Abstract
Legionella pneumophila, an intracellular bacterial pathogen, injects effector proteins into host cells to manipulate cellular processes and promote its survival and proliferation. Here, we reveal a unique mechanism by which the Legionella effector PieF perturbs host mRNA decay by targeting the human CCR4–NOT deadenylase complex. High-resolution cryo-electron microscopy structures and biochemical analyses reveal that PieF binds with nanomolar affinity to the NOT7 and NOT8 catalytic subunits of CCR4–NOT, obstructing RNA access and displacing a catalytic Mg²⁺ ion from the active site. Additionally, PieF prevents NOT7/8 from associating with their partner deadenylases NOT6/6L, inhibiting the assembly of a functional deadenylase complex. Consequently, PieF robustly blocks mRNA poly(A) tail shortening and degradation with striking potency and selectivity for NOT7/8. This inhibition of deadenylation by PieF impedes cell cycle progression in human cells, revealing a novel bacterial strategy to modulate host gene expression.
Similar content being viewed by others
Introduction
Bacterial pathogens have evolved sophisticated RNA regulatory mechanisms to sense and adapt to their environment, evade host immune defenses, and replicate. Neisseria meningitidis and Streptococcus pneumoniae use non-coding RNAs (cia-dependent small RNAs, which are a group of small RNAs that are controlled by the two-component regulatory systems, CiaRH and NrrF, respectively) to regulate key virulence factors, such as capsule formation, metal ion transport, and DNA uptake, allowing adaptation to environmental cues and iron homeostasis during infection1,2.
Beyond regulating their genes, some bacteria manipulate host RNA machinery to enhance their survival. Legionella pneumophila, a Gram-negative bacterium, infects amoebae and human cells using over 300 effector proteins secreted via its Dot/Icm type IV secretion system3. While most effectors are genetically redundant and individually non-essential for intracellular replication, collectively, they enable Legionella to adapt to diverse hosts by modulating cellular processes such as vesicle trafficking, protein synthesis, and cell death4,5.
Legionella infection modifies the expression of 85 human microRNAs (miRNAs), disrupting post-transcriptional regulation in macrophages and promoting bacterial replication6. Intriguingly, Legionella can translocate its small RNAs (sRNAs) into host cells to mimic miRNA functions, downregulating immune regulators and enhancing bacterial survival7. Recently, the effector LegK1 was reported to inhibit small interfering RNA (siRNA) and miRNA activities in human cells through its kinase function and ability to bind Argonaute proteins—critical mediators of miRNA-mediated gene silencing8.
In addition to manipulating miRNA pathways, Legionella subverts host post-transcriptional control by using effectors to inhibit host protein synthesis9,10,11. When infecting macrophages, virulent Legionella secretes effectors that globally inhibit host protein synthesis, blocking IκB production and preventing NF-κB–mediated upregulation of pro-inflammatory cytokines12. Similar strategies are observed in plant pathogens; for example, the Xanthomonas citri effector PthA4 inhibits Citrus sinensis mRNA deadenylase activity to facilitate virulence13.
Bacterial pathogens intricately manipulate host RNA regulatory systems, including targeting mRNA decay and silencing machinery, to facilitate infection and promote their survival. PieF, an effector encoded in a highly dynamic genomic locus of Legionella pneumophila that also encodes other effectors14, was recently shown to interact with the conserved multi-subunit CCR4-NOT complex15,16. This complex is essential in eukaryotic cells for initiating mRNA decay through the enzymatic shortening (deadenylation) of mRNA poly(A) tails17. Here, we reveal the potent and selective mechanism by which PieF targets the host CCR4-NOT complex: by binding with high affinity to its deadenylase subunits, PieF hinders the assembly of a deadenylation-competent CCR4-NOT complex. Overexpression of PieF in human cells induces cell cycle arrest, suggesting that post-infection targeting of host mRNA deadenylation machinery may serve as a strategy to assist with pathogen survival and replication.
Results
Legionella pneumophila PieF effector binds to specific CCR4-NOT subunits
The CCR4-NOT complex contains two deadenylases, NOT7/8 and NOT6/6L, belonging to the DEDD and EEP-type exonuclease families, respectively17 (Fig. 1a). To validate the reported PieF interaction with NOT7/8 and assess potential binding to other CCR4-NOT subunits, we performed a direct interaction assay by pull-down with purified recombinant proteins. N-terminally StrepII-tagged PieF efficiently bound NOT7 and NOT8 with apparent stoichiometry (Fig. 1b). However, NOT9, a non-enzymatic subunit of the CCR4-NOT positioned close to the deadenylases within the complex, did not bind PieF under the same conditions. As a positive control, a fragment of RNF219, a known interactor of the NOT9 subunit, efficiently bound to NOT9. (Fig. 1b)18. Additionally, NOT8, the NOT7 paralog, bound PieF (Fig. 1b).
a Domain organization and schematic representation of the catalytic module of the human CCR4-NOT complex. b Coomassie-stained 15% polyacrylamide gel of the in vitro streptavidin pull-down assay with recombinant N-terminally StrepII-tagged PieF upon incubation with NOT9 module and His6-SUMO-tagged NOT7 and NOT8 deadenylases. * denotes the presence of an N-terminal His6-SUMO-tag. # denotes an N-terminal His6-MBP-tag. Experiment was independently repeated multiple times (n > 3). c Coomassie-stained 15% polyacrylamide gel of the in vitro streptavidin pull-down assay with recombinant N-terminally StrepII-tagged PieF upon incubation with His6-SUMO-tagged NOT7*, NOT6:NOT7*** and NOT6L:NOT7** heterodimers. * denotes presence of an N-terminal His6-SUMO-tag, ** denotes an N-terminal His8-tag. *** denotes untagged NOT7. Experiment was independently repeated multiple times (n > 3). d, e In vitro deadenylation assays with 50 nM of UCUACAU-A20 RNA substrate, His6-SUMO-NOT7 (500 nM; d) and NOT6:NOT7 heterodimer (250 nM; e) without (left panels) and with (right panels) the equimolar to corresponding deadenylase amount of N-terminally His6-tagged PieF. M indicates the tail length RNA marker, and C represents the control sample without the deadenylase. Poly (A) tail length changes were quantified by plotting the most abundant tail length at each time point. Linear regression was used to determine the apparent deadenylation rate (As/min); values are presented as mean ± SE (n = 3). The molecular size markers of right panels are identical to the ones on the left. f Representative isothermal titration calorimetry (ITC) thermograms of the interaction between His6-tagged PieF and His6-SUMO-NOT7. The upper panel shows raw data in (µcal s−1), and the lower panel represents the integration of heat changes associated with each injection (kcal mol−1 of injectant). Data were fitted using a one-site binding model. The parameters of the runs are summarized in Supplementary Table 1. The schematics in (a, d, e) were drawn using Adobe Illustrator 2025. Source data are provided as a Source Data file.
