The enzyme 6-phosphofructokinase (PFK) phosphorylates d-fructose 6-phosphate, producing d-fructose 1,6-bisphosphate. The canonical version—discovered almost 90 years ago—is ATP-dependent, allosterically regulated and catalyses the first committed step in glycolysis. However, beyond this textbook enzyme, there is fascinating functional and structural variety among PFKs across the tree of life. While PFKs are found in two non-homologous superfamilies, here, we review the universally distributed enzymes in one, the Phosphofructokinase Superfamily. We focus on summarising the diversity within this superfamily. A key partition regards the identity of the phosphate donor, which can be ATP or inorganic pyrophosphate (PPi). Considerable insights into functional and evolutionary aspects of the ATP- and PPi-dependent PFKs have come through structural biology, with 45 structures now available in the Protein Data Bank. One recent highlight was the use of cryoEM and molecular dynamics simulations to illuminate the structural basis of allosteric regulation in human liver PFK. Others were to explore interactions of drug-like small molecules with the PFKs from Trypanosoma brucei and human liver, revealing new routes to antibiotics and immune modulators, respectively. In contrast with the ATP-dependent enzymes, PPi-dependent PFKs are typically non-allosteric and catalyse a readily reversible reaction. Some also play an additional physiological role by phosphorylating d-sedoheptulose 7-phosphate. We discuss why these properties are plausibly ancestral. Finally, we also emphasise how much remains to be discovered. For example, the 45 experimentally determined structures are from only 14 species. Nine decades in, it is still a great time to be studying PFK.

The enzyme 6-phosphofructokinase (PFK) was discovered almost 90 years ago in the laboratory overseen by Jakub Karol Parnas [1]. d-Fructose 1,6-bisphosphate (FBP)—the Harden–Young ester—had already been isolated and, in fact, it was the first metabolic intermediate ever to be identified [2]. Thus, it was a major milestone in elucidating the Embden–Meyerhof–Parnas pathway of glycolysis when PFK was revealed as the enzyme responsible for forming FBP, from d-fructose 6-phosphate (F6P) [3].

The earliest characterised versions of PFK use ATP to phosphorylate F6P and are members of Enzyme Commission (EC) class 2.7.1.11. These ATP-dependent PFKs also play important regulatory roles in controlling flux through glycolysis. They catalyse the first committed step in the pathway, thereby committing the cell to use glucose for producing energy. As originally discussed in exquisite detail by Monod and his co-workers, who were studying the enzyme from Escherichia coli, PFKs are textbook examples of co-operative kinetics (with respect to F6P concentration) and allosteric regulation [4]. ADP and GDP are powerful allosteric activators of E. coli PFK (EcoPFK), while phosphoenolpyruvate is an allosteric inhibitor [4]. This allosteric regulation tunes glycolytic flux to reflect the cell’s energy and metabolic requirements.

Given PFK’s role at the critical control point in glycolysis, and as a paradigm of allostery, it is not surprising that it was also one of the first enzymes to have its structure determined by X-ray crystallography and then to be probed by site-directed mutagenesis. In a pioneering series of papers, Evans and his group determined structures of the homotetrameric PFKs from the bacteria Geobacillus stearothermophilus and E. coli [5-7]. Each subunit in the tetramer comprises two Rossmann fold domains. The larger domain binds the substrate ATP, while the smaller one binds F6P (Figure 1B). The F6P-binding domain also contains the binding site for allosteric effectors, at the C-terminal helix (Figure 1B). In the tetramer, each subunit mainly contacts two others, with one contact contributing to the effector site, and the other contributing to the active site (Figure 1C and D). The structures revealed the key active site residues, which were subsequently probed by site-directed mutagenesis of EcoPFK [8,9]. These mutagenesis studies confirmed a catalytic mechanism for phosphoryl transfer involving direct nucleophilic attack of the F6P 1-hydroxyl on the γ-phosphate of ATP. A general base catalyst (Asp127 in EcoPFK) increases the nucleophilicity of the F6P hydroxyl group by abstracting its proton (Figure 1A).

Structure and function in the bacterial ATP-dependent PFKs.

Figure 1:
Structure and function in the bacterial ATP-dependent PFKs.

(A) Simplified mechanism for PFK-catalysed phosphoryl transfer. The general base catalyst is Asp127 in both the E. coli and G. stearothermophilus PFKs (EcoPFK and GstPFK, respectively). Arg162´ from a neighbouring subunit, which interacts with F6P in the active R state but not the T state, is also shown. In the R state GstPFK structure (PDB ID: 4PFK; [6]), the two N–O distances between Arg162´ and the nearest phosphate oxygens on F6P are ~3 Å. (B) Two-domain structure of one protomer from GstPFK (PDB ID: 4PFK), with F6P, ADP and Mg2+ in the active site (rectangle), and ADP and Mg2+ in the effector site (oval). (C) The GstPFK tetramer (PDB ID: 4PFK), with each chain coloured differently. The oval highlights the location of the same effector site from (B), now revealed to be at the interface of the subunits coloured green and pink. (D) A second orientation of the tetramer from panel (C), emphasising the location of the active site from (B) at a different interface (green- and blue-coloured subunits) from the effector site in (C). F6P, fructose 6-phosphate; PFK, 6-phosphofructokinase.

Figure 1:
Structure and function in the bacterial ATP-dependent PFKs.

(A) Simplified mechanism for PFK-catalysed phosphoryl transfer. The general base catalyst is Asp127 in both the E. coli and G. stearothermophilus PFKs (EcoPFK and GstPFK, respectively). Arg162´ from a neighbouring subunit, which interacts with F6P in the active R state but not the T state, is also shown. In the R state GstPFK structure (PDB ID: 4PFK; [6]), the two N–O distances between Arg162´ and the nearest phosphate oxygens on F6P are ~3 Å. (B) Two-domain structure of one protomer from GstPFK (PDB ID: 4PFK), with F6P, ADP and Mg2+ in the active site (rectangle), and ADP and Mg2+ in the effector site (oval). (C) The GstPFK tetramer (PDB ID: 4PFK), with each chain coloured differently. The oval highlights the location of the same effector site from (B), now revealed to be at the interface of the subunits coloured green and pink. (D) A second orientation of the tetramer from panel (C), emphasising the location of the active site from (B) at a different interface (green- and blue-coloured subunits) from the effector site in (C). F6P, fructose 6-phosphate; PFK, 6-phosphofructokinase.

Close modal

Iconically, Evans et al. determined G. stearothermophilus PFK (GstPFK) structures in both the active, high-affinity R state and the allosterically inhibited T state. The R-to-T transition is associated with a 7° rigid body rotation of one dimer of subunits, with respect to the other dimer. This alters interactions at subunit interfaces, which are communicated to the active sites in the tetramer [10]. In the R state, Arg162´ (with the prime indicating an amino acid from a neighbouring subunit) is one of four residues that bind the phosphate of F6P, via a salt bridge (Figure 1A). However, Arg162´ is located at the end of a helical turn that unwinds upon transition to the T state. This causes Arg162´ to swing away so that it is replaced in the active site by Glu161´. The arrival of a negatively charged side chain in the phosphate binding site causes reduced affinity for F6P in the T state. This unwinding is also affected by the binding of molecules in the effector site: unwinding is prevented by the binding of ADP [10].

However, there is significantly more diversity among the structures, functions and regulatory mechanisms of PFK enzymes than these classic studies (and most textbooks) suggest. The purpose of this review is first to differentiate between the two non-homologous superfamilies of PFKs and then to summarise the diversity within the predominant superfamily. While we discuss the diverse sequences, functions and evolutionary events within the superfamily, a primary focus is on insights that are emerging from the increasing number of experimentally determined PFK structures.

PFKs are found in two non-homologous superfamilies

The focus of this review is on the broadly distributed, Rossmann fold-containing superfamily of PFKs. In the InterPro database [11], this group of homologous enzymes is named the Phosphofructokinase Superfamily and it is database entry IPR035966. This superfamily is sometimes referred to as the PFKA family [12,13]. However, it is also important to note that there is a second class of PFKs, sometimes named the PFKB family [12,13]. The PFKB enzymes are not homologues of superfamily IPR035966 (Figure 2). Instead, they are members of the Ribokinase-Like Superfamily of sugar kinases [15], which has InterPro entry number IPR029056. This superfamily contains the minor ATP-dependent PFK from E. coli (PFK-2), as well as the ADP-dependent PFKs from archaea (EC 2.7.1.146) [16]. The Ribokinase-Like Superfamily also contains bacterial and eukaryotic 1-phosphofructokinases (enzymes that phosphorylate fructose 1-phosphate to form FBP), ribokinases, phosphotagatokinases, adenosine kinases and others [15,17]. The interested reader is referred elsewhere for more detailed discussions of the Ribokinase-Like Superfamily [16-18].

A PFK from the Ribokinase-Like Superfamily.

Figure 2:
A PFK from the Ribokinase-Like Superfamily.

Chain A of the homodimeric E. coli PFK-2 (PDB ID: 3UQD; [14]) is shown. The active site contains ATP and F6P (white sticks) and two Mg2+ ions (green spheres) and is located in a Rossmann-like, three-layered α/β/α sandwich domain that characterises the Ribokinase-Like Superfamily. Comparison with Figure 1B reveals the overall lack of similarity with the two-Rossmann structures of the Phosphofructokinase Superfamily. F6P, fructose 6-phosphate; PFK, 6-phosphofructokinase.

Figure 2:
A PFK from the Ribokinase-Like Superfamily.

Chain A of the homodimeric E. coli PFK-2 (PDB ID: 3UQD; [14]) is shown. The active site contains ATP and F6P (white sticks) and two Mg2+ ions (green spheres) and is located in a Rossmann-like, three-layered α/β/α sandwich domain that characterises the Ribokinase-Like Superfamily. Comparison with Figure 1B reveals the overall lack of similarity with the two-Rossmann structures of the Phosphofructokinase Superfamily. F6P, fructose 6-phosphate; PFK, 6-phosphofructokinase.

