated from PDB:7X75 [37], 5TW1 [42], and 4XLR [41], respectively. The initial atomic model of AfsR was generated by AlphaFold [57]. The model of the promoter DNA was built in Coot [58]. The models of RNAP, SARP, and DNA were fitted into the cryo-EM density maps using ChimeraX [59]. The model was refined in Coot and Phenix with secondary structure, rotamer, and Ramachandran restraints [60]. The validation was performed by MolProbity [61]. The Map versus Model FSCs was generated by Phenix (S3 and S14 Figs). The statistics of cryo-EM data processing and refinement were listed in S1 Table.

Supporting information

S1 Fig. In vitro assays of the Streptomyces coelicolor AfsR SARP.

(A) Core promoter sequences used for in vitro assays. The mutations introduced into the afs box are highlighted. The mutated target sites contained mutations in upstream repeat (M1), the downstream repeat (M2), or both repeats (M1M2). The actII-4 promoter was used as a control. (B) Fluorescence polarization assays of the SARP with mutant afs box (M1, M2, and M1M2). Error bars represent mean ± SEM of n = 3 experiments. (C) In vitro MangoIII-based transcription assays with or without 500 nM SARP in the absence of RbpA and CarD. Error bars represent mean ± SEM of n = 3 experiments. The data underlying B and C can be found in S1 Data.

(TIF)

pbio.3002528.s001.tif (1.3MB, tif)
S2 Fig. RbpA and CarD do not influence the transcriptional activation capability of SARP.

In vitro transcription assays with or without RbpA and CarD on the afsS promoter (left) and transcription assays with or without 500 nM SARP in the presence of RbpA and CarD (right). The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3.

(TIF)

pbio.3002528.s002.tif (270.8KB, tif)
S3 Fig. Single-particle cryo-EM analysis of SARP-TIC.

(A) A motion-corrected image (scale bar: 40 nm). (B) 2D classes (scale bar: 100 Å). (C) Data processing pipeline for the dataset of SARP-TIC. The final cryo-EM map of the SARP-TIC was reconstructed using a total of 95,223 single particles and refined to a nominal resolution of 3.35 Å. Local refinement focused on the SARP region generated a 3.77-Å-resolution map. (D) Validation of cryo-EM structural models. Map vs. model FSCs was generated by Phenix (Version 1.19.2). The model-map resolution for the atomic model and cryo-EM map at FSC = 0.5 cutoff was indicated in the figure and reported in S1 Table. The data underlying this figure can be found in S1 Data.

(TIF)

pbio.3002528.s003.tif (2.2MB, tif)
S4 Fig. Comparisons of RNAP.

(A) S. coelicolor SARP-TIC. (B) S. coelicolor RNAPσHrdB-Zur-DNA (PDB ID: 7X75). (C) Mycobacterium tuberculosis RNAP-promoter open complex (PDB ID: 6VVY). (D) Mycobacterium smegmatis TIC (PDB ID: 5VI5).

(TIF)

S5 Fig. Comparison of 2 SARP protomers in SARP-TIC.

Overall conformations of the 2 protomers are essentially the same with an overall rmsd of 0.8 Å.

(TIF)

pbio.3002528.s005.tif (3.7MB, tif)
S6 Fig. Comparison of SARP with EmbR (PDB ID: 2FEZ) comprising an additional C-terminal FHA domain.

(TIF)

pbio.3002528.s006.tif (2.4MB, tif)
S7 Fig. The binding of 2 SARP protomers in afs box results in a 20° bend of the helical axis of the dsDNA.

(TIF)

pbio.3002528.s007.tif (1.9MB, tif)
S8 Fig. Interactions of the downstream SARP with afs box.

(A) Downstream SARP residues involved in interactions with the DNA backbone. (B) Contacts of downstream SARP with specific nucleotides. The residues R92 and T88 make hydrogen bonds (shown as yellow dashed lines) with T-27(nt) and A-25(t), respectively.

(TIF)

pbio.3002528.s008.tif (2.8MB, tif)
S9 Fig. In vitro assays of SARP mutants.

(A) In vitro transcription assays of SARP T63A, E77A, E246A, L247A, and V249A mutants. No obvious difference was observed between the T63A, E77A, E246A, L247A, V249A mutants, and the wild-type protein. n.s. means no significance. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3. (B) SDS-PAGE of E176A, L211A, and L243A mutants. Inclusion bodies were formed when these mutants were expressed under the same conditions as the wild-type protein. P represents precipitation while S represents supernatant. The original gel images can be found in S1 Raw Images.

(TIF)

pbio.3002528.s009.tif (720.5KB, tif)
S10 Fig. Purification and assembly of protein complexes.

(A) Assembly of S. coelicolor RNAP-σHrdB. The peak of RNAP-σHrdB was analyzed by SDS-PAGE. (B) Purification of AfsR. (C) Purification of ΔTPR. (D) Purification of SARP. (E) Purification of RbpA. (F) Purification of CarD. (G) Purification of AfsKΔC. (H) Assembly of AfsRT337A-TIC in the presence of 1 mM ATPγS. The protein compositions in the dotted line boxed fractions are shown in the SDS-PAGE. The original gel images can be found in S1 Raw Images.

(TIF)

S11 Fig. Phosphorylation modulates AfsR transcriptional activation.

(A) LC-MS/MS analysis showing that S22, T337, S391, T506, and S953 are phosphorylated in AfsR by AfsKΔC. The lowercase letter in the peptide sequence indicates phosphorylated residue. “b” and “y” denote peptide fragment ions retaining charges at the N and C terminus, respectively. The subscript numbers indicate their positions in the identified peptide. (B) Sequence alignment of Streptomyces AfsR family members highlighting the consensuses sequences neighboring phosphorylation sites S22, T337, S391, T506, and S953. Orange boxes represent phosphorylation sites. (C) Transcription assays with increasing concentrations of phosphorylated AfsR by AfsKΔC (phos-AfsR) and untreated AfsR. (D) Transcription assays of 500 nM AfsR mutants mimicking dephosphorylation (T337A) and phosphorylation (S22E, T337E, T506E, S953E, and S391E/T337E/T506E/S953E (4E)). Data are presented as mean ± SEM from 3 independent assays. n.s. means no significance; *P < 0.05; **P < 0.01 in comparison with the wild-type AfsR. (E) Representative SEC assay of AfsRT337A and AfsR4E. (F) Fluorescence polarization assay of AfsR4E and dephosphorylated AfsR (dephos-AfsR) with afs box. The concentration of afs box was 10 nM. Error bars represent mean ± SEM of n = 3 experiments. The data underlying A, C, D, E, and F are provided in S1 Data.

(TIF)

pbio.3002528.s011.tif (1.1MB, tif)
S12 Fig. Oligomerization state and DNA-binding capacity of dephosphorylated AfsR are not affected by ATP binding.

(A) SEC assay of dephosphorylated AfsR in the presence (solid line) or absence (dashed line) of 1 mM ATP. The data underlying this figure can be found in S1 Data. (B) Fluorescence polarization assay of dephosphorylated AfsR with afs box DNA in the presence or absence of 1 mM ATP. The concentration of afs box DNA was 10 nM. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3.

(TIF)

pbio.3002528.s012.tif (538.3KB, tif)
S13 Fig. ATPγS activates phosphorylated AfsR and AfsR4E.