NOT7 is a component of the CCR4-NOT catalytic module17,19,20. Beyond its catalytic function, NOT7 connects the endonuclease-exonuclease-phosphatase (EEP) type deadenylase NOT6 (or its paralog NOT6L) to the central MIF4G domain of NOT1 (Fig. 1a)19,21,22. We investigated the potential interaction between PieF and NOT7 within the NOT7:NOT6 and NOT7:NOT6L heterodimer contexts. Pull-down assays showed no detectable interaction between PieF and these heterodimers, indicating that NOT6/6L binding likely precludes PieF association (Fig. 1c). However, a systematic quantitative analysis of the ratios of CCR4-NOT subunits in human cells using mass spectrometry revealed that NOT6 is strikingly substoichiometric compared to other CCR4-NOT subunits23. These data suggest that the predominant form of CCR4-NOT in human cells may be devoid of NOT6/6L subunits leaving NOT7/8 subunits available for potential interaction with PieF.
PieF is a potent and specific inhibitor of NOT7/8 deadenylation activity
We then investigated PieF’s impact on NOT7’s deadenylation activity in isolation and within the NOT6:NOT7 heterodimer. Using an in vitro deadenylation assay, we tested NOT7 and NOT6:NOT7 with and without an equimolar amount of PieF (Fig. 1d,e, Supplementary Fig. 1a,b). The substrate was a 5′-fluorescently-labeled UCUACAU-A20 synthetic RNA.
While the NOT6:NOT7 heterodimer’s activity remained unchanged, NOT7 alone was inhibited ~5-fold by PieF. In the same experimental setup, NOT8 (which has ~3-fold lower basal activity than NOT7) was inhibited ~6-fold by PieF (Supplementary Fig. 1c,d).
Various DEDD-type exoribonucleases are present in cells, including those capable of poly(A) tail shortening, such as poly(A)-specific ribonuclease (PARN)24. We purified recombinant PARN and observed that PARN is not inhibited by PieF, unlike NOT7/8. This suggests that PieF specifically targets the deadenylation activity of the CCR4-NOT complex rather than affecting DEDD-type exoribonucleases in general (Supplementary Fig. 1e).
We used isothermal titration calorimetry (ITC) to determine PieF’s binding affinity for NOT7/8. The measured dissociation constant (Kd) values for NOT7 and NOT8 were ~24 nM, indicating high-affinity binding (Fig. 1f and Supplementary Fig. 1f; Supplementary Table 1).
Given the mutual exclusivity of NOT6 and PieF binding to NOT7 and PieF’s low-nanomolar range affinity, we investigated whether PieF could outcompete NOT6 from the NOT6:NOT7 heterodimer. We purified and immobilized NOT6:NOT7 on beads, then titrated PieF up to 200-fold molar excess. Surprisingly, we did not observe the formation of the NOT7:PieF complex or displacement of NOT6 (Supplementary Fig. 1g).
We employed analytical size exclusion chromatography to test whether PieF interferes with the NOT7 and NOT1 association. We found that NOT7 simultaneously binds the NOT1 MIF4G domain and PieF, forming a NOT1:NOT7:PieF ternary complex (Fig. 2a). Mass photometry analysis revealed a stable 1:1:1 complex within a single symmetrical peak (Fig. 2b).
a Analytical size exclusion chromatography profiles (left panel) and a 15% Coomassie-stained gel (right panel) of the purified components and intermediates of the NOT1:NOT7:PieF complex. The tags used to purify the individual components were preserved during the reconstitution of the complex. Experiment was independently repeated multiple times (n > 3). Source data are provided as a Source Data file. b Mass photometry analysis of the NOT1:NOT7:PieF complex. The tags used for purification were preserved during the reconstitution of the complex. c Cryo-EM reconstruction of the NOT1:NOT7:PieF complex at 2.8 Å resolution. d Cryo-EM reconstruction of the NOT1:NOT8:PieF complex at 3.5 Å resolution. e Superimposed structures of the NOT1:NOT7/8:PieF complexes in cartoon representation. f Secondary structure annotation of PieF structure. The schematics in (a, b) were drawn using Adobe Illustrator 2025.
Cryo-EM reveals how PieF mimics NOT6/6 L and inhibits NOT7/8
To understand the binding dynamics of NOT7 with NOT1, PieF, and NOT6 and to elucidate the mechanism of PieF-mediated NOT7/8 inhibition, we determined the structure of the NOT1:NOT7:PieF complex to 2.8 Å resolution and NOT1:NOT8:PieF complex to 3.5 Å resolution using cryo-electron microscopy (cryo-EM) (Fig. 2c, d; Supplementary Table 2). The high quality of the density map enabled us to build a complete NOT1 MIF4G domain (residues E1093–S1317). NOT7 contains most residues (Q10–G263 out of 285), with only a few flexible residues missing at both termini, while PieF includes all residues except for I19–V25 in the loop connecting the β2 and β3 strands (Fig. 2e, f and Supplementary Fig. 2). Using Foldseek analysis25, we determined that the PieF of Legionella pneumophila has a fold, consisting of a five-stranded beta sheet and two alpha helices, with no close homologs in PDB, CATH50, AFDB-Proteome, and AFDB-SwissProt databases.
PieF and NOT7 form a complex through a bipartite interface with two distinct interaction regions (Fig. 2e, f): the first interface involves PieF’s β5 strand and adjacent loops, while the α3 helix of PieF mediates the second interface.
A striking feature of the latter interface is how PieF binds NOT7 using its α3 helix. This interaction occludes NOT7’s active site and involves PieF’s K124 displacing one of the two catalytic Mg2+ ions. This displacement occurs due to the formation of salt bridges between the side chains of K124PieF with D40NOT7 and D161NOT7 (Fig. 3a, b). K124PieF further hydrogen bonds with the carbonyl group of F156NOT7, and via its carbonyl group, K124PieF interacts with H157NOT7 (Supplementary Fig. 3a). The α3 helix also makes stabilizing hydrophobic contacts: L121PieF interacts with Y160NOT7 and F43NOT7, and Y119PieF interacts with L115NOT7 (Supplementary Fig. 3a).
a Cartoon representation of the NOT7 (blue) active site and part of the helix α3 of PieF (red). Residues K206–E211 were omitted for clarity. The residues involved in the interaction are represented as sticks and the Mg2+ atom is shown as a green sphere. Dashed lines indicate metal coordination (NOT7 and Mg2+) or hydrogen bonding (NOT7 and PieF). D40NOT7 and D161NOT7 are additionally involved in salt bridge interactions with K124PieF. Interaction of K124PieF with D40NOT7 and D161NOT7 displaces a catalytic Mg2+ from the active site, and the NOT7 crystal structure (PDB: 4gmj; transparent blue) is superimposed to illustrate this. b Close-up view of (a). c NOT7:PieF, as represented in (a), superimposed on the crystal structure of the NOT6:NOT7 heterodimer (PDB: 7ax1). The leucine-rich repeat region of NOT6 binds to the same site of NOT7 as PieF. The N-terminal region of NOT6 (solid yellow) extends the binding interface. d Close-up view of (c).
The principal interface between PieF and NOT7 is primarily formed by a section of PieF’s β5 strand (residues S101–F103) and several adjacent residues (K96–G100) immediately preceding it (Supplementary Fig. 3b). The β5 strand of PieF assembles into a short β-sheet with the β3 strand of the NOT7 (residues A48–P50). Additionally, R49NOT7 forms a salt bridge with E117PieF (Supplementary Fig. 3b). In addition to creating a β-sheet, F103PieF along with I88PieF, K96PieF, and I97PieF, establish a hydrophobic network of contacts converging on with L71NOT7. The loop between PieF’s β5 strand and α3 helix (residues F105–E114) also interacts with NOT7, stabilizing the α3 helix. F111PieF makes another set of critical hydrophobic contacts with V47NOT7, while its carbonyl group interacts with R113PieF, stabilizing both F111PieF and S110PieF.