Close modal

The murky world of the Phosphofructokinase Superfamily

Phylogenetic analyses continually confirm a common evolutionary ancestor for the members of the Phosphofructokinase Superfamily [12,19-25]. At the same time, complex patterns of evolution have yielded substantial diversity in the specificity, regulation and structural features of superfamily members. In general, classifying and organising the superfamily has therefore proven complicated.

A significant partition in the superfamily arises from the identity of the phosphate donor. All PFKs were assumed to utilise ATP until 1974, when an enzyme from the amoeba, Entamoeba histolytica, was found to use inorganic pyrophosphate (PPi) instead [26]. More PPi-dependent PFKs have since been discovered in plants, anaerobic bacteria, archaea and unicellular eukaryotes. These enzymes belong to EC class 2.7.1.90, rather than EC 2.7.1.11 like the ATP-dependent PFKs. While the EC numbers are useful descriptors, one caveat is that they are based on function and do not imply homology. Thus ATP-dependent PFKs from the Ribokinase-Like Superfamily, such as E. coli PFK-2, are also members of EC class 2.7.1.11. At the time of writing (November 2024), the BRaunschweig ENzyme DAtabase (BRENDA) lists 381 functional parameters for the PPi-dependent PFKs in EC 2.7.1.90, compared with 717 for the ATP-dependent enzymes in EC 2.7.1.11 [27].

Many ATP-dependent eukaryotic PFKs are significantly more complex than their bacterial homologues such as GstPFK (Figure 1B–D). Duplication and tandem fusion of an ancestral, bacterial-like pfk gene means that the eukaryotic genes are approximately twice as long [28,29]. Subsequent gene duplications in different eukaryotic lineages have also given rise to homologous subunit types and/or tissue-specific isoforms. For example, most yeast PFKs are hetero-oligomers of homologous α and β subunits, often active as α4β4 octamers [30]. Mammals have liver-, muscle- and platelet-specific PFK isoforms (PFKL, PFKM and PFKP, respectively), with subunits forming homo- and hetero-oligomers that include tetramers and higher order assemblies [30,31]. And a whole genome duplication in Quercus rubra, the northern red oak, gave rise to three pairs of pfk paralogues [32]. Finally, an additional gene fusion event occurred during the evolution of yeast PFKs, meaning that the enzymes from Saccharomyces cerevisiae (ScePFK) and Komagataella (formerly Pichia) pastoris (KpaPFK) have an extra N-terminal domain [29]. X-ray crystallography revealed this domain to be a distant structural relative of glyoxalase I in KpaPFK [33].

In terms of surveying this sequence diversity, a key early study performed phylogenetic analyses using 152 PFK proteins [22]. The authors defined eight monophyletic sequence groups. A more recent study sought to place the 11 Trichomonas vaginalis PFKs within their phylogenetic context [25]. A tree constructed from the T. vaginalis PFKs, plus another 180 diverse sequences, now revealed 12 groups, albeit in a broadly similar topology to the earlier work [22].

Together, these two studies emphasised the complex evolutionary history of the Phosphofructokinase Superfamily. They uncovered monophyletic groups containing both ATP-dependent and PPi-dependent PFKs. This is consistent with multiple independent evolutionary events that led to changes in the phosphate donor, in different lineages [22,25,34]. Horizontal gene transfer has also been pervasive. In general, the phylogeny of PFKs does not match the phylogeny of the species in which they are found. For example, five of the eight sequence groups defined by Bapteste et al. [22] contained PFKs from prokaryotes and eukaryotes and there was evidence for transfer both to, and from, bacteria.

The complicated patterns of pfk evolution tend to obfuscate classifications in sequence databases. While the homologous superfamily is well defined as IPR035966 in InterPro, grouping into smaller sequence-based families paints a confusing picture. In InterPro, the Phosphofructokinase Superfamily is currently subdivided into 13 families (Table 1). However, the names of these families are overlapping, and it is difficult to understand what distinguishes any given family. For example, there is an InterPro family named ‘ATP-dependent 6-phosphofructokinase, prokaryotic-type’, and another named ‘ATP-dependent 6-phosphofructokinase, prokaryotic’ (Table 1). Both families contain many of the same proteins. In a similar vein, InterPro family IPR022953 is named ‘ATP-dependent 6-phosphofructokinase’ but contains both ATP-dependent and PPi-dependent PFKs. And family IPR012004 is named ‘Pyrophosphate-dependent phosphofructokinase TP0108-type’ but also contains ATP-dependent PFKs.

Table 1:
InterPro families nested within the Phosphofructokinase Superfamily (IPR035966).
InterPro familyFamily name
IPR022953 ATP-dependent 6-phosphofructokinase 
IPR012003 ATP-dependent 6-phosphofructokinase, prokaryotic type 
IPR012828 ATP-dependent 6-phosphofructokinase, prokaryotic 
IPR009161 ATP-dependent 6-phosphofructokinase, eukaryotic type 
IPR041914 ATP-dependent 6-phosphofructokinase, vertebrate type 
IPR012004 Pyrophosphate-dependent phosphofructokinase TP0108 type 
IPR050929 Phosphofructokinase type A 
IPR011183 Pyrophosphate-dependent phosphofructokinase PfpB 
IPR011404 Pyrophosphate–fructose 6-phosphate 1-phosphotransferase 
IPR012829 Phosphofructokinase mixed-substrate PFK group III 
IPR054846 PPi-dependent phosphofructokinase 
IPR011405 Pyrophosphate-dependent phosphofructokinase SMc01852 type 
IPR011403 Pyrophosphate-dependent phosphofructokinase TM0289 type 
InterPro familyFamily name
IPR022953 ATP-dependent 6-phosphofructokinase 
IPR012003 ATP-dependent 6-phosphofructokinase, prokaryotic type 
IPR012828 ATP-dependent 6-phosphofructokinase, prokaryotic 
IPR009161 ATP-dependent 6-phosphofructokinase, eukaryotic type 
IPR041914 ATP-dependent 6-phosphofructokinase, vertebrate type 
IPR012004 Pyrophosphate-dependent phosphofructokinase TP0108 type 
IPR050929 Phosphofructokinase type A 
IPR011183 Pyrophosphate-dependent phosphofructokinase PfpB 
IPR011404 Pyrophosphate–fructose 6-phosphate 1-phosphotransferase 
IPR012829 Phosphofructokinase mixed-substrate PFK group III 
IPR054846 PPi-dependent phosphofructokinase 
IPR011405 Pyrophosphate-dependent phosphofructokinase SMc01852 type 
IPR011403 Pyrophosphate-dependent phosphofructokinase TM0289 type 

In summary, it is clear that names, sequences and genome annotations can be murky in the world of PFKs. As more PFK sequences are discovered, the boundaries between previously defined groups are tending to become blurrier. It is increasingly difficult to classify PFKs based on sequence alone. The most meaningful classifications are likely to be based on structural and functional features, as discussed in the following sections.

Diversity in function, regulation and physiological roles

Hundreds of PFKs have been biochemically characterised in the past nine decades. At the time of writing, the BRENDA database [27] contained information on ATP-dependent PFKs (EC 2.7.1.11) from 180 different species, as well as PPi-dependent PFKs (EC 2.7.1.90) from 108 species. We used BRENDA as a starting point to review the range of steady-state kinetics parameters reported for members of the Phosphofructokinase Superfamily. We inspected the entries in BRENDA to identify wild-type enzymes that had been assayed for the formation of FBP, from F6P and the predominant phosphate donor (Mg2+-ATP for EC 2.7.1.11; Mg2+-PPi for EC 2.7.1.90), under physiologically relevant conditions. For the ATP-dependent enzymes showing cooperative kinetics, Michaelis constants (KM values) were often determined in the presence of an activator such as AMP [35], thus yielding hyperbolic saturation kinetics. Overall, this snapshot revealed that the ranges for turnover number (kcat) and for the two KM values (F6P and the phosphate donor) are remarkably similar for ATP- and PPi-dependent PFKs (Table 2). The median kcat value in BRENDA was ~110 s-1 for each class of PFK. The median KM values for F6P were both KMF6P = 0.3 mM. With respect to the phosphate donors, the median KMATP was ~0.05 mM and the median KMPPi was ~0.03 mM. In 2018, a global survey of all wild-type enzymes in BRENDA revealed a median kcat of ~10 s-1 and a median KM of ~0.14 mM [46]. Thus, compared with other enzymes, PFKs tend to have higher-than-average turnover numbers and average Michaelis constants.

Table 2:
Indicative steady-state kinetic parameters for wild-type members of the Phosphofructokinase Superfamily catalysing the formation of FBP from F6P.
ParameterValueEnzyme sourceReference
ATP-dependent PFKs 
 Low kcat 49 s-1 Escherichia coli [36
 High kcat 360 s-1 Homo sapiens (muscle isoform) [37
 Low KMF6P 0.016 mM Ricinus communis (castor bean) [38
 High KMF6P 7 mM Fasciola hepatica (liver fluke) [35
 Low KMATP 0.005 mM Cucumis sativus (cucumber) [39
 High KMATP 0.82 mM Bacillus methanolicus [40
PPi-dependent PFKs 
 Low kcat 16 s-1 Methylobacterium nodulans [24
 High kcat 330 s-1 Entamoeba histolytica [41
 Low KMF6P 0.01 mM Naegleria fowleri [42
 High KMF6P 3.8 mM Oryza sativa (rice) [43
 Low KMPPi 0.005 mM Propionibacterium freudenreichii [44
 High KMPPi 0.58 mM Sanseviera trifasciata (snake plant) [45
ParameterValueEnzyme sourceReference
ATP-dependent PFKs 
 Low kcat 49 s-1 Escherichia coli [36
 High kcat 360 s-1 Homo sapiens (muscle isoform) [37
 Low KMF6P 0.016 mM Ricinus communis (castor bean) [38
 High KMF6P 7 mM Fasciola hepatica (liver fluke) [35
 Low KMATP 0.005 mM Cucumis sativus (cucumber) [39
 High KMATP 0.82 mM Bacillus methanolicus [40
PPi-dependent PFKs 
 Low kcat 16 s-1 Methylobacterium nodulans [24
 High kcat 330 s-1 Entamoeba histolytica [41
 Low KMF6P 0.01 mM Naegleria fowleri [42
 High KMF6P 3.8 mM Oryza sativa (rice) [43
 Low KMPPi 0.005 mM Propionibacterium freudenreichii [44
 High KMPPi 0.58 mM Sanseviera trifasciata (snake plant) [45

Allosteric regulation also varies significantly across the Phosphofructokinase Superfamily. In general, the bacterial ATP-dependent PFKs are activated by ADP and inhibited by phosphoenolpyruvate [30]. Some bacterial ATP-dependent PFKs are also inhibited by PPi, often when the organism also possesses a PPi-dependent PFK [47]. However, the strength of this allosteric regulation can vary. For example, the enzyme from Lactobacillus delbrueckii is relatively insensitive to its effectors, ADP and phosphoenolpyruvate, with dissociation constants that are 2–3 orders of magnitude higher than the values for EcoPFK and GstPFK [48].