Transcription assays involving 500 nM phosphorylated AfsR (phos-AfsR) or AfsR4E, with and without preincubation with 1 mM ATP or ATPγS. CK represents the control group without the addition of AfsR. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3.

(TIF)

pbio.3002528.s013.tif (87.2KB, tif)
S14 Fig. Single-particle cryo-EM analysis of AfsRT337A-TIC.

(A) Motion-corrected images (scale bar: 50 nm). (B) 2D classes (scale bar: 100 Å). (C) Data processing pipeline for the dataset of AfsRT337A-TAC. The final cryo-EM map of the AfsRT337A-TIC was reconstructed using a total of 45,208 single particles and refined to a nominal resolution of 3.63 Å. Local refinement focused on the SARP region generated a 5.10-Å-resolution map. Local refinement focused on the AfsR region generated a 8.61-Å-resolution map. (D) Validation of cryo-EM structural models. Map vs. model FSCs was generated by Phenix (Version 1.19.2). The model-map resolution for the atomic model and cryo-EM map at FSC = 0.5 cutoff was indicated in the figure and reported in S1 Table. The data underlying this figure can be found in S1 Data.

(TIF)

pbio.3002528.s014.tif (2.2MB, tif)
S15 Fig. Two views of cryo-EM map and structural model of AfsRT337A-TIC.

The map was generated by merging the consensus map of the full AfsR -TIC and the focused maps of the AfsR. The initial model of RNAP and AfsR was generated from SARP-TIC and AlphaFold, respectively, and fitted into the map using ChimeraX.

(TIF)

pbio.3002528.s015.tif (8.3MB, tif)
S16 Fig. Interactions of regulators with σR4 and −35 element.

(A) SARP. (B) Class I CAP-transcription activation complex (TAC) (PDB ID:6B6H). (C) Class II TAP-TAC (PDB ID: 5I2D). (D) SoxS (PDB ID: 7W5W). (E) RobR (PDB ID: 7VWZ). (F) PafBC (PDB ID: 7P5X). The key residues Arg and Glu supposed to contact DNA are shown as sticks.

(TIF)

pbio.3002528.s016.tif (8.8MB, tif)
S17 Fig. Scheme model summarizing the ATP-dependent and phosphorylation-dependent regulation of AfsR.

Unmodified full-length AfsR is autoinhibited by its C-terminal domains. AfsR is activated by ATP binding and significantly stimulates the production of transcripts. The phosphorylation of AfsR results in increased ATPase activity, potentially counteracting the effects of ATP binding. Phosphorylated AfsR forms more oligomers and shows slightly higher transcriptional activation compared to the unmodified form.

(TIF)

pbio.3002528.s017.tif (1.1MB, tif)
S1 Table. Cryo-EM data collection, refinement, and validation statistics.

(XLSX)

pbio.3002528.s018.xlsx (11.8KB, xlsx)
S2 Table. Phosphorylated peptides in AfsR by mass spectrometric.

(XLSX)

pbio.3002528.s019.xlsx (10.1KB, xlsx)
S3 Table. List of primer sequences used in this study.

(XLSX)

pbio.3002528.s020.xlsx (11.5KB, xlsx)
S1 Data. Numerical values used to generate graphs.

(XLSX)

S1 Raw Images. The original gel images contained in the figures.

(PDF)

pbio.3002528.s022.pdf (13.9MB, pdf)

Acknowledgments

We thank Jing Liu, Xinqiu Guo, and Mengyu Yan at the Instrument Analysis Center (IAC) of Shanghai Jiao Tong University for supporting the cryo-EM data collection.

Abbreviations

BTA

bacterial transcriptional activation

CAP

catabolite activator protein

CRP

cAMP receptor protein

cryo-EM

cryo-electron microscopy

DR

direct repeat

dsDNA

double-stranded DNA

FDR

false discovery rate

FHA

forkhead-associated

FTH

flap tip helix

HTH

helix-turn-helix

MS

mass spectrometry

NBD

nucleotide-binding domain

NOD

nucleotide-binding oligomerization domain

ODB

OmpR-type DNA-binding

RNAP

RNA polymerase

SARP

Streptomyces antibiotic regulatory protein

SEC

size-exclusion chromatography

STAND

signal transduction ATPases with numerous domains

TIC

transcription initiation complex

TPR

tetratricopeptide repeat

TSS

transcription start site

ZBD

zinc-binding domain

Data Availability

Atomic coordinate has been deposited in PDB with accession numbers 8HVR (SARP-TIC) and 8JKE (AfsRT337A-TIC). Four cryo-EM density maps have been deposited in Electron Microscopy Data Bank with accession number EMD-35047 (SARP-TIC), EMD-36370 (AfsRT337A-TIC), EMD-35046 (SARP region from SARP-TIC) and EMD-36369 (SARP region from AfsRT337A-TIC). All data underlying the main and supplementary figures can be found at supplementary S1 Data and S1 Raw Images.

Funding Statement

This work was supported by National Key R&D Program of China (2019YFA0905400) and National Natural Science Foundation of China (32070040, 32370071). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. URLs:https://chinainnovationfunding.eu/national-key-rd-programmes/; https://www.nsfc.gov.cn/.