The strong interactions between NOT6 and NOT7, forming a stable complex with an effectively negligible off-rate, account for PieF’s inability to displace NOT6 from the heterodimer. Specifically, NOT6’s β2 strand (residues E33–I36) forms a β-sheet with NOT7’s β3 strand (residues A48–I51)22. This interaction interface directly overlaps with PieF’s binding site on NOT7 (Fig. 3c, d, Supplementary Fig. 3c, d), underscoring the mutual exclusivity of these protein interactions.
Within the NOT6:NOT7 heterodimer, R49NOT7 plays a crucial role, forming a salt bridge with E35NOT6 and hydrogen bonding with the carbonyl of R12NOT6 (Supplementary Fig. 3c, d). Despite this relatively small overlapping interface, PieF surprisingly cannot displace NOT6 from the NOT6:NOT7 heterodimer. We hypothesized that inability to displace NOT6/6L may be due to an additional binding interface, where the N-terminal loop of NOT6/6L wraps around NOT7’s α2 helix and adjacent loop between the helix and β3 strand (Fig. 3c, d, Supplementary Fig. 3c–f, Supplementary Fig. 4a). To test the hypothesis, we purified the version of the NOT6:NOT7 heterodimer, in which NOT6 lacks residues M1–K28, and did the pull-down assays in presence of the excess of PieF. Even with the truncated N-terminal portion of NOT6, PieF was still unable to outcompete NOT6 from the heterodimer due to the continued stability of the NOT6:NOT7 interface, which likely exhibits very slow dissociation kinetics (Supplementary Fig. 4a–c). In deadenylation assays, we observed no inhibtion of the NOT6Δ1-28:NOT7 heterodimer activity by PieF further corroborating the pull-down results (Supplementary Fig. 4d).
We determined the cryo-EM structure of the NOT1:NOT8:PieF ternary complex at 3.5 Å resolution to further elucidate paralog-specific differences in PieF-mediated deadenylase inhibition. The resulting model includes the same NOT1 and PieF residues as in the NOT7 complex, while NOT8 is nearly complete, missing only the first five and last 29 residues (V6–G263 out of 292) (Fig. 2e, Supplementary Fig. 2). NOT8 associates with NOT1 in a manner nearly identical to NOT7 closely matching previously determined structures (Supplementary Fig. 4e, f)21, suggesting that PieF does not prevent the incorporation of the inhibited deadenylase into the CCR4-NOT complex (Supplementary Fig. 4e, f). The structures of the NOT7 and NOT8 are nearly identical (Supplementary Fig. 4g, h vs 3a,b; r.m.s.d ~0.56 Å over 254 Cα pairs), with a few notable differences. NOT8 does not form a β3 strand; however, R49NOT8 forms a salt bridge with E117PieF, as in the NOT7 structure (Supplementary Fig. 2; Supplementary Figs. 4h vs 3b). NOT8 residues do not contact S99PieF and G100PieF. At the same time, T98PieF interacts with the R173NOT8 (N173NOT7), which additionally interacts with the carbonyl group of K96PieF, a contact absent in the NOT7 structure (Supplementary Figs. 4h vs 3b). Although highly similar, NOT7 and NOT8 exhibit their main structural and sequence divergence beyond residue 265—a region likely flexible and unresolved in either cryo-EM structure (Supplementary Fig. 5).
The C-terminal helix enhances PieF inhibition through synergistic binding
To elucidate the roles of specific PieF residues in their interaction with NOT7 and subsequent inhibition, we employed a structure-guided approach to generate several PieF mutants (Fig. 4a). We then evaluated these mutants through deadenylation assays with NOT7 to assess inhibitory effects and isothermal titration calorimetry (ITC) to measure changes in binding affinity.
a Schematic representation of the mutated residues in PieF used in this study. b In vitro deadenylation assays with 50 nM of UCUACAU-A20 RNA substrate, 500 nM of His6-SUMO-NOT7 and 500 nM of His6-tagged PieFK124A, PieFK124R, PieF 5 M, PieF 2 M, and PieF 3 M. The molecular size markers on all gels are identical to those shown on the leftmost panel. Experiment was independently repeated multiple times (n = 3). Source data are provided as a Source Data file. c Violin plots showing the poly(A) tail distribution in HEK293T cells expressing transfected GFP control (green), GFP-NOT7D40A (blue), GFP-PieF (dark red), and GFP-PieF 5M (orange) based on mRNA poly(A) length estimates by nanopore sequencing. Modal poly(A) tail length values are next to the corresponding violin plot. Box plots show median value (solid bold line) and 1st and 3rd quartiles, and whiskers represent 1.5 × IQR (interquartile range). n is the number of basecalled reads in each sample. d Cumulative cell counts at day 2, day 4, and day 6 after transfection and overexpression of a control GFP construct (green) compared to GFP-NOT7D40A (blue), GFP-PieF WT (dark red) or GFP-PieF 5 M (orange) constructs. Values are presented as mean ± standard deviation (n = 3). Source data are provided as a Source Data file. e The Legionella pneumophila effector PieF or NOT6 bind to NOT7 within the CCR4-NOT complex in a mutually exclusive manner. PieF binding impairs the assembly and deadenylation activity of the CCR4-NOT complex, leading to longer poly(A) tails, increased mRNA levels, and defects in cell cycle progression. The schematics were drawn using Adobe Illustrator 2025.
We first examined PieFK124A to assess if removing salt bridges with D40NOT7 and D161NOT7 would fully relieve inhibition and tested PieFK124R for functional substitution. Under identical conditions to wild-type PieF, deadenylation assays showed NOT7’s activity with PieFK124A was ~2.5-fold faster than with wild-type PieF, while ITC revealed ~23-fold weaker binding (Fig. 4b, Supplementary Fig. 6b, c). PieFK124R had milder effects: ~1.2-fold faster deadenylation and ~4-fold decreased affinity (Fig. 4b, Supplementary Fig. 6b, d). These results indicate salt bridge interactions with catalytic residues are crucial for PieF’s total inhibitory capacity.
To further investigate the residues crucial for PieF’s interaction with and inhibition of NOT7, we created a PieF 5M mutant (Fig. 4a; orange circles): two mutations in the helix (E117R and L121E; 2M, brown rim) and three around the β5 strand (T98E, F111E, and R113E; 3M, dark blue rim). PieF 5M showed no inhibition of NOT7 deadenylation or detectable binding (Fig. 4b, Supplementary Fig. 6b, e). We then wondered which mutations were responsible for the loss of PieF-mediated inhibition: those in the helix (2M) or around the β5 strand (3M). To address this, we tested the 2M and 3M mutants separately. PieF 2M behaved similarly to PieFK124A, with ~2.5-fold faster deadenylation and ~18-fold weaker binding than wild-type PieF (Fig. 4b, Supplementary Fig. 6b, f). PieF 3 M mirrored PieF 5M, showing no inhibition or binding to NOT7 (Fig. 4b, Supplementary Fig. 6b, g). These results suggest that, although the PieF C-terminal helix contributes to inhibition, full inhibitory function depends on stabilization and proper orientation by proximal residues.