Allosteric regulation in the eukaryotic ATP-dependent PFKs tends to be significantly more complex and was reviewed by Schöneberg et al. [30]. Over 20 effectors have been identified for various eukaryotic enzymes, with AMP and fructose 2,6-bisphosphate joining ADP as important physiological activators. ATP is both a substrate and an allosteric inhibitor, with metabolites such as citrate and lactate potentiating its inhibitory effect [49]. Recently, a powerful new mass spectrometry method was described for high-throughput discovery of protein-metabolite interactions [50]. This confirmed existing metabolite interactions of the human liver and platelet PFK isoforms, while also providing evidence for a dozen or so new interactions. These interactions—with metabolites such as 2,3-bisphosphoglycerate, CTP and 5-hydroxy-l-tryptophan—require further biochemical investigation; however, the implication is that more allosteric effectors remain to be discovered for the eukaryotic ATP-dependent PFKs.

In contrast with the ATP-dependent enzymes, many PPi-dependent PFKs—including all those characterised from bacteria—are non-allosteric. For example, the PFK from Acetivibrio thermocellus (formerly Clostridium thermocellum) was recently tested in the presence of 11 different metabolites at two different concentrations, and none had a significant effect on activity compared with the ATP-dependent PFK from the same organism [47]. However, fructose 2,6-bisphosphate activates the PPi-dependent PFKs from plants including rice, potato, banana and strawberry [43,51-53]. AMP was also shown to be an activator of the PPi-dependent PFK from the amoeba, Naegleria fowleri [42].

The identity of the phosphate donor has significant metabolic and physiological implications. ATP-dependent PFKs catalyse a reaction (F6P + ATP → FBP + ADP) that is far from equilibrium and therefore effectively irreversible in vivo (although its reverse has been assayed in vitro; [54]). Not only is the forward reaction highly favourable under standard conditions (ΔG°′ = –17.8 ± 1.3 kJ mol-1), but a further thermodynamic driving force comes from the intracellular (ATP)/(ADP) ratio, which is typically around 10 [55,56]. In contrast, the standard Gibbs free energy of PPi-dependent phosphotransfer is smaller (ΔG°′ = –4.6 ± 1.4 kJ mol-1) and the intracellular ratio of PPi to Pi is typically around 0.1. This lower driving force means the glycolytic pathway lies closer to thermodynamic equilibrium in organisms with PPi-dependent PFKs [56]. As a result, PPi-dependent PFKs are often assayed in both the forward (F6P + PPi → FBP + Pi) and reverse (FBP + Pi → F6P+PPi) directions. Careful comparisons of the steady-state kinetic parameters in each direction have been carried out for the PPi-dependent PFKs from the unicellular eukaryote Giardia lamblia [57], and the bacteria Methylobacterium nodulans and Methylosinus trichosporium [24]. For all three enzymes, kcat is identical (or near-identical) in the forward and reverse directions. However, each enzyme shows a lower KM for the bisphosphorylated substrate, FBP, meaning that in each case the overall catalytic efficiency (kcat/KM) is 3- to 13-fold higher for the reverse reaction (Supplementary Table S1).

An interesting corollary is that organisms with a reversible, PPi-dependent PFK could use it for both the glycolytic forward reaction and the gluconeogenic reverse reaction in vivo [12,58]. This is in contrast with organisms with ATP-dependent PFKs, which employ a separate enzyme, fructose-1,6-bisphosphatase (FBPase), for hydrolysing FBP to F6P in gluconeogenesis. Thus, one might expect the gene for FBPase to be inactivated or lost from organisms with PPi-dependent PFKs. While PPi-dependent PFKs have been proposed to play this bifunctional role in Amycolatopsis methanolica [59] and Porphyromonas gingivalis [60], these anaerobic bacteria also contain the gene for a dedicated FBPase. It remains to be determined whether any organism uses a PPi-dependent PFK as its primary enzyme for both glycolysis and gluconeogenesis.

A third physiological role for PPi-dependent PFKs lies in a variation of the classical pentose phosphate pathway. Four decades ago, Reeves et al. discovered that E. histolytica lacks a transaldolase for catalysing sedoheptulose 7-phosphate + glyceraldehyde 3-phosphate → erythrose 4-phosphate + F6P. Instead, the E. histolytica PFK (EhiPFK) phosphorylates the seven-carbon metabolite ᴅ-sedoheptulose 7-phosphate (S7P), forming sedoheptulose 1,7-bisphosphate (SBP). The resulting SBP can be cleaved into erythrose 4-phosphate and dihydroxyacetone phosphate by the glycolytic enzyme, fructose bisphosphate aldolase, effectively replacing the absent transaldolase [61]. More recently, this SBP pathway was also shown to operate in cellulolytic clostridia lacking transaldolase genes [62]. Metabolomics experiments confirmed a significant intracellular SBP pool in Clostridium thermosuccinogenes. The PPi-dependent PFKs from C. thermosuccinogenes and A. thermocellus were also purified and shown to be highly active. Their kcat/KM values were > 106 s-1 M-1 when F6P was the substrate, and reduced by only four- to six-fold with S7P [62].

Some PPi-dependent PFKs also have a fourth physiological role, which is to maintain PPi homeostasis. Most organisms possess a pyrophosphatase (PPase) to hydrolyse PPi and prevent its toxic accumulation. However, A. thermocellus does not contain a PPase and the gene for its PPi-dependent PFK could only be deleted when a gene for a PPase was also introduced [63]. Similarly, in the parasite Toxoplasma gondii, the PPi-dependent PFK—but not the ATP-dependent PFK—was shown to be essential for growth [64]. Depleting the PPi-dependent PFK led to increased PPi levels and reduced protein synthesis by 40%. The fitness defect could be rescued by over-expressing PPase, demonstrating an essential role for the PPi-dependent PFK in maintaining PPi homeostasis [64].

A dive into the structural universe of ATP-dependent PFKs

At the time of writing, there are 45 structures of Phosphofructokinase Superfamily enzymes available in the Protein Data Bank (PDB) [65]. We have compiled basic information about each of these structures in Supplementary Table S2. Perhaps surprisingly—given the central role of PFK in glycolysis and its rich history of biochemical research—the 45 structures are from only 14 different organisms. Forty of the structures are ATP-dependent PFKs and only five are PPi-dependent PFKs.

Within the set of ATP-dependent PFKs, 18 structures are from five species of bacteria and they all have the same architecture first glimpsed by Evans et al. [5,7]. Each enzyme is active as a homotetramer, with subunits of 319 to 330 residues in the different species. The subunit structures from each species are near-identical, with an average RMSD between any two chains of approximately 0.6 Å (Figure 3). GstPFK has remained a particularly important exemplar, with eight structures in the PDB. These have confirmed that the physiologically relevant allosteric regulators, ADP and phosphoenolpyruvate, occupy the same effector site in the R state and T state, respectively [6,10,66].

Many bacterial PFKs have near-identical structures.

Figure 3:
Many bacterial PFKs have near-identical structures.

Structures of the ATP-dependent PFKs from five bacterial species have been determined. Representative subunits from each of these are shown superimposed. The rectangle highlights the active site and the oval highlights the effector site. PFKs, 6-phosphofructokinases.

Figure 3:
Many bacterial PFKs have near-identical structures.

Structures of the ATP-dependent PFKs from five bacterial species have been determined. Representative subunits from each of these are shown superimposed. The rectangle highlights the active site and the oval highlights the effector site. PFKs, 6-phosphofructokinases.

Close modal

The second major group of ATP-dependent PFK structures is from the yeasts S. cerevisiae and K. pastoris, and the mammals Oryctolagus cuniculus (rabbit) and Homo sapiens. There are 16 PFK structures from these eukaryotes in the PDB (Supplementary Table S2). As a result of the tandem duplication in their evolutionary history (vide supra), their subunits are 750–1000 residues in length. The detailed molecular architecture of these enzymes was first revealed in 2011, with near-simultaneous publication of the KpaPFK structure [33] and then the S. cerevisiae and rabbit muscle structures [67]. The latter structure was a homodimer, rather than the physiologically active homotetramer. On the other hand, the yeast enzymes were determined in active, heterooligomeric forms. ScePFK contained two α subunits and two β subunits [67]. Most impressively, the structure of KpaPFK also contains a third subunit, γ, and was solved as a (αβγ)4 dodecamer of almost 1 MDa [33]. The structure also revealed that the γ subunit had most likely evolved from an ancient S-adenosylmethionine-dependent methyltransferase.

The first low-resolution structure of a human PFK (PFKM at 6 Å resolution) was reported in 2014 [68], followed by higher resolution PFKP structures the following year [69,70]. The structure of the human liver isoform, PFKL, was the last of this group to be reported and also the first PFK to be determined by cryoEM rather than X-ray crystallography [71].