References

  • 1.Flardh K, Buttner MJ. Streptomyces morphogenetics: dissecting differentiation in a filamentous bacterium. Nat Rev Microbiol. 2009;7(1):36–49. doi: 10.1038/nrmicro1968 . [DOI] [PubMed] [Google Scholar]
  • 2.Bibb MJ. Regulation of secondary metabolism in streptomycetes. Curr Opin Microbiol. 2005;8(2):208–215. doi: 10.1016/j.mib.2005.02.016 . [DOI] [PubMed] [Google Scholar]
  • 3.Li Y, Zhang J, Zheng J, Guan H, Liu W, Tan H. Co-expression of a SARP Family Activator ChlF2 and a Type II Thioesterase ChlK Led to High Production of Chlorothricin in Streptomyces antibioticus DSM 40725. Front Bioeng Biotechnol. 2020;8:1013. Epub 20200821. doi: 10.3389/fbioe.2020.01013 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Yan YS, Yang YQ, Zhou LS, Zhang L, Xia HY. MilR3, a unique SARP family pleiotropic regulator in Streptomyces bingchenggensis. Arch Microbiol. 2022;204(10):631. doi: 10.1007/s00203-022-03240-x . [DOI] [PubMed] [Google Scholar]
  • 5.Martinez-Hackert E, Stock AM. The DNA-binding domain of OmpR: crystal structures of a winged helix transcription factor. Structure. 1997;5(1):109–124. doi: 10.1016/s0969-2126(97)00170-6 . [DOI] [PubMed] [Google Scholar]
  • 6.Alderwick LJ, Molle V, Kremer L, Cozzone AJ, Dafforn TR, Besra GS, et al. Molecular structure of EmbR, a response element of Ser/Thr kinase signaling in Mycobacterium tuberculosis. Proc Natl Acad Sci U S A. 2006;103(8):2558–2563. doi: 10.1073/pnas.0507766103 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Krause J, Handayani I, Blin K, Kulik A, Mast Y. Disclosing the Potential of the SARP-Type Regulator PapR2 for the Activation of Antibiotic Gene Clusters in Streptomycetes. Front Microbiol. 2020;11:225. Epub 2020/03/07. doi: 10.3389/fmicb.2020.00225 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Wietzorrek A, Bibb M. A novel family of proteins that regulates antibiotic production in streptomycetes appears to contain an OmpR-like DNA-binding fold. Mol Microbiol. 1997;25(6):1181–1184. doi: 10.1046/j.1365-2958.1997.5421903.x . [DOI] [PubMed] [Google Scholar]
  • 9.Tanaka A, Takano Y, Ohnishi Y, Horinouchi S. AfsR recruits RNA polymerase to the afsS promoter: a model for transcriptional activation by SARPs. J Mol Biol. 2007;369(2):322–333. Epub 2007/04/17. doi: 10.1016/j.jmb.2007.02.096 . [DOI] [PubMed] [Google Scholar]
  • 10.Lee PC, Umeyama T, Horinouchi S. afsS is a target of AfsR, a transcriptional factor with ATPase activity that globally controls secondary metabolism in Streptomyces coelicolor A3(2). Mol Microbiol. 2002;43(6):1413–1430. Epub 2002/04/16. doi: 10.1046/j.1365-2958.2002.02840.x . [DOI] [PubMed] [Google Scholar]
  • 11.Lian W, Jayapal KP, Charaniya S, Mehra S, Glod F, Kyung YS, et al. Genome-wide transcriptome analysis reveals that a pleiotropic antibiotic regulator, AfsS, modulates nutritional stress response in Streptomyces coelicolor A3(2). BMC Genomics. 2008;9:56. Epub 2008/01/31. doi: 10.1186/1471-2164-9-56 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Horinouchi S. AfsR as an integrator of signals that are sensed by multiple serine/threonine kinases in Streptomyces coelicolor A3(2). J Ind Microbiol Biotechnol. 2003;30(8):462–467. Epub 2003/07/29. doi: 10.1007/s10295-003-0063-z . [DOI] [PubMed] [Google Scholar]
  • 13.Santos-Beneit F, Rodríguez-García A, Martín JF. Overlapping binding of PhoP and AfsR to the promoter region of glnR in Streptomyces coelicolor. Microbiol Res. 2012;167(9):532–535. Epub 2012/04/03. doi: 10.1016/j.micres.2012.02.010 . [DOI] [PubMed] [Google Scholar]
  • 14.Martín JF, Liras P. The Balance Metabolism Safety Net: Integration of Stress Signals by Interacting Transcriptional Factors in Streptomyces and Related Actinobacteria. Front Microbiol. 2019;10:3120. Epub 20200122. doi: 10.3389/fmicb.2019.03120 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Sawai R, Suzuki A, Takano Y, Lee PC, Horinouchi S. Phosphorylation of AfsR by multiple serine/threonine kinases in Streptomyces coelicolor A3(2). Gene. 2004;334:53–61. doi: 10.1016/j.gene.2004.02.046 . [DOI] [PubMed] [Google Scholar]
  • 16.Browning DF, Busby SJ. Local and global regulation of transcription initiation in bacteria. Nat Rev Microbiol. 2016;14(10):638–650. doi: 10.1038/nrmicro.2016.103 . [DOI] [PubMed] [Google Scholar]
  • 17.Lee DJ, Minchin SD, Busby SJ. Activating transcription in bacteria. Annu Rev Microbiol. 2012;66:125–152. doi: 10.1146/annurev-micro-092611-150012 . [DOI] [PubMed] [Google Scholar]
  • 18.Smidova K, Zikova A, Pospisil J, Schwarz M, Bobek J, Vohradsky J. DNA mapping and kinetic modeling of the HrdB regulon in Streptomyces coelicolor. Nucleic Acids Res. 2019;47(2):621–633. doi: 10.1093/nar/gky1018 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Jeong Y, Kim JN, Kim MW, Bucca G, Cho S, Yoon YJ, et al. The dynamic transcriptional and translational landscape of the model antibiotic producer Streptomyces coelicolor A3(2). Nat Commun. 2016;7:11605. Epub 2016/06/03. doi: 10.1038/ncomms11605 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Santos-Beneit F, Rodriguez-Garcia A, Sola-Landa A, Martin JF. Cross-talk between two global regulators in Streptomyces: PhoP and AfsR interact in the control of afsS, pstS and phoRP transcription. Mol Microbiol. 2009;72(1):53–68. doi: 10.1111/j.1365-2958.2009.06624.x . [DOI] [PubMed] [Google Scholar]
  • 21.Santos-Beneit F, Rodriguez-Garcia A, Martin JF. Complex transcriptional control of the antibiotic regulator afsS in Streptomyces: PhoP and AfsR are overlapping, competitive activators. J Bacteriol. 2011;193(9):–2251. Epub 2011/03/08. doi: 10.1128/JB.01462-10 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Liu B, Hong C, Huang RK, Yu Z, Steitz TA. Structural basis of bacterial transcription activation. Science. 2017;358(6365):947–951. Epub 2017/11/18. doi: 10.1126/science.aao1923 . [DOI] [PubMed] [Google Scholar]
  • 23.Feng Y, Zhang Y, Ebright RH. Structural basis of transcription activation. Science. 2016;352(6291):1330–1333. doi: 10.1126/science.aaf4417 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Shi W, Jiang Y, Deng Y, Dong Z, Liu B. Visualization of two architectures in class-II CAP-dependent transcription activation. PLoS Biol. 2020;18(4):e3000706. Epub 2020/04/21. doi: 10.1371/journal.pbio.3000706 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Leipe DD, Koonin EV, Aravind L. STAND, a class of P-loop NTPases including animal and plant regulators of programmed cell death: multiple, complex domain architectures, unusual phyletic patterns, and evolution by horizontal gene transfer. J Mol Biol. 2004;343(1):1–28. Epub 2004/09/24. doi: 10.1016/j.jmb.2004.08.023 . [DOI] [PubMed] [Google Scholar]
  • 26.Dorstyn L, Akey CW, Kumar S. New insights into apoptosome structure and function. Cell Death Differ. 2018;25(7):1194–1208. Epub 2018/05/17. doi: 10.1038/s41418-017-0025-z . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Urbach JM, Ausubel FM. The NBS-LRR architectures of plant R-proteins and metazoan NLRs evolved in independent events. Proc Natl Acad Sci U S A. 2017;114(5):1063–1068. Epub 2017/01/18. doi: 10.1073/pnas.1619730114 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Alcantara C, Sarmiento-Rubiano LA, Monedero V, Deutscher J, Perez-Martinez G, Yebra MJ. Regulation of Lactobacillus casei sorbitol utilization genes requires DNA-binding transcriptional activator GutR and the conserved protein GutM. Appl Environ Microbiol. 2008;74(18):5731–5740. Epub 2008/08/05. doi: 10.1128/AEM.00230-08 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Liu P, Danot O, Richet E. A dual role for the inducer in signalling by MalT, a signal transduction ATPase with numerous domains (STAND). Mol Microbiol. 2013;90(6):1309–1323. Epub 2013/10/19. doi: 10.1111/mmi.12434 . [DOI] [PubMed] [Google Scholar]
  • 30.Shakeri R, Kheirollahi A, Davoodi J. Apaf-1: Regulation and function in cell death. Biochimie. 2017;135:111–125. Epub 2017/02/14. doi: 10.1016/j.biochi.2017.02.001 . [DOI] [PubMed] [Google Scholar]