PieF disrupts mRNA poly(A) tail control and impairs cell cycle progression in human cells
To assess PieF’s impact on deadenylation in human cells, we expressed GFP-tagged PieF in HEK293T cells and measured transcriptome-wide poly(A) tail lengths using nanopore direct RNA sequencing. Despite PieF’s lack of effect on preassembled NOT6/6L:NOT7/8 heterodimers in vitro, GFP-PieF expression in cells dramatically increased modal mRNA poly(A) tail length from ~54 to ~150 adenosines, consistent with a highly pronounced CCR4-NOT-mediated deadenylation impairment (Fig. 4c, green vs red). In vitro assays confirmed that GFP-PieF inhibits NOT7 as effectively as untagged PieF, corroborating the cellular data (Supplementary Fig. 6h).
To confirm that the observed deadenylation impairment resulted from CCR4-NOT dysfunction, we overexpressed GFP-tagged NOT7D40A, a well-characterized dominant-negative mutant26. The modal poly(A) tail length with GFP-NOT7D40A (~146 adenosines) closely resembled that observed with GFP-PieF (~150 adenosines), demonstrating that PieF overexpression phenocopies NOT7 inactivation (Fig. 4c, blue). Crucially, cells overexpressing the GFP-PieF 5 M mutant exhibited mRNA poly(A) tail lengths similar to GFP alone (~55 vs. ~54 adenosines), consistent with the PieF inhibitory effect being critically dependent on its stable binding to NOT7 (Fig. 4c, orange).
Control of mRNA poly(A) tail length plays a crucial role in cell cycle progression regulation27. Since poly(A) tail length is primarily controlled by CCR4-NOT deadenylation activity, we investigated how PieF affects cell cycle progression by inhibiting this activity. We observed that cells expressing GFP-tagged PieF and NOT7D40A proliferated more slowly than those expressing GFP or GFP-PieF 5 M controls. A minor decrease in viability (from 95 to 85%) upon GFP-PieF and GFP-NOT7D40A overexpression only partially explains these observations (Fig. 4d and Supplementary Fig. 7a).
To assess PieF’s impact on cell cycle progression via CCR4-NOT inhibition, we used DAPI staining and flow cytometry to quantify DNA content across cell cycle phases (G0/G1, S, G2, M). This method allowed us to evaluate how PieF and its variants affected cell cycle distribution. Cells expressing GFP-PieF showed a ~10% decrease in G0/G1 phase frequency and a comparable increase in S phase frequency compared to GFP-expressing cells (Supplementary Fig. 7b, c, red vs green). GFP-NOT7D40A-expressing cells exhibited a similar reduction in the G0/G1 phase and an increase in S phase frequencies relative to GFP vector control, consistent with a critical role of regulated mRNA decay and translational repression in maintaining normal cell cycle progression (Supplementary Figs. 7b, c, blue vs green). Notably, GFP-PieF 5M control cells reversed this trend, emphasizing that PieF’s binding to NOT7 is essential for modulating the cell cycle (Supplementary Figs. 7b, c, orange vs red). These results suggest that L. pneumophila employs PieF to inhibit CCR4-NOT-mediated mRNA deadenylation as an additional strategy to disrupt cell cycle signaling cascades and control the host cell cycle.
Discussion
We show that PieF manipulates host mRNA decay by inhibiting the NOT7/8 deadenylase paralogs of the CCR4-NOT complex. This inhibition is achieved through physical occlusion of the NOT7/8 active site and displacement of a catalytic Mg²⁺ ion, combined with the prevention of native deadenylation complex assembly by blocking NOT6/6L binding to NOT7/8. The CCR4-NOT complex assembles via interactions among its subunits, with NOT6/6L binding to NOT7/8—scaffolded by the NOT1 subunit—being crucial for forming a functional deadenylase complex. Quantitative proteomic analyses28 show that NOT6 is present at significantly lower levels (~700 copies per cell) compared to other subunits such as NOT7 (~15,000 copies per cell), NOT1 (~18,000 copies per cell), and NOT9 (~19,000 copies per cell), suggesting that NOT6 availability may indeed be limiting relative to other complex subunits23.
Since PieF cannot outcompete NOT6 once it is bound to NOT7, its inhibitory effect is likely exerted under cellular conditions where there is a strong bias in the distribution of NOT6/6L and NOT7/8 subunits, favoring the latter. By sequestering the exposed NOT7/8 subunits in partially assembled CCR4-NOT, PieF may prevent their association with the limiting amounts of NOT6/6L. It is particularly intriguing that PieF binds to NOT7/8 in such a way that NOT7/8 can still dock onto the MIF4G domain of NOT1, thereby trapping the entire CCR4-NOT complex in a catalytically incompetent state (Fig. 4e). In this PieF-bound state, CCR4-NOT may function in a dominant-negative manner, impairing its critical roles in mRNA stability regulation and expression. This interference underscores the importance of the assembly dynamics and turnover of the CCR4-NOT complex in regulating mRNA decay. When PieF is expressed in human cells, it leads to striking transcriptome-wide perturbations in polyadenylation, including poly(A) tail lengthening, reduced mRNA decay, and consequent mRNA accumulation. The downstream effects of this disruption of mRNA homeostasis include impaired cell proliferation and delayed cell cycle S phase progression.
Our biochemical, structural, and cellular evidence suggests that L. pneumophila PieF has evolved to target CCR4-NOT-mediated deadenylation by structurally mimicking the NOT7/8-interacting region of the leucine-rich repeat domain of NOT6/6L. Notably, the K124 residue, which plays a pivotal role in PieF’s inhibitory properties, is unique among Legionella species (Supplementary Figs. 8a, 9 and 10). Interestingly, the inhibitory strategy of PieF parallels that of certain bacterial toxin-antitoxin systems29,30. In these complexes, the antitoxin protein neutralizes its RNase toxin by physically occluding the active site and displacing the catalytic Mg²⁺ ion with a positively charged side chain. For example, the FitA antitoxin inserts an Arg residue into the FitB toxin’s active site to expel the Mg²⁺ required for catalysis31. Similarly, the VapB antitoxin uses an Arg (R79) to eject the Mg²⁺ cofactor from the VapC toxin32. These parallels underscore a convergent inhibitory strategy among bacterial pathogens, whereby PieF, similar to bacterial antitoxins, inactivates NOT7’s enzymatic core by displacing its catalytic metal ion.
Although sequence homologs of PieF are found in other bacteria, the absence of this critical K124 residue (Supplementary Figs. 8a, 9 and 10) suggests that these homologs may represent examples of a conserved effector structure being used for different biological roles. As is typical with many effectors, redundancy in these systems complicates the identification of the specific advantage conferred by PieF’s subtle disruption of mRNA homeostasis, mainly since the low levels of PieF produced during natural infection and deletion of the pie operon do not prevent L. pneumophila infections in model systems14.
L. pneumophila replicates in Acanthamoeba castellanii and numerous other protist hosts, where impairment of host proliferation is essential for successful replication of the bacterium33,34. Although direct manipulation of cell cycle progression by targeting cyclins and cyclin-dependent kinases is a common feature of intracellular pathogens34,35, the signaling cascade components can vary significantly across species, making this approach less effective across diverse hosts. In contrast, the NOT7/8 complex is highly conserved across a wide range of L. pneumophila hosts—from protists to higher eukaryotes (Supplementary Figs. 5, 8b and 11)—establishing it as an effective target for mRNA homeostasis manipulation with a single effector sequence. By injecting PieF, L. pneumophila gains the means to manipulate mRNA levels directly and the expression of genes involved in cell cycle progression, adding another regulatory mechanism to its control over host metabolism and life cycle. Thus, while PieF is not essential for infection, its ability to target NOT7/8 may confer an evolutionary advantage by expanding the pathogen’s range of potential hosts.