Other than the accessory γ subunit of KpaPFK, each subunit of the yeast and mammalian PFKs contains four Rossmann folds (Figure 4A). The two N-terminal Rossmann folds contain the active site. Binding modes for the substrates, F6P and ATP, have been retained from the bacterial ATP-dependent enzymes [33,67]. On the other hand, catalytic activity has been lost from the C-terminal half of each monomer. Instead, this part of the structure is now important for stability [73] and for the complex allosteric regulation of these enzymes (Figure 4B). Schöneberg et al. have written an excellent review of the allosteric effector sites [30], to which the interested reader is referred for more detail. In brief, at the former C-terminal active site the ATP-binding pocket has been lost entirely, while the F6P-binding pocket (the F' site) has evolved to become the sugar-binding effector site that is important for activation by fructose 2,6-bisphosphate (Figure 4A). Two new effector sites have evolved in the yeast and mammalian enzymes, which have no counterparts in the bacterial PFKs (Figure 4B). The first of these, site N1 (also named site 2 in PFKL; [72]) binds ATP as an allosteric inhibitor [33,67]. The second novel site N2 (named site 1 in PFKL) is an activation site that binds ADP [67,72].

The complex architecture of some eukaryotic ATP-dependent PFKs.

Figure 4:
The complex architecture of some eukaryotic ATP-dependent PFKs.

(A) The human PFKP enzyme (PDB ID: 4XZ2, [69]). The four chains in the homotetramer are coloured blue, green, yellow and red. The lighter shaded part of each subunit is the N-terminal half containing the active site. The darker shaded part is the C-terminal (regulatory) half. The active site in the blue subunit is highlighted with an oval. The four F´ regulatory sites (boxed) are on the inner surface of the tetramer and occupied by fructose-1,6-bisphosphate in this structure. (B) Locations of the effector sites in human PFKL [72]. The image shows the T state structure (PDB ID: 8W2H). Chains are coloured as described for panel (A). The oval shows the position of an active site. The rectangles indicate effector sites and are labelled according to the two common naming conventions [69,72]. F6P in the active site and ADP in effector site N2 were positioned by overlaying the R state structure (PDB ID: 8W2G). PFK, 6-phosphofructokinase.

Figure 4:
The complex architecture of some eukaryotic ATP-dependent PFKs.

(A) The human PFKP enzyme (PDB ID: 4XZ2, [69]). The four chains in the homotetramer are coloured blue, green, yellow and red. The lighter shaded part of each subunit is the N-terminal half containing the active site. The darker shaded part is the C-terminal (regulatory) half. The active site in the blue subunit is highlighted with an oval. The four F´ regulatory sites (boxed) are on the inner surface of the tetramer and occupied by fructose-1,6-bisphosphate in this structure. (B) Locations of the effector sites in human PFKL [72]. The image shows the T state structure (PDB ID: 8W2H). Chains are coloured as described for panel (A). The oval shows the position of an active site. The rectangles indicate effector sites and are labelled according to the two common naming conventions [69,72]. F6P in the active site and ADP in effector site N2 were positioned by overlaying the R state structure (PDB ID: 8W2G). PFK, 6-phosphofructokinase.

Close modal

Until recently, no yeast or mammalian PFK had been visualised in both the R state and the T state. Therefore, it was an important advance when Lynch et al. used cryoEM to determine structures of human PFKL in both states, elucidating key details of the R to T transition [72]. As with the R-to-T transition in the bacterial enzyme [10], a 7° rotation disrupts the F6P-binding part of the active site. However, this rotation is between PFKL monomers and along a different axis than the conformational change in GstPFK.

Critically, the T state PFKL structure captured ATP in effector site 3 (Figure 4B), which is in an equivalent location to the effector site in GstPFK. Compared with the PFKL R state (in which site 3 was unoccupied), ATP binding led to localised unfolding of an α helix and shifted the positions of two arginine residues (Arg201 and Arg292). In the R state, these residues interact with the phosphate of the substrate, F6P, in an adjoining active site [72]. Thus, ATP at site 3 is an important inhibitory signal that contributes to disrupting the F6P binding pocket. Furthermore, a phosphate occupies site 3 in previously determined R state PFKP structures [69,70]. Because phosphate ions have been shown to increase PFK activity [74], Lynch et al. suggested that this effector site may play dual activating and inhibitory roles—akin to the single effector site in GstPFK [72].

An interesting corollary is to ask how GstPFK binds ADP to maintain the R state, whereas PFKL binds ATP to maintain the T state. Overlaying the two enzymes reveals quite different binding poses for the two effector molecules. While the phosphates of the two effectors occupy equivalent positions, the adenosine moiety in PFKL is rotated by approximately 180° compared with its position in the GstPFK effector site (Supplementary Figure S1). Thus, the location of the effector site has been conserved through evolution, but the specific interactions made by the bacterial and eukaryotic enzymes have diverged.

The R and T state structures of PFKL have also helped to resolve the role of a C-terminal tail of approximately 20 residues. Deleting this tail was known to lock PFKP in an active conformation [69]. In the PFKL structures, it was disordered in the R state, but in the T state, it adopted an extended conformation that bridged across the regulatory and catalytic domains [72]. Molecular dynamics simulations confirmed that this positioning of the tail stabilises the T state [72]. Interestingly, the C-terminus is also known to become phosphorylated when PFKL is activated in response to insulin signalling [75]. This phosphorylation may displace the tail and favour transition to the active R state [72].

Overall, it is clear that the tandem duplication associated with these eukaryotic PFKs has potentiated the evolution of sophisticated regulatory mechanisms and that future work will continue to reveal more details of these.

Trypanosoma brucei PFK: an unusual structure makes a novel drug target

In addition to the enzymes described above, another well-characterised, eukaryotic, ATP-dependent PFK is from the protozoan that causes African sleeping sickness, Trypanosoma brucei [76]. The T. brucei PFK (TbrPFK) was the first eukaryotic PFK to have its structure determined [77], and there are now six structures available in the PDB (Supplementary Table S2). It is an outlier because it is not a product of the gene duplication that characterises other eukaryotic ATP-dependent PFKs. Instead, it is closer in sequence to the PPi-dependent PFKs [78]. With 487-residue subunits, it also contains extra elements of structure that set it apart from shorter bacterial ATP-dependent PFKs. These structural decorations include a novel N-terminal domain and an extra 20-residue loop that is also found in some PPi-dependent PFKs [77]. The ATP-bound structure of the TbrPFK dimer-of-dimers also revealed a C-terminal extension dubbed the ‘reaching arm’, which adopts a long α-helical conformation, extends across the inter-dimer interface, and forms a lid over the effector binding site [34]. The only known allosteric activator of TbrPFK is AMP, which binds at the same effector site found in ATP-dependent bacterial PFKs [79]. However, AMP is rotated by 180° compared with ADP occupying the equivalent site in EcoPFK.

The glycolytic enzymes of T. brucei have attracted attention as drug targets because the form of the parasite in the bloodstream of infected humans is wholly dependent on glycolysis for producing ATP [80]. Recently, the unusual structural features of TbrPFK were used to guide a noteworthy drug discovery campaign [81]. McNae et al. developed a series of allosteric inhibitors that do not compete with ATP binding in the active site but do prevent a mobile loop from adopting the R state conformation (Figure 5). The inhibitors showed IC50 values as low as 30 nM for inhibiting TbrPFK. Most impressively, the three best compounds had faster parasite kill times than existing drugs in the clinic, did not significantly inhibit human PFKs in vitro and rapidly cleared parasites in a mouse model of T. brucei infection [81].

Structure-based design of TbrPFK inhibitors that cure sleeping sickness in mice.

Figure 5:
Structure-based design of TbrPFK inhibitors that cure sleeping sickness in mice.

(A) The TbrPFK homotetramer, with chains coloured yellow, green, purple and orange. The box shows the location of an active site. The structure (PDB ID: 6QU5) contains ATP, displayed as thick sticks. The dichlorophenyl lead compound, CTCB-12, is shown in pink, with its chlorine atoms in green, in the adjacent allosteric pocket. (B) Zoomed-in view of the boxed region in (A) showing the dichlorophenyl ring of CTCB-12 occupying the allosteric pocket. Two key residues from the mobile activating loop, Asp231 and Leu232, are in black. By blocking the movement of Leu232 into the allosteric pocket, CTCB-12 prevents the mobile loop from adopting its R state conformation in which Asp231, as well as the catalytic base Asp229, are positioned for catalysis. The figure is reproduced from reference [81] under a Creative Commons Attribution 4.0 International License (CC BY 4.0; https://creativecommons.org/licenses/by/4.0/). PFK, 6-phosphofructokinase.

Figure 5:
Structure-based design of TbrPFK inhibitors that cure sleeping sickness in mice.

(A) The TbrPFK homotetramer, with chains coloured yellow, green, purple and orange. The box shows the location of an active site. The structure (PDB ID: 6QU5) contains ATP, displayed as thick sticks. The dichlorophenyl lead compound, CTCB-12, is shown in pink, with its chlorine atoms in green, in the adjacent allosteric pocket. (B) Zoomed-in view of the boxed region in (A) showing the dichlorophenyl ring of CTCB-12 occupying the allosteric pocket. Two key residues from the mobile activating loop, Asp231 and Leu232, are in black. By blocking the movement of Leu232 into the allosteric pocket, CTCB-12 prevents the mobile loop from adopting its R state conformation in which Asp231, as well as the catalytic base Asp229, are positioned for catalysis. The figure is reproduced from reference [81] under a Creative Commons Attribution 4.0 International License (CC BY 4.0; https://creativecommons.org/licenses/by/4.0/). PFK, 6-phosphofructokinase.