  • 31.Chou WC, Jha S, Linhoff MW, Ting JP. The NLR gene family: from discovery to present day. Nat Rev Immunol. 2023. Epub 2023/03/28. doi: 10.1038/s41577-023-00849-x . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Wu Y, Sun Y, Richet E, Han Z, Chai J. Structural basis for negative regulation of the Escherichia coli maltose system. Nat Commun. 2023;14(1):4925. Epub 2023/08/16. doi: 10.1038/s41467-023-40447-y . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Lisa MN, Cvirkaite-Krupovic V, Richet E, Andre-Leroux G, Alzari PM, Haouz A, et al. Double autoinhibition mechanism of signal transduction ATPases with numerous domains (STAND) with a tetratricopeptide repeat sensor. Nucleic Acids Res. 2019;47(7):3795–3810. Epub 2019/02/23. doi: 10.1093/nar/gkz112 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Boyaci H, Chen J, Jansen R, Darst SA, Campbell EA. Structures of an RNA polymerase promoter melting intermediate elucidate DNA unwinding. Nature. 2019;565(7739):382–385. doi: 10.1038/s41586-018-0840-5 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Lilic M, Darst SA, Campbell EA. Structural basis of transcriptional activation by the Mycobacterium tuberculosis intrinsic antibiotic-resistance transcription factor WhiB7. Mol Cell. 2021;(14):2875–86.e5. Epub 20210624. doi: 10.1016/j.molcel.2021.05.017 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Jensen D, Manzano AR, Rammohan J, Stallings CL, Galburt EA. CarD and RbpA modify the kinetics of initial transcription and slow promoter escape of the Mycobacterium tuberculosis RNA polymerase. Nucleic Acids Res. 2019;47(13):6685–6698. doi: 10.1093/nar/gkz449 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Yang X, Wang Y, Liu G, Deng Z, Lin S, Zheng J. Structural basis of Streptomyces transcription activation by zinc uptake regulator. Nucleic Acids Res. 2022;50(14):8363–8376. Epub 2022/07/25. doi: 10.1093/nar/gkac627 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lilic M, Chen J, Boyaci H, Braffman N, Hubin EA, Herrmann J, et al. The antibiotic sorangicin A inhibits promoter DNA unwinding in a Mycobacterium tuberculosis rifampicin-resistant RNA polymerase. Proc Natl Acad Sci U S A. 2020;117(48):30423–30432. doi: 10.1073/pnas.2013706117 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Hubin EA, Lilic M, Darst SA, Campbell EA. Structural insights into the mycobacteria transcription initiation complex from analysis of X-ray crystal structures. Nat Commun. 2017;8:16072. doi: 10.1038/ncomms16072 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Muller AU, Kummer E, Schilling CM, Ban N, Weber-Ban E. Transcriptional control of mycobacterial DNA damage response by sigma adaptation. Sci Adv. 2021;7(49):eabl4064. Epub 2021/12/02. doi: 10.1126/sciadv.abl4064 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Bae B, Chen J, Davis E, Leon K, Darst SA, Campbell EA. CarD uses a minor groove wedge mechanism to stabilize the RNA polymerase open promoter complex. Elife. 2015;4:e08505. doi: 10.7554/eLife.08505 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Hubin EA, Fay A, Xu C, Bean JM, Saecker RM, Glickman MS, et al. Structure and function of the mycobacterial transcription initiation complex with the essential regulator RbpA. Elife. 2017;6:e22520. doi: 10.7554/eLife.22520 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Shin Y, Qayyum MZ, Pupov D, Esyunina D, Kulbachinskiy A, Murakami KS. Structural basis of ribosomal RNA transcription regulation. Nat Commun. 2021;12(1):528. Epub 2021/01/24. doi: 10.1038/s41467-020-20776-y . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Marquenet E, Richet E. Conserved motifs involved in ATP hydrolysis by MalT, a signal transduction ATPase with numerous domains from Escherichia coli. J Bacteriol. 2010;192(19):5181–5191. Epub 2010/08/10. doi: 10.1128/JB.00522-10 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Marquenet E, Richet E. How integration of positive and negative regulatory signals by a STAND signaling protein depends on ATP hydrolysis. Mol Cell. 2007;28(2):187–199. Epub 2007/10/30. doi: 10.1016/j.molcel.2007.08.014 . [DOI] [PubMed] [Google Scholar]
  • 46.Blin K, Pascal Andreu V, de Los Santos ELC, Del Carratore F, Lee SY, Medema MH, et al. The antiSMASH database version 2: a comprehensive resource on secondary metabolite biosynthetic gene clusters. Nucleic Acids Res. 2019;47(D1):D625–d30. doi: 10.1093/nar/gky1060 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Garg RP, Parry RJ. Regulation of valanimycin biosynthesis in Streptomyces viridifaciens: characterization of VlmI as a Streptomyces antibiotic regulatory protein (SARP). Microbiology (Reading). 2010;156(Pt 2):472–83. Epub 20091105. doi: 10.1099/mic.0.033167-0 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Ye S, Braña AF, González-Sabín J, Morís F, Olano C, Salas JA, et al. New Insights into the Biosynthesis Pathway of Polyketide Alkaloid Argimycins P in Streptomyces argillaceus. Front Microbiol. 2018;9:252. Epub 20180216. doi: 10.3389/fmicb.2018.00252 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Sun L, Zeng J, Cui P, Wang W, Yu D, Zhan J. Manipulation of two regulatory genes for efficient production of chromomycins in Streptomyces reseiscleroticus. J Biol Eng. 2018;12:9. Epub 2018/07/07. doi: 10.1186/s13036-018-0103-x . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Shi J, Wang L, Wen A, Wang F, Zhang Y, Yu L, et al. Structural basis of three different transcription activation strategies adopted by a single regulator SoxS. Nucleic Acids Res. 2022;50(19):11359–11373. doi: 10.1093/nar/gkac898 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Shi J, Wang F, Li F, Wang L, Xiong Y, Wen A, et al. Structural basis of transcription activation by Rob, a pleiotropic AraC/XylS family regulator. Nucleic Acids Res. 2022;50(10):5974–5987. doi: 10.1093/nar/gkac433 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Cortes T, Schubert OT, Rose G, Arnvig KB, Comas I, Aebersold R, et al. Genome-wide mapping of transcriptional start sites defines an extensive leaderless transcriptome in Mycobacterium tuberculosis. Cell Rep. 2013;5(4):1121–1131. doi: 10.1016/j.celrep.2013.10.031 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Poon KK, Chu JC, Wong SL. Roles of glucitol in the GutR-mediated transcription activation process in Bacillus subtilis: glucitol induces GutR to change its conformation and to bind ATP. J Biol Chem. 2001;276(32):29819–29825. Epub 2001/06/08. doi: 10.1074/jbc.M100905200 . [DOI] [PubMed] [Google Scholar]
  • 54.Hubin EA, Tabib-Salazar A, Humphrey LJ, Flack JE, Olinares PD, Darst SA, et al. Structural, functional, and genetic analyses of the actinobacterial transcription factor RbpA. Proc Natl Acad Sci U S A. 2015;112(23):7171–7176. doi: 10.1073/pnas.1504942112 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sudalaiyadum Perumal A, Vishwakarma RK, Hu Y, Morichaud Z, Brodolin K. RbpA relaxes promoter selectivity of M. tuberculosis RNA polymerase. Nucleic Acids Res. 2018;46(19):10106–10118. doi: 10.1093/nar/gky714 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Punjani A, Rubinstein JL, Fleet DJ, Brubaker MA. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat Methods. 2017;14(3):290–296. doi: 10.1038/nmeth.4169 . [DOI] [PubMed] [Google Scholar]
  • 57.Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596(7873):583–589. doi: 10.1038/s41586-021-03819-2 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Casanal A, Lohkamp B, Emsley P. Current developments in Coot for macromolecular model building of Electron Cryo-microscopy and Crystallographic Data. Protein Sci. 2020;29(4):1069–1078. doi: 10.1002/pro.3791 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Pettersen EF, Goddard TD, Huang CC, Meng EC, Couch GS, Croll TI, et al. UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci. 2021;30(1):70–82. doi: 10.1002/pro.3943 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Liebschner D, Afonine PV, Baker ML, Bunkoczi G, Chen VB, Croll TI, et al. Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr D Struct Biol. 2019;75(Pt 10):861–877. doi: 10.1107/S2059798319011471 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Williams CJ, Headd JJ, Moriarty NW, Prisant MG, Videau LL, Deis LN, et al. MolProbity: More and better reference data for improved all-atom structure validation. Protein Sci. 2018;27(1):293–315. doi: 10.1002/pro.3330 . [DOI] [PMC free article] [PubMed] [Google Scholar]