In conclusion, we propose that PieF’s unique inhibitory interface could serve as a template for developing potent, highly specific deadenylase inhibitors, given the crucial roles of the CCR4-NOT complex in various biological and disease processes ranging from T cell quiescence maintenance36 to metastatic progression37.
Methods
DNA constructs
The DNA constructs are summarized in Supplementary Table 3.
Protein production and purification
His6-PieF was produced in E. coli BL21(DE3) Star cells using 1 L of autoinduction media at 37 °C overnight. The cells were harvested and resuspended in a lysis buffer containing 50 mM potassium phosphate pH 7.5, 300 mM NaCl, 25 mM imidazole, and lysed by sonication. The lysate was clarified by centrifugation at 40,000 × g for 45 min and loaded on a 1 ml nickel-charged IMAC column (Cytiva). The bound protein was washed first with 20 column volumes of lysis buffer and then with five column volumes of lysis buffer supplemented with 35 mM imidazole. PieF was eluted from the column in 0.5 ml fractions with the same buffer supplemented with 250 mM imidazole. Peak fractions from nickel affinity chromatography were then loaded and eluted on Superdex 75 26/600 (Cytiva) equilibrated in a buffer containing 10 mM HEPES/NaOH pH 7.5, 200 mM NaCl, 2 mM DTT. The peak fractions were then pooled together, concentrated to 6–12 mg/ml, flash-frozen in liquid nitrogen, and stored at −80 °C.
All other His6-PieF constructs, His6-NOT1(E1093–S1317), His6-SUMO-NOT7, and His6-Strep-SUMO-NOT7 were produced and purified identically to the wild-type PieF.
His6-GFP-PieF and His6-SUMO-NOT8 were produced in E. coli BL21(DE3) Star cells using 2 L of autoinduction media at 20 °C overnight. His6-GFP-PieF and His6-SUMO-NOT8 were purified identically to the wild-type PieF.
Purifications of the NOT6:NOT7 heterodimer and NOT9 module were described previously19,38. His6-NOT6:StrepII-SUMO-NOT7 was purified identically to NOT6:NOT7, with the only difference being that tags were not cleaved. NOT6Δ1-28:Strep-NOT7 was purified as described previously22.
NOT6L:His8-NOT7 was produced in Spodoptera frugiperda Sf21 insect cells using the MultiBac baculovirus expression system. The Sf21 cells were grown to a density of 2 × 106 cells/ml at 27 °C in Sf900II medium (Thermo Fisher Scientific), infected with the V1 NOT6L:His8-NOT7 stock of baculovirus, and harvested 48 h after they stopped dividing. Cells were resuspended in lysis buffer (50 mM HEPES/NaOH pH 7.5, 500 mM NaCl) and lysed by sonication. The lysate was cleared by centrifugation at 40,000 × g for 1 h at 4 °C, filtered through 0.45 µm syringe-driven filters (Millipore), and loaded on a 5 ml nickel-charged IMAC column (Cytiva) equilibrated in buffer containing 50 mM potassium phosphate pH 7.5, 300 mM NaCl and 25 mM imidazole. After several washing steps, the NOT6L:His8-NOT7 heterodimer was eluted with the same buffer supplemented with 250 mM imidazole. The top 5 ml were loaded and eluted on Superdex 200 16/600 equilibrated in 10 mM HEPES/NaOH pH7.5; 200 mM NaCl; 2 mM DTT. The peak fractions were then pooled, concentrated to 7 mg/ml, flash-frozen in liquid nitrogen, and stored at −80 °C.
The expression plasmid of human PARN was a kind gift from Perry Blackshear. His6-PARN was produced in E. coli BL21 (DE3) Star cells using 1 L of autoinduction media at 20 °C overnight. The cells were harvested and resuspended in a lysis buffer containing 50 mM HEPES/NaOH pH 7.5, 1000 mM NaCl, 5% (v/v) glycerol, 25 mM imidazole, and lysed by sonication. The lysate was clarified by centrifugation at 40,000 × g for 40 min and loaded on a 5 ml nickel-charged IMAC column (Cytiva). Contaminants were removed by washing with lysis buffer supplemented with 40 mM imidazole, and PARN was eluted in lysis buffer supplemented with 250 mM imidazole. PARN was further purified by size exclusion chromatography on a Superdex 200 26/600 column (Cytiva) in a buffer containing 10 mM HEPES/NaOH pH 7.5, 200 mM NaCl, 5% (v/v) glycerol, 2 mM DTT. The peak fractions were then pooled, concentrated to 0.9 mg/ml, flash-frozen in liquid nitrogen, and stored at −80 °C.
Production and reconstitution of ternary complexes for cryo-EM
For reconstitution of the NOT1:NOT7:PieF complex, all three proteins were purified in parallel, as described above, and then mixed after the nickel-affinity chromatography step. The proteins were left to incubate on ice overnight. The next day, the complex was loaded and eluted on Superdex 200 26/600 (Cytiva) equilibrated in a buffer containing 10 mM HEPES/NaOH pH 7.5, 200 mM NaCl, 2 mM DTT. Reconstitution of the NOT1:NOT8:PieF was done identically to the NOT7 version of the complex. NOT1:NOT7:PieF and NOT1:NOT8:PieF were concentrated to 6.5 mg/ml and 8.3 mg/ml, respectively, mixed with 0.7 mM fluorinated octyl maltoside (Anatrace), and adsorbed onto glow-discharged holey carbon-coated grids (Quantifoil 300 mesh, Au R1.2/1.3) for 10 s. Grids were then blotted for 2 s at 100% humidity at 10 °C and frozen in liquid ethane using a Vitrobot Mark IV (Thermo Fisher Scientific).
Cryo-EM data collection, data processing, and model building
Both NOT1:NOT7:PieF and NOT1:NOT8:PieF datasets were collected in counted mode (in Electron Event Representation format) on a CFEG-equipped Titan Krios G4 (Thermo Fisher Scientific) operating at 300 kV and × 165,000 magnification. A Selectris X imaging filter (Thermo Fisher Scientific) with slit width set to 10 eV on a Falcon 4i direct detection camera (Thermo Fisher Scientific) with a calibrated pixel size of 0.732 Å was used. Movies were collected at a total dose of 55.6 e−/Å2 (NOT1:NOT7:PieF) or 52.4 e−/Å2 (NOT1:NOT8:PieF), both fractionated to ~1 e−/Å2/fraction for motion correction.
Patched motion correction, CTF parameter estimation, particle picking, extraction, and initial 2D classification were performed in SIMPLE 3.039. All downstream processing was done in cryoSPARC or RELION, using the csparc2star.py script in the UCSF pyem package to convert between formats40,41,42. Global resolution was estimated from gold-standard Fourier shell correlations (FSCs) using the 0.143 criterion, and local resolution estimation was calculated within cryoSPARC using an FSC threshold of 0.5.