Close modal

The PPi-dependent PFKs: decorated structures and insights into phosphate donor specificity

There are currently five structures of PPi-dependent PFKs available in the PDB (Supplementary Table S2). Four are from bacteria: Borrelia burgdorferi (two structures); Nitrosospira multiformis; and Marinobacter aquaeolei. The fifth is one of the PFK isoforms from the protozoan, T. vaginalis. Of these, just one of the B. burgdorferi PFK (BbuPFK) structures has an associated publication [13], although the T. vaginalis structure (TvaPFK) was only deposited in late 2024. The PPi-dependent PFKs have the two-Rossmann architecture of their ATP-dependent bacterial homologues. However, a variety of insertions and extensions mean that their subunits are also larger than those shown in Figure 3.

The landmark study on BbuPFK, the enzyme from the Lyme disease pathogen, revealed a highly decorated homodimer of 555-residue subunits [13]. The two Rossmann-containing domains adopt a highly similar structure to EcoPFK (RMSD = 1.2 Å). However, BbuPFK has extensions of 70 residues at the N-terminus and 34 residues at the C-terminus, with the latter structural element disrupting the effector binding site seen in EcoPFK (Figure 6A). BbuPFK also contains an 83-residue insertion that forms a novel domain, comprising four α helices and a β hairpin that is positioned over the active site (Figure 6A). The β hairpin plays a role in conferring PPi specificity, with the side chain of His384 positioned such that it would block binding of an ATP ribose group [13].

Superposition of PPi-dependent PFKs with the ATP-dependent E. coli enzyme.

Figure 6:
Superposition of PPi-dependent PFKs with the ATP-dependent E. coli enzyme.

(A) A protomer of BbuPFK (PDB ID: 2F48) in purple, aligned with EcoPFK (PDB ID: 1PFK) in grey. In BbuPFK, the C-terminal extension and the helical domain insertion are shown in a darker shade. The effector site in EcoPFK, with ADP and Mg2+ bound, is indicated by the rectangle. This site is disrupted by the C-terminal extension in BbuPFK. The β hairpin from the BbuPFK domain insertion, which closes over the active site, is shown in the oval. (B) Superposition of TvaPFK (PDB ID: 9DQM) in pink with EcoPFK (PDB ID: 1PFK) in grey, in the region of the effector site. The TvaPFK structure has AMP (pink sticks) bound in a putative effector site that is distinct from the site where ADP (white sticks) binds in EcoPFK. The TvaPFK structure also includes a bound phosphate (circled) that overlays the α-phosphate of ADP in the EcoPFK structure. PFK, 6-phosphofructokinase; PPi, inorganic pyrophosphate.

Figure 6:
Superposition of PPi-dependent PFKs with the ATP-dependent E. coli enzyme.

(A) A protomer of BbuPFK (PDB ID: 2F48) in purple, aligned with EcoPFK (PDB ID: 1PFK) in grey. In BbuPFK, the C-terminal extension and the helical domain insertion are shown in a darker shade. The effector site in EcoPFK, with ADP and Mg2+ bound, is indicated by the rectangle. This site is disrupted by the C-terminal extension in BbuPFK. The β hairpin from the BbuPFK domain insertion, which closes over the active site, is shown in the oval. (B) Superposition of TvaPFK (PDB ID: 9DQM) in pink with EcoPFK (PDB ID: 1PFK) in grey, in the region of the effector site. The TvaPFK structure has AMP (pink sticks) bound in a putative effector site that is distinct from the site where ADP (white sticks) binds in EcoPFK. The TvaPFK structure also includes a bound phosphate (circled) that overlays the α-phosphate of ADP in the EcoPFK structure. PFK, 6-phosphofructokinase; PPi, inorganic pyrophosphate.

Close modal

The BbuPFK structure also shed light on previous sequence analyses [19] and mutagenesis experiments [82], which had identified putative signatures of phosphate donor specificity (Supplementary Table S3). The first of these is an invariant glycine in ATP-dependent PFKs, found in the motif GGDG. BbuPFK and other PPi-dependent PFKs possess GGDD instead. The negatively charged aspartate that differs between the two (Asp177 in BbuPFK) occupies the site of the ATP α-phosphate. Not only does this prevent ATP from binding, but it also ensures that PPi binds in the correct register; that is, it binds in a position corresponding to the β- and γ-phosphates of ATP [13]. Secondly, in BbuPFK, Lys203 is found in the PPi-PFK signature motif, PKTIDXD. Glycine usually replaces the lysine in ATP-dependent PFKs [82]. In the solved structure of BbuPFK, Lys203 makes a salt bridge with the sulfate in the active site that mimics one of the phosphates of PPi. More specifically, the sulfate occupies the position where the β-phosphate of ATP would be found in an ATP-dependent PFK [13]. In an unpublished BbuPFK structure (PDB ID: 2F48), the lysine is instead 3.1 Å from the phosphate at the 1-position of the bound product, FBP. This appears to confirm a role for Lys203 in positioning PPi and also in aiding with polarisation of the phosphoanhydride bond that is cleaved during catalysis [13]. Interestingly, the ‘PPi-like’ ATP-dependent enzyme, TbrPFK, also has a lysine in this position (Supplementary Table S3), showing that the P(G/K)TIDXD motif is not, a priori, an indicator of phosphate donor specificity [34]. Finally, the BbuPFK structure highlighted the role of Asn181 in sterically preventing the adenine moiety of ATP from binding in the active site [13]. In all solved structures of ATP-dependent PFKs, the equivalent residue is a glycine. While it is not conserved as asparagine in PPi-dependent PFKs, it is always larger than glycine (Supplementary Table S3).

TvaPFK (denoted TvPPi-PFK1 in [25]) was first purified over 25 years ago and is a homotetramer of 434-residue subunits [19]. While a full description of its structure (PDB ID: 9DQM) must necessarily await an accompanying publication, it is tantalising to note that this eukaryotic PPi-dependent enzyme contains extra helices in the FBP-binding domain. These contribute to a binding site that is occupied by AMP in the deposited structure (Figure 6B) and which appears to be in a similar location to site N1 in the eukaryotic ATP-dependent enzymes (Figure 4B). Further, an inorganic phosphate in the TvaPFK structure occupies the same position as the α-phosphate of ADP in the bacterial effector site (Figure 6B), perhaps making it equivalent to site 3 in human PFKL (Figure 4B).

The nature of the ancestral PFK

In the sections above, we summarised the enormous diversity in phosphate donor specificity, reaction specificity, allosteric regulation and structural innovations in the Phosphofructokinase Superfamily, across the tree of life. A longstanding question therefore concerns the nature of the superfamily’s common ancestor [12,58,83]. ATP-dependent bacterial enzymes such as EcoPFK have the smallest subunits of the known PFKs, and in their classic phylogenetic study, Bapteste et al. used parsimony to conclude that the ancestral PFK was also ATP-dependent [22]. On the other hand, EcoPFK is allosterically regulated, and allostery is a derived rather than ancestral trait [84]. Thus, others have hypothesised that the ancestral PFK was non-allosteric and PPi-dependent [20,85], despite the structural decorations on modern-day enzymes such as BbuPFK. Of note, the non-allosteric PPi-dependent PFKs from the archaeon, Thermoproteus tenax, and the bacterium, Dictyoglomus thermophilum, are minimally sized homodimers that are approximately the same size as EcoPFK (337 and 345 residues per subunit, versus 319 for EcoPFK), and may represent descendants of the most ancient lineage [20,86].

In eukaryotes, there is strong evidence for frequent phosphate donor switching during evolution. Activity with ATP was introduced into the PPi-dependent EhiPFK by a single point mutation, showing that this enzyme contains a latent ATP-binding site [82]. In the other direction, sequence and structure analyses have shown that the ATP-dependent TbrPFK evolved from a PPi-dependent ancestor [34]. However, it is useful to recall that eukaryotes only arose about two billion years ago [87]; by the time they appeared, prokaryotes had already been evolving for over two billion years [88].

Glycolysis and gluconeogenesis are among the most ancient metabolic pathways on Earth, and they originated when all life was strictly anaerobic [89,90]. The phylogenetic distribution of extant PPi-dependent PFKs—across eukaryotes, bacteria and archaea—also correlates with an anaerobic lifestyle. Having this type of PFK is beneficial in fermentative metabolism because, compared with ATP-dependent PFKs, PPi-dependent PFKs deliver a net yield of one extra ATP per molecule of glucose in glycolysis [58]. In the primordial world, pyrophosphate was also likely to have been present as an essential component of energy metabolism [91]. Thus, in this ancient anaerobic environment, there is likely to have been a strong selective advantage for a PPi-dependent PFK.

Finally, the first organisms had small genomes and, consequently, restricted enzyme repertoires [92]. A widely accepted model posits a patchwork metabolism in these cells, with multifunctional enzymes catalysing reactions in multiple pathways [93,94]. As discussed above, PPi-dependent PFKs catalyse phosphotransfer in the gluconeogenic direction (forming F6P and PPi) with similar efficiency to the glycolytic reaction (Supplementary Table S1). Not only that, but gluconeogenesis is thought to have evolved before glycolysis [95], and some PPi-dependent PFKs functionally replace the transaldolase in another ancient branch of metabolism, the pentose phosphate pathway [62]. Thus, it is conceivable that a primordial cell made use of one multifunctional, PPi-dependent enzyme in three physiological roles.

On balance, it is the non-allosteric, multifunctional, PPi-dependent PFKs that possess an array of ancestor-like characteristics. However, as discussed in the previous section, the number of PPi-dependent PFK structures is currently low. Further exploration of ancestor-like PPi-dependent PFKs will undoubtedly generate new hypotheses on plausible evolutionary routes to ATP dependence, allostery and narrowed substrate range.

Future directions: diversity, dynamics, diseases and drugs

Within the Phosphofructokinase Superfamily, hundreds of functional parameters and dozens of structures are now available in BRENDA and the PDB, respectively. At the same time, there are still important unanswered questions and exciting opportunities for future research.

We have highlighted the taxonomic bias in our structural knowledge of PFKs, which mirrors the bias for all other enzymes [96]. For example, no superfamily members from plants or archaea have had their structures determined. Similarly, the available bacterial structures are from only three of the 175 recognised phyla in the Genome Taxonomy Database (release 220; [97]). Given the structural diversity that has already been described from 14 (taxonomically non-representative) species, it seems certain that new structures will continue to reveal domain fusions, insertions and subunit compositions that underpin complex regulatory behaviour and, ultimately, the adaptation of organisms to their environment.