Decision Letter 0

Richard Hodge

29 Sep 2023

Dear Dr Zheng,

Thank you for submitting your manuscript entitled "Structural and functional dissection of AfsR, a transcription activator of Streptomyces antibiotic regulatory protein family" for consideration as a Research Article by PLOS Biology.

Your manuscript has now been evaluated by the PLOS Biology editorial staff, as well as by an academic editor with relevant expertise, and I am writing to let you know that we would like to send your submission out for external peer review.

However, before we can send your manuscript to reviewers, we need you to complete your submission by providing the metadata that is required for full assessment. To this end, please login to Editorial Manager where you will find the paper in the 'Submissions Needing Revisions' folder on your homepage. Please click 'Revise Submission' from the Action Links and complete all additional questions in the submission questionnaire.

Once your full submission is complete, your paper will undergo a series of checks in preparation for peer review. After your manuscript has passed the checks it will be sent out for review. To provide the metadata for your submission, please Login to Editorial Manager (https://www.editorialmanager.com/pbiology) within two working days, i.e. by Oct 01 2023 11:59PM.

If your manuscript has been previously peer-reviewed at another journal, PLOS Biology is willing to work with those reviews in order to avoid re-starting the process. Submission of the previous reviews is entirely optional and our ability to use them effectively will depend on the willingness of the previous journal to confirm the content of the reports and share the reviewer identities. Please note that we reserve the right to invite additional reviewers if we consider that additional/independent reviewers are needed, although we aim to avoid this as far as possible. In our experience, working with previous reviews does save time.

If you would like us to consider previous reviewer reports, please edit your cover letter to let us know and include the name of the journal where the work was previously considered and the manuscript ID it was given. In addition, please upload a response to the reviews as a 'Prior Peer Review' file type, which should include the reports in full and a point-by-point reply detailing how you have or plan to address the reviewers' concerns.

During the process of completing your manuscript submission, you will be invited to opt-in to posting your pre-review manuscript as a bioRxiv preprint. Visit http://journals.plos.org/plosbiology/s/preprints for full details. If you consent to posting your current manuscript as a preprint, please upload a single Preprint PDF.

Feel free to email us at plosbiology@plos.org if you have any queries relating to your submission.

Kind regards,

Richard

Richard Hodge, PhD

Senior Editor, PLOS Biology

rhodge@plos.org

PLOS

Empowering researchers to transform science

Carlyle House, Carlyle Road, Cambridge, CB4 3DN, United Kingdom

ORCiD I plosbio.org I @PLOSBiology I Blog

California (U.S.) corporation #C2354500, based in San Francisco

Decision Letter 1

Richard Hodge

11 Dec 2023

Dear Dr Zheng,

Thank you for your patience while your manuscript "Structural and functional dissection of AfsR, a transcription activator of Streptomyces antibiotic regulatory protein family" went through peer-review at PLOS Biology. Please accept my sincere apologies for the long delays that you have experienced during the peer review process. Your manuscript has now been evaluated by the PLOS Biology editors, an Academic Editor with relevant expertise, and by two independent reviewers.

In light of the reviews, which you will find at the end of this email, we are pleased to offer you the opportunity to address the comments from the reviewers in a revision that we anticipate should not take you very long. Reviewer #2 notes that a control assay for the in vitro transcription assays should be provided to show that the effects are specific to transcription initiation at the afsS promoter. We will then assess your revised manuscript and your response to the reviewers' comments with our Academic Editor aiming to avoid further rounds of peer-review, although might need to consult with the reviewers, depending on the nature of the revisions.

We expect to receive your revised manuscript within 3 months. Please email us (plosbiology@plos.org) if you have any questions or concerns, or would like to request an extension.

At this stage, your manuscript remains formally under active consideration at our journal; please notify us by email if you do not intend to submit a revision so that we withdraw the manuscript.

**IMPORTANT - SUBMITTING YOUR REVISION**

Your revisions should address the specific points made by each reviewer. Please submit the following files along with your revised manuscript:

1. A 'Response to Reviewers' file - this should detail your responses to the editorial requests, present a point-by-point response to all of the reviewers' comments, and indicate the changes made to the manuscript.

*NOTE: In your point-by-point response to the reviewers, please provide the full context of each review. Do not selectively quote paragraphs or sentences to reply to. The entire set of reviewer comments should be present in full and each specific point should be responded to individually.

You should also cite any additional relevant literature that has been published since the original submission and mention any additional citations in your response.

2. In addition to a clean copy of the manuscript, please also upload a 'track-changes' version of your manuscript that specifies the edits made. This should be uploaded as a "Revised Article with Changes Highlighted " file type.

*Resubmission Checklist*

When you are ready to resubmit your revised manuscript, please refer to this resubmission checklist: https://plos.io/Biology_Checklist

To submit a revised version of your manuscript, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' where you will find your submission record.

Please make sure to read the following important policies and guidelines while preparing your revision:

*Published Peer Review*

Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

*PLOS Data Policy*

Please note that as a condition of publication PLOS' data policy (http://journals.plos.org/plosbiology/s/data-availability) requires that you make available all data used to draw the conclusions arrived at in your manuscript. If you have not already done so, you must include any data used in your manuscript either in appropriate repositories, within the body of the manuscript, or as supporting information (N.B. this includes any numerical values that were used to generate graphs, histograms etc.). For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5

*Blot and Gel Data Policy*

We require the original, uncropped and minimally adjusted images supporting all blot and gel results reported in an article's figures or Supporting Information files. We will require these files before a manuscript can be accepted so please prepare them now, if you have not already uploaded them. Please carefully read our guidelines for how to prepare and upload this data: https://journals.plos.org/plosbiology/s/figures#loc-blot-and-gel-reporting-requirements

*Protocols deposition*

To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

Thank you again for your submission to our journal. We hope that our editorial process has been constructive thus far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Richard

Richard Hodge, PhD

Senior Editor, PLOS Biology

rhodge@plos.org

----------------------------------------------------------------

REVIEWS:

Reviewer #1: In this paper, Wang and Zheng et. al. characterized the structure and mechanism of AfsR stimulated transcription. They determined high-resolution structure of Sarp domain of AfsR as well as full-length AfsR bound to transcription initiation complex (TIC), which together with biochemical assays illustrated how AfsR interacts with and stimulates transcription. In addition, biochemical investigation suggested a multi-layer regulation scheme of AfsR, which are duel controlled by ATP binding and phosphorylation. The paper presents significant new insights to Sarp-domain mediated transcription activation and could be general interests to broad audience. I only have several minor suggestions.