The cryo-EM processing workflow for NOT1:NOT7:PieF is outlined in Supplementary Fig. 12. Briefly, particles were subjected to two rounds of reference-free 2D classification (k = 300) using a 140 Å soft circular mask. Four volumes were generated from a 225,824 particle subset of the 2D-cleaned particles after multi-class ab initio reconstruction using a maximum resolution cutoff of 7 Å. Output volumes were low-pass-filtered to 7 Å and used as references for a 4-class heterogeneous refinement against the full 2D-cleaned particle set (1,062,166 particles). Particles (417,581) from the most populated and structured class were selected and non-uniform refined against their corresponding volume lowpass-filtered to 15 Å, generating a 3.0 Å map. Bayesian polishing was performed in RELION followed by non-uniform refinement against the pre-polished volume, lowpass-filtered to 15 Å, yielding a 2.9 Å volume that was further improved to 2.8 Å after local and global CTF refinement (fitting tilt and trefoil).
The cryo-EM processing workflow for NOT1:NOT8:PieF is outlined in Supplementary Fig. 13. Briefly, particles were subjected to two rounds of reference-free 2D classification (k = 200) using a 150 Å soft circular mask. Selected particles (1,384,683) were then subjected to heterogeneous refinement against the four ab initio volumes generated from the NOT1:NOT7:PieF dataset, low-pass filtered to 7 Å. Particles (407,968) from the most populated and structured class were selected and non-uniformly refined against their corresponding volume lowpass-filtered to 15 Å, generating a 3.5 Å map. Bayesian polishing followed by 2D classification (k = 100, 150 Å soft circular mask) resulted in the selection of 200,675 particles. These particles were non-uniform refined against their corresponding pre-polished volume, low-pass-filtered to 15 Å, yielding a 3.5 Å volume that could be marginally improved by subsequent local CTF refinement. The “tightTarget” model of deepEMhancer was used on the final NOT1:NOT8:PieF half maps for postprocessing to improve map connectivity and side chain density43.
Atomic models for NOT1:NOT7:PieF and NOT1:NOT8:PieF were generated first by docking AlphaFold2-Multimer models into their respective cryo-EM volumes, followed by manual rebuilding and real-space refinement in Coot v0.9.8.344,45. Both models were further refined in real space using PHENIX with rotamer, Ramachandran restraints, and with or without the secondary structure restraints (the latter applied only for NOT1:NOT8:PieF) against either the global B-factor sharpened map of NOT1:NOT7:PieF or the deepEMhancer map for NOT1:NOT8:PieF, yielding the models described in Supplementary Table 243. All models were validated using MolProbity within PHENIX46,47.
Analytical size exclusion chromatography
Individual proteins His6-SUMO-NOT7 (10000 pmol), His6-NOT1(E1093–S1317) (15000 pmol), His6-PieF (15000 pmol) and complexes His6-SUMO-NOT7/His6-NOT1(E1093–S1317), His6-SUMO-NOT7/His6-PieF, His6-SUMO-NOT7/His6-NOT1(E1093–S1317)/His6-PieF (mixed at the same ratio to His6-SUMO-NOT7) were loaded on Superdex 200 10/300 (Cytiva) equilibrated buffer containing 10 mM HEPES/NaOH pH 7.5, 200 mM NaCl, 2 mM DTT. Complexes were incubated on ice for 30 min before loading on the column. Detection of the complexes and individual proteins was tracked using UV absorbance at 280 nm. The content of the peak fractions was analyzed using SDS-PAGE electrophoresis followed by Coomassie staining. Chromatograms were exported and re-plotted in one graph using Prism10.
Mass photometry
Mass photometry of His6-NOT1(E1093–S1317):His6-SUMO-NOT7:His6-PieF was performed using the Refeyn TwoMP mass photometry instrument in a buffer containing 10 mM HEPES/NaOH pH 7.5, 200 mM NaCl, 2 mM DTT. Molecular weight calibrations were performed using two protein oligomer solutions, β-amylase (56, 112, and 224 kDa) and Thyroglobulin (670 kDa). The data acquisition was performed with AcquireMP (v2023 R1.1) software and data analysis using DiscoverMP (v2023 R1.2) software.
Deadenylation assays
Deadenylation reactions were carried out at 37 °C in a buffer containing 20 mM PIPES/NaOH pH 7.0, 40 mM NaCl, 10 mM KCl and 2 mM Mg(OAc)2. A purified His6-SUMO-NOT7 (500 nM), His6-SUMO-NOT8 (500 nM), or NOT6:NOT7 heterodimer (250 nM) was mixed with the PieF variant at the equimolar ratio and left to incubate on ice for 15 min. In the assays with PARN (2 nM), PieF was added in 20-fold excess. To start the reaction, synthetic 5′-fluorescein-labeled UCUACAU-A20 substrate (50 nM) was added. To stop the reaction at the corresponding time point, 3x reaction volumes of RNA loading dye were added (95% [v/v] deionized formamide, 17.5 mM EDTA pH 8, 0.01% [w/v] bromophenol blue). The products were resolved on a denaturing TBE-urea polyacrylamide gel, which was subsequently imaged using Sapphire FL Biomolecular Imager (Azure Biosystems) or Amersham Typhoon Biomolecular Imager (Cytiva). Poly(A) tail length changes were quantified by plotting the most abundant tail length at each time point48. The apparent deadenylation rate (As/min) was estimated using linear regression.
Isothermal titration calorimetry
ITC measurements were performed using an iTC200 calorimeter (Malvern Panalytical/MicroCal, Netherlands/USA) at 25 °C. Recombinant purified protein samples were dialyzed against 10 mM HEPES pH 7.5, 200 mM NaCl buffer at 4 °C, and final concentrations determined by UV absorbance at 280 nm. Each experiment consisted of 18 injections of 2.1 µl of 200 µM His6-tagged PieF variant into the cell containing 20 µM His6-SUMO-tagged NOT7 or NOT8 made 180 s apart. These were preceded by an initial 0.5 µl injection to accommodate the reagent’s interaction during pre-titration thermal equilibration at the tip of the syringe. The initial injection was removed during data processing. Reference titrations were subtracted from experimental data, and thermodynamic parameters were determined using a one-site binding model in Origin 7.0 Malvern/MicroCal data analysis software. All ITC run parameters are summarized in Supplementary Table 1. Representative thermograms are shown in Fig. 1, Supplementary Figs. 1 and 6.