While structural biology has been essential for elucidating the basis of the R-to-T allosteric transition, a noteworthy recent advance was the use of molecular dynamics to aid with understanding this transition in human PFKP and PFKL [72,98]. Computational approaches for identifying the networks of residues in proteins that transmit allosteric signals are advancing rapidly [99]. Combined with increasingly powerful molecular dynamics simulations [100], an exciting opportunity now exists to understand the allosteric transition at hitherto unseen levels of spatial and temporal resolution. Moreover, with atomistic understanding comes programmability. In the era of synthetic biology, there is significant interest in designing allosterically switchable properties into proteins [101,102]. Much of metabolic engineering involves programming carbon flux. There is considerable potential to build on groundbreaking work [56] by designing exactly the right PFK for the desired outcome in microbial cell factories.

Finally, our increasing knowledge of PFK biochemistry will likely translate into the clinic. As discussed above, TbrPFK has already been validated as a target for promising anti-trypanosome drugs [81]. In humans, Tarui disease (also known as glycogen storage disease type VII) is a rare autosomal recessive disorder associated with the absence of PFK activity in skeletal muscle [103,104]. Structures of the human isoforms are rationalising disease-causing mutations [69]. As an important regulator of glycolytic flux, PFK has also attracted attention because glucose metabolism is dysregulated in a range of conditions including cancer, diabetes and Alzheimer’s disease [98,105]. Determining the structure of human PFKP enabled 44 somatic mutations from cancers to be mapped onto it, leading to the prediction that 28 of these would affect activity, three of which were validated experimentally [70]. One specific example is the PFKP mutation D564N—identified in colon cancer—which was shown to decrease enzyme activity but was proposed to favour cancer cell survival in tumour microenvironments where oxidative stress is present [98]. On the other hand, up-regulation of PFKL expression is directly connected to the Warburg effect, in which cancerous cells increase glycolysis independent of the oxygen level [106]. This has recently been explored in hepatocellular carcinoma (HCC), in which the E3 ubiquitin ligase A20 usually acts as a tumour suppressor by targeting PFKL for degradation. Malignant transformation results in reduced expression of A20, which in turn causes PFKL to accumulate. This facilitates cell proliferation and migration, HCC growth and reduced patient survival [107].

Future therapeutic approaches may therefore lie in designing small molecules that either activate or deactivate PFK, depending on the particular disease. For example, a small molecule, NA-11, inhibited the pro-inflammatory oxidative burst in neutrophils specifically by binding at effector site N2 in PFKL (Figure 4B) and activating the enzyme [71]. The mechanism of action suggested a wholly novel approach to controlling inflammatory disorders such as acute respiratory distress syndrome, and potentially treating other autoimmune diseases. On the other hand, a drug repurposing screen identified an antipsychotic medication, penfluridol, that specifically inhibited PFKL in oesophageal squamous cell carcinoma, suppressing tumour growth and inducing apoptosis [108].

Our biochemical knowledge of the Phosphofructokinase Superfamily has come a long way in the past 90 years. Nevertheless, the diversity of structures and functions across the tree of life and the complex role of PFK at the heart of the metabolic network ensure that much still remains to be discovered.

The authors declare that there are no competing interests associated with this manuscript.

W.M.P. acknowledges funding from New Zealand’s Marsden Fund (grant no. 18-VUW-050). J.A.C. was supported by a doctoral scholarship from Victoria University of Wellington.

J.A.C.: Conceptualisation, Data curation, Formal analysis, Investigation, Visualisation, Writing—original draft; Writing—review and editing. W.M.P.: Conceptualisation, Data curation, Formal analysis, Funding acquisition, Investigation, Visualisation, Writing—original draft; Writing—review and editing.