Minor:

1. Scale bars are needed for the 2D classification images in Fig S2&10,.

2. Local densities of SARP domain and C-terminal domains of AfsR, possibly in FigS2&10 or in additional figures, are needed to show the quality of the maps, especially at the DNA binding region and C-terminal domains.

3. A minor suggestion, possibly color the protein complexes and DNAs with different color in Fig S3 will be helpful to better illustrate the architecture of TICs.

4. In the TIC structure with SARP domain (Fig 2), the author claims the two protomer are nearly identical. A structural comparison will be helpful. In addition, protomer 1 has sequence specific DNA interactions and the promotor. How about protomer 2? Does it help with sequence recognition and transcription activation? In addition, do both protomers exert similar degree of DNA bending?

5. A scheme model to summarize the ATP dependent and phosphorylation dependent regulation of AfsR will be helpful.

Reviewer #2: This manuscript reveals some key aspects of how AfsR, a member of the Streptomyces antibiotic regulatory protein family binds at a target promoter to activate transcription, including some very interesting and novel contacts with various subunits of RNA polymerase. The work is of particular interest to anyone working with the regulation and expression of antibiotic biosynthetic clusters since AfsR is related to many key activators, and the structural model presented is likely to be broadly applicable to this family. The works is also of wider interest to the transcription field, showing another example of how flexibility in key RNAP subunits alpha-CTD and sigma can accommodate varied scaffolds that lead to activation. The work does not explore in any detail mechanistic questions of activation and the SARP role in the sequential process of initiation, but this is beyond the scope. Overall the presentation is very good and it reads very well. I have a few specific comments.

1. The CACTGT -10 is close but not consensus since a T at the first position is most common. In fact it is argued in the paper that SARPs can compensate for poor -35 elements, however since most promoters have poor -35 elements, it is just as valid to suggest that at specific SARP-regulated promoters, the promoters lack activity in the absence of the SARP due to other "poor" promoter elements e.g. the -10 element.

2. Can the authors show that the in vitro transcription assays used are specific to transcription initiation at the afsS promoter rather than an increase in general activity of the RNAP as seen with end-to-end transcription on the DNA template. The assay with a control promoter would be useful here.

3. Line 106. I think the authors mean disinhibition by ATP>

4. Line 114. The Fig 1 C/D are mislabelled.

5. Line 115. It should be made clear that thre SARP domain was used to build the TIC.

6. Line 123. In the context of activation in Streptomyces it is important to know whether the SARP can activate transcription in the absence of RbpA and CarD. It seems an obvious omission and the paper would benefit from the inclusion.

7. The DNA scaffold used to form the TIC includes dsDNA around the -10 region, so it would be important to emphasise that the DNA is melted in the final TIC - ie that the complex has isomerized to the open complex.

8. Line 292 and associated figure. The effect of ATP on limited proteolysis patterns of dephosphorylated AfsR vs phosphorylated AfsR is not that convincing and key differences could be highlighted. From my interpretation, the bigger difference appears to be between dephos-AfsR vs phos-AfsR rather than the effect of ATP.

9. Line 295 It is proposed that ATP hydrolysis might explain why the phosphorylated AfsR is not activated by ATP binding - what happens when a non-hydrolysable analogue of ATP is used?

10. Line 331 The phrase that SARP functions as an adaptor to physically bond the -35 and sigma R4 is a bit ambiguous and might suggest that sigma makes normal contacts with -35 that are held in place by SARP. In fact the SARP provides an alternative docking point for R4 meaning that a -35 region is not required.

Decision Letter 2

Richard Hodge

22 Jan 2024

Dear Dr Zheng,

Thank you for your patience while we considered your revised manuscript "Structural and functional dissection of AfsR, a transcription activator of Streptomyces antibiotic regulatory protein family" for publication as a Research Article at PLOS Biology. This revised version of your manuscript has been evaluated by the PLOS Biology editors and the Academic Editor.

Based on our Academic Editor's assessment of your revision, we are likely to accept this manuscript for publication, provided you satisfactorily address the following data and other policy-related requests.

IMPORTANT - please attend to the following:

a) Please change your Title to the following: "Structural and functional characterization of AfsR, a SARP family transcriptional activator of antibiotic biosynthesis in Streptomyces"

b) Please address my Data Policy requests below; specifically, we need you to supply the numerical values underlying Figs 1CDE, 3F, 4C, 5ACEF, S1BC, S2, S3D, S9A, S11ACDEF, S12AB, S13, S14D, either as a supplementary data file or as a permanent DOI’d deposition. I note that you already have a supplementary data file, "SourceData.xlsx," which seems to include some of this data. Please clarify this, and include any additional data required to support the Figures.

c) Please cite the location of the data clearly in all relevant main and supplementary Figure legends, e.g. re-name "SourceData.xlsx" as "S1_Data/xlsx" and write “The data underlying this Figure can be found in S1 Data” in each relevant Figure legend.

d) Please make any custom code available, either as a supplementary file or as part of an online deposition.

As you address these items, please take this last chance to review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the cover letter that accompanies your revised manuscript.

We expect to receive your revised manuscript within two weeks.

To submit your revision, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' to find your submission record. Your revised submission must include the following:

- a cover letter that should detail your responses to any editorial requests, if applicable, and whether changes have been made to the reference list

- a Response to Reviewers file that provides a detailed response to the reviewers' comments (if applicable, if not applicable please do not delete your existing 'Response to Reviewers' file.)

- a track-changes file indicating any changes that you have made to the manuscript.

NOTE: If Supporting Information files are included with your article, note that these are not copyedited and will be published as they are submitted. Please ensure that these files are legible and of high quality (at least 300 dpi) in an easily accessible file format. For this reason, please be aware that any references listed in an SI file will not be indexed. For more information, see our Supporting Information guidelines:

https://journals.plos.org/plosbiology/s/supporting-information

*Published Peer Review History*

Please note that you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

*Press*

Should you, your institution's press office or the journal office choose to press release your paper, please ensure you have opted out of Early Article Posting on the submission form. We ask that you notify us as soon as possible if you or your institution is planning to press release the article.

*Protocols deposition*

To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

Please do not hesitate to contact me should you have any questions.

Sincerely,

Roli Roberts

Roland G Roberts PhD

Senior Editor

rroberts@plos.org

PLOS Biology

on behalf of

Richard Hodge,

Senior Editor,

rhodge@plos.org,

PLOS Biology

------------------------------------------------------------------------

DATA POLICY:

You may be aware of the PLOS Data Policy, which requires that all data be made available without restriction: http://journals.plos.org/plosbiology/s/data-availability. For more information, please also see this editorial: http://dx.doi.org/10.1371/journal.pbio.1001797

Note that we do not require all raw data. Rather, we ask that all individual quantitative observations that underlie the data summarized in the figures and results of your paper be made available in one of the following forms:

1) Supplementary files (e.g., excel). Please ensure that all data files are uploaded as 'Supporting Information' and are invariably referred to (in the manuscript, figure legends, and the Description field when uploading your files) using the following format verbatim: S1 Data, S2 Data, etc. Multiple panels of a single or even several figures can be included as multiple sheets in one excel file that is saved using exactly the following convention: S1_Data.xlsx (using an underscore).