Streptavidin pull-down assays
StrepII-tagged (C-terminally) MBP, StrepII-tagged (N-terminally) PieF, StrepII-tagged (C-terminally) and MBP-tagged (N-terminally) RNF219 (residues S434–Q600), StrepII-tagged (C-terminally) His6-SUMO-tagged (N-terminally) NOT1 (E1093–S1317) were produced in E. coli BL21 (DE3) Star cells (Thermo Fisher Scientific) grown in autoinduction medium overnight at 37 °C. Cells were resuspended in binding buffer (50 mM HEPES/NaOH pH 7.5, 200 mM NaCl) and lysed by sonication. Lysate was cleared by centrifugation at 40,000 × g for 1 h at 4 °C. The cleared lysates were incubated with StrepTactin Sepharose resin (Cytiva) for 1 h at 4 °C. After the incubation, beads were washed three times with wash buffer (50 mM HEPES/NaOH pH 7.5, 200 mM NaCl, 0.03% [v/v] Tween) and once with binding buffer. In pull-down from Fig. 1b, 2000 pmol of purified His6-SUMO-NOT7, His6-SUMO-NOT8, and NOT9 module were added to bead-bound proteins and incubated for 1 h at 4 °C. In Fig. 1c, 500 pmol of His6-SUMO-NOT7, NOT6:NOT7 and NOT6L:NOT7 were added. In the competition pull-down assay in Supplementary Fig. 1g, 500 pmol of His6-NOT6:StrepII-SUMO-NOT7 were immobilized, and 1–200 fold molar excess of His6-PieF was added. In the pull-down assay in Supplementary Fig. 4b, 500 pmol of MBP-Strep, His6-StrepII-SUMO-NOT7, His6-NOT6:StrepII-SUMO-NOT7, NOT6Δ1-28:StrepII-NOT7 were immobilized and 100-fold molar excess of PieF was added. In the competition pull-down assay in Supplementary Fig. 4c, 500 pmol of NOT6Δ1-28:StrepII-NOT7 were immobilized, and 1–200 fold molar excess of PieF was added. After the incubation, bound proteins were washed three times with the binding buffer. Beads were incubated for 1 h at 4 °C in 40 μl of binding buffer supplemented with 50 mM biotin to elute the complexes. The eluted proteins were analyzed using SDS-polyacrylamide gel electrophoresis followed by Coomassie blue staining.
Transfection, flow cytometry, and total RNA isolation
HEK293T cells were maintained in Dulbecco modified Eagle medium (DMEM) (Gibco) containing 10% (v/v) fetal bovine serum (FBS) and 1× GlutaMAX (Gibco), at 37 °C with 5% CO2. HEK293T cells were seeded at a density of 0.5 × 106 cells per well in six-well plates. The next day, cells were transfected with 10 μg of plasmid in a complex with 10 μl lipofectamine 2000 (Invitrogen); after 16 h the media was replaced. Cells were reseeded on day 2 and day 4 and counted on day 2, day 4, and day 6 after transfection. For cell cycle analysis, cells were resuspended in PBS and stained with fixable Viability Dye eFluor 780 (ThermoFisher) before fixation and permeabilization using the manufacturer’s instructions (BD Biosciences), followed by staining with DAPI and analysis by flow cytometry. Data were analyzed with FlowJo v10.9.0 software. Total RNA was extracted from one well of a 6-well plate on day 4 after transfection using TRI Reagent (Zymo) and Direct-zol RNA Microprep kit (Zymo) following the manufacturer’s instructions.
Nanopore sequencing and data analysis
Libraries for direct RNA sequencing using library kit SQK-RNA004 (Oxford Nanopore Technologies) and sequenced with PromethION (Oxford Nanopore Technologies) using FLO-PRO004RA flow cells. One flow cell was used for each sample. Basecalling and poly(A) tail length estimates were performed using Dorado v0.0.7 (Oxford Nanopore Technologies) on a GPU-enabled computing cluster. The violin plots of mRNA poly(A) tail length distributions were computed using the ggplot2 package in R v4.3.2.
Resource availability
Further information and requests for resources and reagents should be directed to and will be fulfilled by E.V., subject to a completed materials transfer agreement.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Data availability
The data supporting the findings of this study are available from the corresponding authors upon request. The generated cryo-EM maps and PDB codes associated with different structures are deposited in the EMDB and PDB databases under accession codes 47689 / 47690 and 9E7T / 9E7U, respectively. The nanopore direct RNA sequencing data are deposited in the NCBI Sequence Read Archive (SRA) database under accession code PRJNA1263024. All raw data associated with the gel images and deadenylation assays is provided in Source Data file.
References
Eichner, H., Karlsson, J. & Loh, E. The emerging role of bacterial regulatory RNAs in disease. Trends Microbiol. 30, 959–972 (2022).
Halfmann, A., Kovács, M., Hakenbeck, R. & Brückner, R. Identification of the genes directly controlled by the response regulator CiaR in Streptococcus pneumoniae: five out of 15 promoters drive expression of small non-coding RNAs. Mol. Microbiol. 66, 110–126 (2007).
Ensminger, A. W. Legionella pneumophila, armed to the hilt: justifying the largest arsenal of effectors in the bacterial world. Curr. Opin. Microbiol. 29, 74–80 (2016).
Qiu, J. & Luo, Z.-Q. Legionella and Coxiella effectors: strength in diversity and activity. Nat. Rev. Microbiol. 15, 591–605 (2017).
Du Toit, A. The effector repertoire of Legionella. Nat. Rev. Microbiol. 17, 126 (2019).
Herkt, C. E. et al. A MicroRNA network controls Legionella pneumophila replication in human macrophages via LGALS8 and MX1. MBio 11, e03155 (2020).
Sahr, T. et al. Translocated Legionella pneumophila small RNAs mimic eukaryotic microRNAs targeting the host immune response. Nat. Commun. 13, 762 (2022).
Toinon, J. et al. A Legionella pneumophila effector impedes host gene silencing to promote virulence. bioRxiv 2022.11.16.516792 https://doi.org/10.1101/2022.11.16.516792 (2022).
Fontana, M. F., Shin, S. & Vance, R. E. Activation of host mitogen-activated protein kinases by secreted Legionella pneumophila effectors that inhibit host protein translation. Infect. Immun. 80, 3570–3575 (2012).
Subramanian, A. et al. A Legionella toxin exhibits tRNA mimicry and glycosyl transferase activity to target the translation machinery and trigger a ribotoxic stress response. Nat. Cell Biol. 25, 1600–1615 (2023).
Syriste, L. et al. An acetyltransferase effector conserved across Legionella species targets the eukaryotic eIF3 complex to modulate protein translation. MBio 15, e0322123 (2024).
Fontana, M. F. et al. Secreted bacterial effectors that inhibit host protein synthesis are critical for induction of the innate immune response to virulent Legionella pneumophila. PLoS Pathog. 7, e1001289 (2011).
Shimo, H. M. et al. Role of the citrus sinensis RNA deadenylase CsCAF1 in citrus canker resistance. Mol. Plant Pathol. 20, 1105–1118 (2019).
Ninio, S., Celli, J. & Roy, C. R. A Legionella pneumophila effector protein encoded in a region of genomic plasticity binds to Dot/Icm-modified vacuoles. PLoS Pathog. 5, e1000278 (2009).
Mount, H. O. et al. A comprehensive two-hybrid analysis to explore the Legionella pneumophila effector-effector interactome. mSystems 9, e0100424 (2024).
Mount, H. O., Urbanus, M. L., Zangari, F., Gingras, A.-C. & Ensminger, A. W. The Legionella pneumophila effector PieF modulates mRNA stability through association with eukaryotic CCR4-NOT. mSphere 10, e0089124 (2025).
Raisch, T. & Valkov, E. Regulation of the multisubunit CCR4-NOT deadenylase in the initiation of mRNA degradation. Curr. Opin. Struct. Biol. 77, 102460 (2022).
Poetz, F. et al. RNF219 attenuates global mRNA decay through inhibition of CCR4-NOT complex-mediated deadenylation. Nat. Commun. 12, 7175 (2021).
Raisch, T. et al. Reconstitution of recombinant human CCR4-NOT reveals molecular insights into regulated deadenylation. Nat. Commun. 10, 3173 (2019).