BRENDA

BRaunschweig ENzyme DAtabase

EC

Enzyme Commission

FBP

fructose 1,6-bisphosphate

FBPase

fructose 1,6-bisphosphatase

F6P

fructose 6-phosphate

PDB

Protein Data Bank

PFK

6-phosphofructokinase

PFKL

liver isoform of mammalian phosphofructokinase

PFKM

muscle isoform of mammalian phosphofructokinase

PFKP

platelet isoform of mammalian phosphofructokinase

SBP

sedoheptulose 1,7-bisphosphate

S7P

sedoheptulose 7-phosphate

1.
Ostern
,
P.
,
Guthke
,
J.A.
and
Terszakowec
,
J
. (
1936
)
Über die Bildung des Hexose-monophosphorsäure-esters und dessen Umwandlung in Fructose-diphosphorsäure-ester im Muskel
.
Biol. Chem.
243
,
9
37
https://doi.org/10.1515/bchm2.1936.243.1-3.9
2.
Young
,
W.J.
and
Harden
,
A
. (
1909
)
The hexosephosphate formed by yeast-juice from hexose and phosphate
.
Proc. R. Soc. Lond. B.
81
,
528
545
https://doi.org/10.1098/rspb.1909.0051
3.
Barańska
,
J.
,
Dzugaj
,
A.
and
Kwiatkowska-Korczak
,
J
. (
2007
) Embden-Meyerhof-Parnas, the first metabolic pathway: the fate of prominent Polish biochemist Jakub Karol Parnas.
In
Comprehensive Biochemistry
(
Semenza
,
G.
, ed),
vol
.
45
,
pp
.
157
207
4.
Blangy
,
D.
,
Buc
,
H.
and
Monod
,
J
. (
1968
)
Kinetics of the allosteric interactions of phosphofructokinase from Escherichia coli
.
J. Mol. Biol.
31
,
13
35
https://doi.org/10.1016/0022-2836(68)90051-x
5.
Evans
,
P.R.
and
Hudson
,
P.J
. (
1979
)
Structure and control of phosphofructokinase from Bacillus stearothermophilus
.
Nature
279
,
500
504
https://doi.org/10.1038/279500a0
6.
Evans
,
P.R.
,
Farrants
,
G.W.
and
Hudson
,
P.J
. (
1981
)
Phosphofructokinase: structure and control
.
Philos. Trans. R. Soc. Lond B: Biol. Sci.
293
,
53
62
https://doi.org/10.1098/rstb.1981.0059
7.
Shirakihara
,
Y.
and
Evans
,
P.R
. (
1988
)
Crystal structure of the complex of phosphofructokinase from Escherichia coli with its reaction products
.
J. Mol. Biol.
204
,
973
994
https://doi.org/10.1016/0022-2836(88)90056-3
8.
Hellinga
,
H.W.
and
Evans
,
P.R
. (
1987
)
Mutations in the active site of Escherichia coli phosphofructokinase
.
Nature
327
,
437
439
https://doi.org/10.1038/327437a0
9.
Berger
,
S.A.
and
Evans
,
P.R
. (
1992
)
Site-directed mutagenesis identifies catalytic residues in the active site of Escherichia coli phosphofructokinase
.
Biochemistry
31
,
9237
9242
https://doi.org/10.1021/bi00153a017
10.
Schirmer
,
T.
and
Evans
,
P.R
. (
1990
)
Structural basis of the allosteric behaviour of phosphofructokinase
.
Nature
343
,
140
145
https://doi.org/10.1038/343140a0
11.
Paysan-Lafosse
,
T.
,
Blum
,
M.
,
Chuguransky
,
S.
,
Grego
,
T.
,
Pinto
,
B.L.
,
Salazar
,
G.A.
et al
. (
2023
)
InterPro in 2022
.
Nucleic Acids Res.
51
,
D418
D427
https://doi.org/10.1093/nar/gkac993
12.
Ronimus
,
R.S.
and
Morgan
,
H.W
. (
2001
)
The biochemical properties and phylogenies of phosphofructokinases from extremophiles
.
Extremophiles
5
,
357
373
https://doi.org/10.1007/s007920100215
13.
Moore
,
S.A.
,
Ronimus
,
R.S.
,
Roberson
,
R.S.
and
Morgan
,
H.W
. (
2002
)
The structure of a pyrophosphate-dependent phosphofructokinase from the Lyme disease spirochete Borrelia burgdorferi
.
Structure
10
,
659
671
https://doi.org/10.1016/s0969-2126(02)00760-8
14.
Murillo-López
,
J.
,
Zinovjev
,
K.
,
Pereira
,
H.
,
Caniuguir
,
A.
,
Garratt
,
R.
,
Babul
,
J.
et al
. (
2019
)
Studying the phosphoryl transfer mechanism of the E. coli phosphofructokinase-2: from X-ray structure to quantum mechanics/molecular mechanics simulations
.
Chem. Sci.
10
,
2882
2892
https://doi.org/10.1039/c9sc00094a
15.
Bork
,
P.
,
Sander
,
C.
and
Valencia
,
A
. (
1993
)
Convergent evolution of similar enzymatic function on different protein folds: the hexokinase, ribokinase, and galactokinase families of sugar kinases
.
Protein Sci.
2
,
31
40
https://doi.org/10.1002/pro.5560020104
16.
Guixé
,
V.
and
Merino
,
F
. (
2009
)
The ADP-dependent sugar kinase family: kinetic and evolutionary aspects
.
IUBMB Life
61
,
753
761
https://doi.org/10.1002/iub.217
17.
Park
,
J.
and
Gupta
,
R.S
. (
2008
)
Adenosine kinase and ribokinase--the RK family of proteins
.
Cell. Mol. Life Sci.
65
,
2875
2896
https://doi.org/10.1007/s00018-008-8123-1
18.
Cabrera
,
R.
,
Baez
,
M.
,
Pereira
,
H.M.
,
Caniuguir
,
A.
,
Garratt
,
R.C.
and
Babul
,
J
. (
2011
)
The crystal complex of phosphofructokinase-2 of Escherichia coli with fructose-6-phosphate
.
J. Biol. Chem.
286
,
5774
5783
https://doi.org/10.1074/jbc.M110.163162
19.
Mertens
,
E.
,
Ladror
,
U.S.
,
Lee
,
J.A.
,
Miretsky
,
A.
,
Morris
,
A.
,
Rozario
,
C.
et al
. (
1998
)
The pyrophosphate-dependent phosphofructokinase of the protist, Trichomonas vaginalis, and the evolutionary relationships of protist phosphofructokinases
.
J. Mol. Evol.
47
,
739
750
https://doi.org/10.1007/pl00006433
20.
Siebers
,
B.
,
Klenk
,
H.P.
and
Hensel
,
R
. (
1998
)
PPᵢ-dependent phosphofructokinase from Thermoproteus tenax, an archaeal descendant of an ancient line in phosphofructokinase evolution
.
J. Bacteriol.
180
,
2137
2143
https://doi.org/10.1128/JB.180.8.2137-2143.1998
21.
Müller
,
M.
,
Lee
,
J.A.
,
Gordon
,
P.
,
Gaasterland
,
T.
and
Sensen
,
C.W
. (
2001
)
Presence of prokaryotic and eukaryotic species in all subgroups of the PPᵢ-dependent group II phosphofructokinase protein family
.
J. Bacteriol.
183
,
6714
6716
https://doi.org/10.1128/JB.183.22.6714-6716.2001
22.
Bapteste
,
E.
,
Moreira
,
D.
and
Philippe
,
H
. (
2003
)
Rampant horizontal gene transfer and phospho-donor change in the evolution of the phosphofructokinase
.
Gene
318
,
185
191
https://doi.org/10.1016/s0378-1119(03)00797-2
23.
Reshetnikov
,
A.S.
,
Rozova
,
O.N.
,
Khmelenina
,
V.N.
,
Mustakhimov
,
I.I.
,
Beschastny
,
A.P.
,
Murrell
,
J.C.
et al
. (
2008
)
Characterization of the pyrophosphate-dependent 6-phosphofructokinase from Methylococcus capsulatus Bath
.
FEMS Microbiol. Lett.
288
,
202
210
https://doi.org/10.1111/j.1574-6968.2008.01366.x
24.
Rozova
,
O.N.
,
Khmelenina
,
V.N.
and
Trotsenko
,
Y.A
. (
2012
)
Characterization of recombinant PPᵢ-dependent 6-phosphofructokinases from Methylosinus trichosporium OB3b and Methylobacterium nodulans ORS 2060
.
Biochem. Mosc.
77
,
288
295
https://doi.org/10.1134/S0006297912030078
25.
Rada
,
P.
,
Makki
,
A.R.
,
Zimorski
,
V.
,
Garg
,
S.
,
Hampl
,
V.
,
Hrdý
,
I.
et al
. (
2015
)
N-terminal presequence-independent import of phosphofructokinase into hydrogenosomes of Trichomonas vaginalis
.
Eukaryotic Cell
14
,
1264
1275
https://doi.org/10.1128/EC.00104-15
26.
Reeves
,
R.E.
,
South
,
D.J.
,
Blytt
,
H.J.
and
Warren
,
L.G
. (
1974
)
Pyrophosphate: ᴅ-fructose 6-phosphate 1-phosphotransferase
.
J. Biol. Chem.
249
,
7737
7741
https://doi.org/10.1016/S0021-9258(19)42029-2
27.
Chang
,
A.
,
Jeske
,
L.
,
Ulbrich
,
S.
,
Hofmann
,
J.
,
Koblitz
,
J.
,
Schomburg
,
I.
et al
. (
2021
)
BRENDA, the ELIXIR core data resource in 2021: new developments and updates
.
Nucleic Acids Res.
49
,
D498
D508
https://doi.org/10.1093/nar/gkaa1025
28.
Poorman
,
R.A.
,
Randolph
,
A.
,
Kemp
,
R.G.
and
Heinrikson
,
R.L
. (
1984
)
Evolution of phosphofructokinase—gene duplication and creation of new effector sites
.
Nature
309
,
467
469
https://doi.org/10.1038/309467a0
29.
Heinisch
,
J.
,
Ritzel
,
R.G.
,
Borstel
,
R.C.
,
Aguilera
,
A.
,
Rodicio
,
R.
and
Zimmermann
,
F.K
. (
1989
)
The phosphofructokinase genes of yeast evolved from two duplication events
.
Gene
78
,
309
321
https://doi.org/10.1016/0378-1119(89)90233-3
30.
Schöneberg
,
T.
,
Kloos
,
M.
,
Brüser
,
A.
,
Kirchberger
,
J.
and
Sträter
,
N
. (
2013
)
Structure and allosteric regulation of eukaryotic 6-phosphofructokinases
.
Biol. Chem.
394
,
977
993
https://doi.org/10.1515/hsz-2013-0130
31.
Dunaway
,
G.A.
,
Kasten
,
T.P.
,
Sebo
,
T.
and
Trapp
,
R
. (
1988
)
Analysis of the phosphofructokinase subunits and isoenzymes in human tissues
.
Biochem. J.
251
,
677
683
https://doi.org/10.1042/bj2510677
32.
Kim
,
T.L.
,
Lim
,
H.
,
Denison
,
M.I.J.
,
Natarajan
,
S.
and
Oh
,
C
. (
2023
)
Genome-wide identification of the PFK gene family and their expression analysis in Quercus rubra
.
Front. Genet.
14
,
1289557
https://doi.org/10.3389/fgene.2023.1289557
33.
Sträter
,
N.
,
Marek
,
S.
,
Kuettner
,
E.B.
,
Kloos
,
M.
,
Keim
,
A.
,
Brüser
,
A.
et al
. (
2011
)
Molecular architecture and structural basis of allosteric regulation of eukaryotic phosphofructokinases
.
FASEB J.
25
,
89
98
https://doi.org/10.1096/fj.10-163865
34.
McNae
,
I.W.
,
Martinez-Oyanedel
,
J.
,
Keillor
,
J.W.
,
Michels
,
P.A.M.
,
Fothergill-Gilmore
,
L.A.
and
Walkinshaw
,
M.D
. (
2009
)
The crystal structure of ATP-bound phosphofructokinase from Trypanosoma brucei reveals conformational transitions different from those of other phosphofructokinases
.
J. Mol. Biol.
385
,
1519
1533
https://doi.org/10.1016/j.jmb.2008.11.047
35.
Kamemoto
,
E.S.
and
Mansour
,
T.E
. (
1986
)
Phosphofructokinase in the liver fluke Fasciola hepatica. Purification and kinetic changes by phosphorylation
.
J. Biol. Chem.
261
,
4346
4351
https://doi.org/10.1016/S0021-9258(17)35667-3
36.
Wang
,
X.
and
Kemp
,
R.G
. (
2001
)
Reaction path of phosphofructo-1-kinase is altered by mutagenesis and alternative substrates
.
Biochemistry
40
,
3938
3942