2) Deposition in a publicly available repository. Please also provide the accession code or a reviewer link so that we may view your data before publication.

Regardless of the method selected, please ensure that you provide the individual numerical values that underlie the summary data displayed in the following figure panels as they are essential for readers to assess your analysis and to reproduce it: Figs 1CDE, 3F, 4C, 5ACEF, S1BC, S2, S3D, S9A, S11ACDEF, S12AB, S13, S14D. NOTE: the numerical data provided should include all replicates AND the way in which the plotted mean and errors were derived (it should not present only the mean/average values).

IMPORTANT: Please also ensure that figure legends in your manuscript include information on where the underlying data can be found, and ensure your supplemental data file/s has a legend.

Please ensure that your Data Statement in the submission system accurately describes where your data can be found.

------------------------------------------------------------------------

CODE POLICY

Per journal policy, as the code that you have generated is important to support the conclusions of your manuscript, we require that you make it available without restrictions upon publication. Please ensure that the code is sufficiently well documented and reusable, and that your Data Statement in the Editorial Manager submission system accurately describes where your code can be found.

------------------------------------------------------------------------

DATA NOT SHOWN?

- Please note that per journal policy, we do not allow the mention of "data not shown", "personal communication", "manuscript in preparation" or other references to data that is not publicly available or contained within this manuscript. Please either remove mention of these data or provide figures presenting the results and the data underlying the figure(s).

------------------------------------------------------------------------

Decision Letter 3

Richard Hodge

29 Jan 2024

Dear Dr Zheng,

Thank you for the submission of your revised Research Article "Structural and functional characterization of AfsR, a SARP family transcriptional activator of antibiotic biosynthesis in Streptomyces" for publication in PLOS Biology. On behalf of my colleagues and the Academic Editor, Yunsun Nam, I'm pleased to say that we can in principle accept your manuscript for publication, provided you address any remaining formatting and reporting issues. These will be detailed in an email you should receive within 2-3 business days from our colleagues in the journal operations team; no action is required from you until then. Please note that we will not be able to formally accept your manuscript and schedule it for publication until you have completed any requested changes.

Please take a minute to log into Editorial Manager at http://www.editorialmanager.com/pbiology/, click the "Update My Information" link at the top of the page, and update your user information to ensure an efficient production process.

PRESS: We frequently collaborate with press offices. If your institution or institutions have a press office, please notify them about your upcoming paper at this point, to enable them to help maximise its impact. If the press office is planning to promote your findings, we would be grateful if they could coordinate with biologypress@plos.org. If you have previously opted in to the early version process, we ask that you notify us immediately of any press plans so that we may opt out on your behalf.

We also ask that you take this opportunity to read our Embargo Policy regarding the discussion, promotion and media coverage of work that is yet to be published by PLOS. As your manuscript is not yet published, it is bound by the conditions of our Embargo Policy. Please be aware that this policy is in place both to ensure that any press coverage of your article is fully substantiated and to provide a direct link between such coverage and the published work. For full details of our Embargo Policy, please visit http://www.plos.org/about/media-inquiries/embargo-policy/.

Thank you again for choosing PLOS Biology for publication and supporting Open Access publishing. We look forward to publishing your study. 

Sincerely, 

Roli Roberts

Roland Roberts PhD

Senior Editor

PLOS Biology

rroberts@plos.org

on behalf of

Richard Hodge, PhD,

Senior Editor

PLOS Biology

rhodge@plos.org

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    S1 Fig. In vitro assays of the Streptomyces coelicolor AfsR SARP.

    (A) Core promoter sequences used for in vitro assays. The mutations introduced into the afs box are highlighted. The mutated target sites contained mutations in upstream repeat (M1), the downstream repeat (M2), or both repeats (M1M2). The actII-4 promoter was used as a control. (B) Fluorescence polarization assays of the SARP with mutant afs box (M1, M2, and M1M2). Error bars represent mean ± SEM of n = 3 experiments. (C) In vitro MangoIII-based transcription assays with or without 500 nM SARP in the absence of RbpA and CarD. Error bars represent mean ± SEM of n = 3 experiments. The data underlying B and C can be found in S1 Data.

    (TIF)

    pbio.3002528.s001.tif (1.3MB, tif)
    S2 Fig. RbpA and CarD do not influence the transcriptional activation capability of SARP.

    In vitro transcription assays with or without RbpA and CarD on the afsS promoter (left) and transcription assays with or without 500 nM SARP in the presence of RbpA and CarD (right). The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3.

    (TIF)

    pbio.3002528.s002.tif (270.8KB, tif)
    S3 Fig. Single-particle cryo-EM analysis of SARP-TIC.

    (A) A motion-corrected image (scale bar: 40 nm). (B) 2D classes (scale bar: 100 Å). (C) Data processing pipeline for the dataset of SARP-TIC. The final cryo-EM map of the SARP-TIC was reconstructed using a total of 95,223 single particles and refined to a nominal resolution of 3.35 Å. Local refinement focused on the SARP region generated a 3.77-Å-resolution map. (D) Validation of cryo-EM structural models. Map vs. model FSCs was generated by Phenix (Version 1.19.2). The model-map resolution for the atomic model and cryo-EM map at FSC = 0.5 cutoff was indicated in the figure and reported in S1 Table. The data underlying this figure can be found in S1 Data.

    (TIF)

    pbio.3002528.s003.tif (2.2MB, tif)
    S4 Fig. Comparisons of RNAP.

    (A) S. coelicolor SARP-TIC. (B) S. coelicolor RNAPσHrdB-Zur-DNA (PDB ID: 7X75). (C) Mycobacterium tuberculosis RNAP-promoter open complex (PDB ID: 6VVY). (D) Mycobacterium smegmatis TIC (PDB ID: 5VI5).

    (TIF)

    S5 Fig. Comparison of 2 SARP protomers in SARP-TIC.

    Overall conformations of the 2 protomers are essentially the same with an overall rmsd of 0.8 Å.

    (TIF)

    pbio.3002528.s005.tif (3.7MB, tif)
    S6 Fig. Comparison of SARP with EmbR (PDB ID: 2FEZ) comprising an additional C-terminal FHA domain.

    (TIF)

    pbio.3002528.s006.tif (2.4MB, tif)
    S7 Fig. The binding of 2 SARP protomers in afs box results in a 20° bend of the helical axis of the dsDNA.

    (TIF)

    pbio.3002528.s007.tif (1.9MB, tif)
    S8 Fig. Interactions of the downstream SARP with afs box.

    (A) Downstream SARP residues involved in interactions with the DNA backbone. (B) Contacts of downstream SARP with specific nucleotides. The residues R92 and T88 make hydrogen bonds (shown as yellow dashed lines) with T-27(nt) and A-25(t), respectively.

    (TIF)

    pbio.3002528.s008.tif (2.8MB, tif)
    S9 Fig. In vitro assays of SARP mutants.

    (A) In vitro transcription assays of SARP T63A, E77A, E246A, L247A, and V249A mutants. No obvious difference was observed between the T63A, E77A, E246A, L247A, V249A mutants, and the wild-type protein. n.s. means no significance. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3. (B) SDS-PAGE of E176A, L211A, and L243A mutants. Inclusion bodies were formed when these mutants were expressed under the same conditions as the wild-type protein. P represents precipitation while S represents supernatant. The original gel images can be found in S1 Raw Images.