Bianchin, C., Mauxion, F., Sentis, S., Séraphin, B. & Corbo, L. Conservation of the deadenylase activity of proteins of the Caf1 family in human. RNA 11, 487–494 (2005).
Petit, A.-P. et al. The structural basis for the interaction between the CAF1 nuclease and the NOT1 scaffold of the human CCR4-NOT deadenylase complex. Nucleic Acids Res. 40, 11058–11072 (2012).
Chen, Y., Khazina, E., Izaurralde, E. & Weichenrieder, O. Crystal structure and functional properties of the human CCR4-CAF1 deadenylase complex. Nucleic Acids Res. 49, 6489–6510 (2021).
Lau, N.-C. et al. Human Ccr4-Not complexes contain variable deadenylase subunits. Biochem. J. 422, 443–453 (2009).
Virtanen, A., Henriksson, N., Nilsson, P. & Nissbeck, M. Poly(A)-specific ribonuclease (PARN): an allosterically regulated, processive and mRNA cap-interacting deadenylase. Crit. Rev. Biochem. Mol. Biol. 48, 192–209 (2013).
van Kempen, M. et al. Fast and accurate protein structure search with Foldseek. Nat. Biotechnol. 42, 243–246 (2024).
Jonstrup, A. T., Andersen, K. R., Van, L. B. & Brodersen, D. E. The 1.4-A crystal structure of the S. pombe Pop2p deadenylase subunit unveils the configuration of an active enzyme. Nucleic Acids Res. 35, 3153–3164 (2007).
Park, J.-E., Yi, H., Kim, Y., Chang, H. & Kim, V. N. Regulation of poly(A) tail and translation during the somatic cell cycle. Mol. Cell 62, 462–471 (2016).
Beck, M. et al. The quantitative proteome of a human cell line. Mol. Syst. Biol. 7, 549 (2011).
Blower, T. R., Salmond, G. P. C. & Luisi, B. F. Balancing at survival’s edge: the structure and adaptive benefits of prokaryotic toxin-antitoxin partners. Curr. Opin. Struct. Biol. 21, 109–118 (2011).
Sofos, N., Xu, K., Dedic, E. & Brodersen, D. E. Cut to the chase-regulating translation through RNA cleavage. Biochimie 114, 10–17 (2015).
Mattison, K., Wilbur, J. S., So, M. & Brennan, R. G. Structure of FitAB from Neisseria gonorrhoeae bound to DNA reveals a tetramer of toxin-antitoxin heterodimers containing pin domains and ribbon-helix-helix motifs. J. Biol. Chem. 281, 37942–37951 (2006).
Dienemann, C., Bøggild, A., Winther, K. S., Gerdes, K. & Brodersen, D. E. Crystal structure of the VapBC toxin-antitoxin complex from Shigella flexneri reveals a hetero-octameric DNA-binding assembly. J. Mol. Biol. 414, 713–722 (2011).
Boamah, D. K., Zhou, G., Ensminger, A. W. & O’Connor, T. J. From many hosts, one accidental pathogen: the diverse protozoan hosts of Legionella. Front. Cell. Infect. Microbiol. 7, 477 (2017).
Mengue, L. et al. Legionella pneumophila prevents proliferation of its natural host Acanthamoeba castellanii. Sci. Rep. 6, 36448 (2016).
Sol, A. et al. Legionella pneumophila translocated translation inhibitors are required for bacterial-induced host cell cycle arrest. Proc. Natl. Acad. Sci. USA 116, 3221–3228 (2019).
Hwang, S. S. et al. mRNA destabilization by BTG1 and BTG2 maintains T cell quiescence. Science 367, 1255–1260 (2020).
Faraji, F. et al. Post-transcriptional control of tumor cell autonomous metastatic potential by CCR4-NOT deadenylase CNOT7. PLoS Genet. 12, e1005820 (2016).
Sgromo, A. et al. A CAF40-binding motif facilitates recruitment of the CCR4-NOT complex to mRNAs targeted by Drosophila Roquin. Nat. Commun. 8, 14307 (2017).
Caesar, J. et al. SIMPLE 3.0. Stream single-particle cryo-EM analysis in real time. J. Struct. Biol. X 4, 100040 (2020).
Punjani, A., Zhang, H. & Fleet, D. J. Non-uniform refinement: adaptive regularization improves single-particle cryo-EM reconstruction. Nat. Methods 17, 1214–1221 (2020).
Zivanov, J., Nakane, T. & Scheres, S. H. W. A Bayesian approach to beam-induced motion correction in cryo-EM single-particle analysis. IUCrJ 6, 5–17 (2019).
Asarnow, D., Palovcak, E. & Cheng, Y. UCSF pyem v0. 5. Zenodo https://doi.org/10.5281/zenodo.3576630 (2019).
Sanchez-Garcia, R. et al. DeepEMhancer: a deep learning solution for cryo-EM volume post-processing. Commun. Biol. 4, 874 (2021).
Brown, A. et al. Tools for macromolecular model building and refinement into electron cryo-microscopy reconstructions. Acta Crystallogr. Sect. D Biol. Crystallogr. 71, 136–153 (2015).
Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).
Williams, C. J. et al. MolProbity: more and better reference data for improved all-atom structure validation. Protein Sci. 27, 293–315 (2018).
Afonine, P. V. et al. Real-space refinement in PHENIX for cryo-EM and crystallography. Acta Crystallogr. Sect. D Struct. Biol. 74, 531–544 (2018).
Webster, M. W., Stowell, J. A. W., Tang, T. T. L. & Passmore, L. A. Analysis of mRNA deadenylation by multi-protein complexes. Methods 126, 95–104 (2017).
Acknowledgements
We are grateful to Tom Misteli, Bobby Hogg, and Robin Stanley for critical comments on the manuscript. We also thank members of the Valkov and Lea laboratories for helpful discussions, as well as our colleagues at the RNA Biology Laboratory and the Center for Structural Biology for their support and advice. This study was supported by the Intramural Research Program of the National Institutes of Health. F.P. was supported by a Walter Benjamin postdoctoral fellowship from the German Research Foundation (Deutsche Forschungsgemeinschaft) [Project number 531520533].
Funding
Open access funding provided by the National Institutes of Health.
Author information
Authors and Affiliations
Contributions
Conceptualization: E.V. Methodology: Y.L., E.V., J.C.D., D.J.T., C.T.P. Investigation: Y.L., C.T.P., D.J.T., J.C.D., F.P., S.G.T. Resources: F.P., A.L.V. Formal analysis: Y.L., J.C.D., D.J.T., C.T.P., S.G.T., S.M.L., E.V. Validation: Y.L., C.T.P., A.L.V. Visualization: Y.L., J.D., D.J.T., C.T.P., S.M.L., E.V. Data curation: E.V., J.C.D., Y.L. Project administration: E.V. Supervision: E.V., S.M.L. Software: E.V. Writing—original draft: Y.L., J.C.D., D.J.T., E.V., S.M.L. Writing—review and editing: E.V., Y.L., S.M.L., C.T.P., J.C.D.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Ditlev Brodersen and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Levdansky, Y., Deme, J.C., Turner, D.J. et al. Intracellular pathogen effector reprograms host gene expression by inhibiting mRNA decay. Nat Commun 16, 6452 (2025). https://doi.org/10.1038/s41467-025-61194-2
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-61194-2