    (TIF)

    pbio.3002528.s009.tif (720.5KB, tif)
    S10 Fig. Purification and assembly of protein complexes.

    (A) Assembly of S. coelicolor RNAP-σHrdB. The peak of RNAP-σHrdB was analyzed by SDS-PAGE. (B) Purification of AfsR. (C) Purification of ΔTPR. (D) Purification of SARP. (E) Purification of RbpA. (F) Purification of CarD. (G) Purification of AfsKΔC. (H) Assembly of AfsRT337A-TIC in the presence of 1 mM ATPγS. The protein compositions in the dotted line boxed fractions are shown in the SDS-PAGE. The original gel images can be found in S1 Raw Images.

    (TIF)

    S11 Fig. Phosphorylation modulates AfsR transcriptional activation.

    (A) LC-MS/MS analysis showing that S22, T337, S391, T506, and S953 are phosphorylated in AfsR by AfsKΔC. The lowercase letter in the peptide sequence indicates phosphorylated residue. “b” and “y” denote peptide fragment ions retaining charges at the N and C terminus, respectively. The subscript numbers indicate their positions in the identified peptide. (B) Sequence alignment of Streptomyces AfsR family members highlighting the consensuses sequences neighboring phosphorylation sites S22, T337, S391, T506, and S953. Orange boxes represent phosphorylation sites. (C) Transcription assays with increasing concentrations of phosphorylated AfsR by AfsKΔC (phos-AfsR) and untreated AfsR. (D) Transcription assays of 500 nM AfsR mutants mimicking dephosphorylation (T337A) and phosphorylation (S22E, T337E, T506E, S953E, and S391E/T337E/T506E/S953E (4E)). Data are presented as mean ± SEM from 3 independent assays. n.s. means no significance; *P < 0.05; **P < 0.01 in comparison with the wild-type AfsR. (E) Representative SEC assay of AfsRT337A and AfsR4E. (F) Fluorescence polarization assay of AfsR4E and dephosphorylated AfsR (dephos-AfsR) with afs box. The concentration of afs box was 10 nM. Error bars represent mean ± SEM of n = 3 experiments. The data underlying A, C, D, E, and F are provided in S1 Data.

    (TIF)

    pbio.3002528.s011.tif (1.1MB, tif)
    S12 Fig. Oligomerization state and DNA-binding capacity of dephosphorylated AfsR are not affected by ATP binding.

    (A) SEC assay of dephosphorylated AfsR in the presence (solid line) or absence (dashed line) of 1 mM ATP. The data underlying this figure can be found in S1 Data. (B) Fluorescence polarization assay of dephosphorylated AfsR with afs box DNA in the presence or absence of 1 mM ATP. The concentration of afs box DNA was 10 nM. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3.

    (TIF)

    pbio.3002528.s012.tif (538.3KB, tif)
    S13 Fig. ATPγS activates phosphorylated AfsR and AfsR4E.

    Transcription assays involving 500 nM phosphorylated AfsR (phos-AfsR) or AfsR4E, with and without preincubation with 1 mM ATP or ATPγS. CK represents the control group without the addition of AfsR. The data underlying this figure can be found in S1 Data; error bars, SEM; n = 3.

    (TIF)

    pbio.3002528.s013.tif (87.2KB, tif)
    S14 Fig. Single-particle cryo-EM analysis of AfsRT337A-TIC.

    (A) Motion-corrected images (scale bar: 50 nm). (B) 2D classes (scale bar: 100 Å). (C) Data processing pipeline for the dataset of AfsRT337A-TAC. The final cryo-EM map of the AfsRT337A-TIC was reconstructed using a total of 45,208 single particles and refined to a nominal resolution of 3.63 Å. Local refinement focused on the SARP region generated a 5.10-Å-resolution map. Local refinement focused on the AfsR region generated a 8.61-Å-resolution map. (D) Validation of cryo-EM structural models. Map vs. model FSCs was generated by Phenix (Version 1.19.2). The model-map resolution for the atomic model and cryo-EM map at FSC = 0.5 cutoff was indicated in the figure and reported in S1 Table. The data underlying this figure can be found in S1 Data.

    (TIF)

    pbio.3002528.s014.tif (2.2MB, tif)
    S15 Fig. Two views of cryo-EM map and structural model of AfsRT337A-TIC.

    The map was generated by merging the consensus map of the full AfsR -TIC and the focused maps of the AfsR. The initial model of RNAP and AfsR was generated from SARP-TIC and AlphaFold, respectively, and fitted into the map using ChimeraX.

    (TIF)

    pbio.3002528.s015.tif (8.3MB, tif)
    S16 Fig. Interactions of regulators with σR4 and −35 element.

    (A) SARP. (B) Class I CAP-transcription activation complex (TAC) (PDB ID:6B6H). (C) Class II TAP-TAC (PDB ID: 5I2D). (D) SoxS (PDB ID: 7W5W). (E) RobR (PDB ID: 7VWZ). (F) PafBC (PDB ID: 7P5X). The key residues Arg and Glu supposed to contact DNA are shown as sticks.

    (TIF)

    pbio.3002528.s016.tif (8.8MB, tif)
    S17 Fig. Scheme model summarizing the ATP-dependent and phosphorylation-dependent regulation of AfsR.

    Unmodified full-length AfsR is autoinhibited by its C-terminal domains. AfsR is activated by ATP binding and significantly stimulates the production of transcripts. The phosphorylation of AfsR results in increased ATPase activity, potentially counteracting the effects of ATP binding. Phosphorylated AfsR forms more oligomers and shows slightly higher transcriptional activation compared to the unmodified form.

    (TIF)

    pbio.3002528.s017.tif (1.1MB, tif)
    S1 Table. Cryo-EM data collection, refinement, and validation statistics.

    (XLSX)

    pbio.3002528.s018.xlsx (11.8KB, xlsx)
    S2 Table. Phosphorylated peptides in AfsR by mass spectrometric.

    (XLSX)

    pbio.3002528.s019.xlsx (10.1KB, xlsx)
    S3 Table. List of primer sequences used in this study.

    (XLSX)

    pbio.3002528.s020.xlsx (11.5KB, xlsx)
    S1 Data. Numerical values used to generate graphs.

    (XLSX)

    S1 Raw Images. The original gel images contained in the figures.

    (PDF)

    pbio.3002528.s022.pdf (13.9MB, pdf)
    Attachment

    Submitted filename: Response to Reviewers.docx

    pbio.3002528.s023.docx (34.2KB, docx)
    Attachment

    Submitted filename: Response to Reviewers.docx

    pbio.3002528.s024.docx (34.2KB, docx)

    Data Availability Statement

    Atomic coordinate has been deposited in PDB with accession numbers 8HVR (SARP-TIC) and 8JKE (AfsRT337A-TIC). Four cryo-EM density maps have been deposited in Electron Microscopy Data Bank with accession number EMD-35047 (SARP-TIC), EMD-36370 (AfsRT337A-TIC), EMD-35046 (SARP region from SARP-TIC) and EMD-36369 (SARP region from AfsRT337A-TIC). All data underlying the main and supplementary figures can be found at supplementary S1 Data and S1 Raw Images.


    Articles from PLOS Biology are provided here courtesy of PLOS

    RESOURCES