https://doi.org/10.1021/bi002709o
37.
Ferreras
,
C.
,
Hernández
,
E.D.
,
Martínez-Costa
,
O.H.
and
Aragón
,
J.J
. (
2009
)
Subunit interactions and composition of the fructose 6-phosphate catalytic site and the fructose 2,6-bisphosphate allosteric site of mammalian phosphofructokinase
.
J. Biol. Chem.
284
,
9124
9131
https://doi.org/10.1074/jbc.M807737200
38.
Knowles
,
V.L.
,
Greyson
,
M.F.
and
Dennis
,
D.T
. (
1990
)
Characterization of ATP-dependent fructose 6-phosphate 1-phosphotransferase isozymes from leaf and endosperm tissues of Ricinus communis
.
Plant Physiol.
92
,
155
159
https://doi.org/10.1104/pp.92.1.155
39.
Cawood
,
M.E.
,
Botha
,
F.C.
and
Small
,
J.G.C
. (
1988
)
Properties of the phosphofructokinase isoenzymes from germinating cucumber seeds
.
J. Plant Physiol.
132
,
204
209
https://doi.org/10.1016/S0176-1617(88)80162-7
40.
Le
,
S.B.
,
Heggeset
,
T.M.B.
,
Haugen
,
T.
,
Nærdal
,
I.
and
Brautaset
,
T
. (
2017
)
6-phosphofructokinase and ribulose-5-phosphate 3-epimerase in methylotrophic Bacillus methanolicus ribulose monophosphate cycle
.
Appl. Microbiol. Biotechnol.
101
,
4185
4200
https://doi.org/10.1007/s00253-017-8173-0
41.
Deng
,
Z.
,
Wang
,
X.
and
Kemp
,
R.G
. (
2000
)
Site-directed mutagenesis of the fructose 6-phosphate binding site of the pyrophosphate-dependent phosphofructokinase of Entamoeba histolytica
.
Arch. Biochem. Biophys.
380
,
56
62
https://doi.org/10.1006/abbi.2000.1881
42.
Mertens
,
E.
,
De Jonckheere
,
J.
and
Van Schaftingen
,
E
. (
1993
)
Pyrophosphate-dependent phosphofructokinase from the amoeba Naegleria fowleri, an AMP-sensitive enzyme
.
Biochem. J.
292
,
797
803
https://doi.org/10.1042/bj2920797
43.
Enomoto
,
T.
,
Ohyama
,
H.
and
Kodama
,
M
. (
1992
)
Purification and characterization of pyrophosphate: d-fructose 6-phosphate 1-phosphotransferase from rice seedlings
.
Biosci. Biotechnol. Biochem.
56
,
251
255
https://doi.org/10.1271/bbb.56.251
44.
Bertagnolli
,
B.L.
and
Cook
,
P.F
. (
1984
)
Kinetic mechanism of pyrophosphate-dependent phosphofructokinase from Propionibacterium freudenreichii
.
Biochemistry
23
,
4101
4108
https://doi.org/10.1021/bi00313a014
45.
Kowalczyk
,
S.
,
Januszewska
,
B.
,
Cymerska
,
E.
and
Maslowski
,
P
. (
1984
)
The occurrence of inorganic pyrophosphate: d‐fructose‐6‐phosphate 1‐phosphotransferase in higher plants. I. Initial characterization of partially purified enzyme from Sanseviera trifasciata leaves
.
Physiol. Plant.
60
,
31
37
https://doi.org/10.1111/j.1399-3054.1984.tb04245.x
46.
Davidi
,
D.
,
Longo
,
L.M.
,
Jabłońska
,
J.
,
Milo
,
R.
and
Tawfik
,
D.S
. (
2018
)
A bird’s-eye view of enzyme evolution: chemical, physicochemical, and physiological considerations
.
Chem. Rev.
118
,
8786
8797
https://doi.org/10.1021/acs.chemrev.8b00039
47.
Kuil
,
T.
,
Nurminen
,
C.M.K.
and
Maris
,
A.J.A
. (
2023
)
Pyrophosphate as allosteric regulator of ATP-phosphofructokinase in Clostridium thermocellum and other bacteria with ATP- and PPi-phosphofructokinases
.
Arch. Biochem. Biophys.
743
,
109676
https://doi.org/10.1016/j.abb.2023.109676
48.
Paricharttanakul
,
N.M.
,
Ye
,
S.
,
Menefee
,
A.L.
,
Javid-Majd
,
F.
,
Sacchettini
,
J.C.
and
Reinhart
,
G.D
. (
2005
)
Kinetic and structural characterization of phosphofructokinase from Lactobacillus bulgaricus
.
Biochemistry
44
,
15280
15286
https://doi.org/10.1021/bi051283g
49.
Sola-Penna
,
M.
,
Da Silva
,
D.
,
Coelho
,
W.S.
,
Marinho-Carvalho
,
M.M.
and
Zancan
,
P
. (
2010
)
Regulation of mammalian muscle type 6-phosphofructo-1-kinase and its implication for the control of the metabolism
.
IUBMB Life
62
,
791
796
https://doi.org/10.1002/iub.393
50.
Hicks
,
K.G.
,
Cluntun
,
A.A.
,
Schubert
,
H.L.
,
Hackett
,
S.R.
,
Berg
,
J.A.
,
Leonard
,
P.G.
et al
. (
2023
)
Protein-metabolite interactomics of carbohydrate metabolism reveal regulation of lactate dehydrogenase
.
Science
379
,
996
1003
https://doi.org/10.1126/science.abm3452
51.
Podestá
,
F.E.
and
Plaxton
,
W.C
. (
2003
)
Fluorescence study of ligand binding to potato tuber pyrophosphate-dependent phosphofructokinase: evidence for competitive binding between fructose-1,6-bisphosphate and fructose-2,6-bisphosphate
.
Arch. Biochem. Biophys.
414
,
101
107
https://doi.org/10.1016/s0003-9861(03)00157-7
52.
Turner
,
W.L.
and
Plaxton
,
W.C
. (
2003
)
Purification and characterization of pyrophosphate- and ATP-dependent phosphofructokinases from banana fruit
.
Planta
217
,
113
121
https://doi.org/10.1007/s00425-002-0962-7
53.
Basson
,
C.E.
,
Groenewald
,
J.H.
,
Kossmann
,
J.
,
Cronjé
,
C.
and
Bauer
,
R
. (
2011
)
Upregulation of pyrophosphate: fructose 6-phosphate 1-phosphotransferase (PFP) activity in strawberry
.
Transgenic Res.
20
,
925
931
https://doi.org/10.1007/s11248-010-9451-0
54.
Fernandes
,
P.M.
,
Kinkead
,
J.
,
McNae
,
I.W.
,
Bringaud
,
F.
,
Michels
,
P.A.M.
and
Walkinshaw
,
M.D
. (
2019
)
The kinetic characteristics of human and trypanosomatid phosphofructokinases for the reverse reaction
.
Biochem. J.
476
,
179
191
https://doi.org/10.1042/BCJ20180635
55.
Noor
,
E.
,
Bar-Even
,
A.
,
Flamholz
,
A.
,
Reznik
,
E.
,
Liebermeister
,
W.
and
Milo
,
R
. (
2014
)
Pathway thermodynamics highlights kinetic obstacles in central metabolism
.
PLOS Comput. Biol.
10
, e1003483 https://doi.org/10.1371/journal.pcbi.1003483
56.
Hon
,
S.
,
Jacobson
,
T.
,
Stevenson
,
D.M.
,
Maloney
,
M.I.
,
Giannone
,
R.J.
,
Hettich
,
R.L.
et al
. (
2022
)
Increasing the thermodynamic driving force of the phosphofructokinase reaction in Clostridium thermocellum
.
Appl. Environ. Microbiol.
88
, e01258 https://doi.org/10.1128/aem.01258-22
57.
Phillips
,
N.F.
and
Li
,
Z
. (
1995
)
Kinetic mechanism of pyrophosphate-dependent phosphofructokinase from Giardia lamblia
.
Mol. Biochem. Parasitol.
73
,
43
51
https://doi.org/10.1016/0166-6851(95)00087-h
58.
Mertens
,
E
. (
1991
)
Pyrophosphate-dependent phosphofructokinase, an anaerobic glycolytic enzyme?
FEBS Lett.
285
,
1
5
https://doi.org/10.1016/0014-5793(91)80711-b
59.
Alves
,
A.M.
,
Euverink
,
G.J.
,
Hektor
,
H.J.
,
Hessels
,
G.I.
,
Vlag
,
J.
,
Vrijbloed
,
J.W
, et al
. (
1994
)
Enzymes of glucose and methanol metabolism in the actinomycete Amycolatopsis methanolica
.
J. Bacteriol.
176
,
6827
6835
https://doi.org/10.1128/jb.176.22.6827-6835.1994
60.
Arimoto
,
T.
,
Ansai
,
T.
,
Yu
,
W.
,
Turner
,
A.J.
and
Takehara
,
T
. (
2002
)
Kinetic analysis of PPᵢ-dependent phosphofructokinase from Porphyromonas gingivalis
.
FEMS Microbiol. Lett.
207
,
35
38
https://doi.org/10.1111/j.1574-6968.2002.tb11024.x
61.
Susskind
,
B.M.
,
Warren
,
L.G.
and
Reeves
,
R.E
. (
1982
)
A pathway for the interconversion of hexose and pentose in the parasitic amoeba Entamoeba histolytica
.
Biochem. J.
204
,
191
196
https://doi.org/10.1042/bj2040191
62.
Koendjbiharie
,
J.G.
,
Hon
,
S.
,
Pabst
,
M.
,
Hooftman
,
R.
,
Stevenson
,
D.M.
,
Cui
,
J
, et al
. (
2020
)
The pentose phosphate pathway of cellulolytic clostridia relies on 6-phosphofructokinase instead of transaldolase
.
J. Biol. Chem.
295
,
1867
1878
https://doi.org/10.1074/jbc.RA119.011239
63.
Koendjbiharie
,
J.G.
,
Kuil
,
T.
,
Nurminen
,
C.M.K.
and
Maris
,
A.J.A
. (
2024
)
The 6-phosphofructokinase reaction in Acetivibrio thermocellus is both ATP- and pyrophosphate-dependent
.
Metab. Eng.
86
,
41
54
https://doi.org/10.1016/j.ymben.2024.09.002
64.
Yang
,
X.
,
Yin
,
X.
,
Liu
,
J.
,
Niu
,
Z.
,
Yang
,
J.
and
Shen
,
B
. (
2022
)
Essential role of pyrophosphate homeostasis mediated by the pyrophosphate-dependent phosphofructokinase in Toxoplasma gondii
.
PLOS Pathog.
18
, e1010293 https://doi.org/10.1371/journal.ppat.1010293
65.
Berman
,
H.
,
Henrick
,
K.
and
Nakamura
,
H
. (
2003
)
Announcing the worldwide Protein Data Bank
.
Nat. Struct. Mol. Biol.
10
,
980
980
https://doi.org/10.1038/nsb1203-980
66.
Mosser
,
R.
,
Reddy
,
M.C.M.
,
Bruning
,
J.B.
,
Sacchettini
,
J.C.
and
Reinhart
,
G.D
. (
2013
)
Redefining the role of the quaternary shift in Bacillus stearothermophilus phosphofructokinase
.
Biochemistry
52
,
5421
5429
https://doi.org/10.1021/bi4002503
67.
Banaszak
,
K.
,
Mechin
,
I.
,
Obmolova
,
G.
,
Oldham
,
M.
,
Chang
,
S.H.
,
Ruiz
,
T.
et al
. (
2011
)
The crystal structures of eukaryotic phosphofructokinases from baker’s yeast and rabbit skeletal muscle
.
J. Mol. Biol.
407
,
284
297
https://doi.org/10.1016/j.jmb.2011.01.019
68.
Kloos
,
M.
,
Brüser
,
A.
,
Kirchberger
,
J.
,
Schöneberg
,
T.
and
Sträter
,
N
. (
2014
)
Crystallization and preliminary crystallographic analysis of human muscle phosphofructokinase, the main regulator of glycolysis
.
Acta Crystallogr. F. Struct. Biol. Commun.
70
,
578
582
https://doi.org/10.1107/S2053230X14008723
69.
Kloos
,
M.
,
Brüser
,
A.
,
Kirchberger
,
J.
,
Schöneberg
,
T.
and
Sträter
,
N
. (
2015
)
Crystal structure of human platelet phosphofructokinase-1 locked in an activated conformation
.
Biochem. J.
469
,
421
432
https://doi.org/10.1042/BJ20150251
70.
Webb
,
B.A.
,
Forouhar
,
F.
,
Szu
,
F.E.
,
Seetharaman
,
J.
,
Tong
,
L.
and
Barber
,
D.L
. (
2015
)
Structures of human phosphofructokinase-1 and atomic basis of cancer-associated mutations
.
Nature
523
,
111
114
https://doi.org/10.1038/nature14405
71.
Amara
,
N.
,
Cooper
,
M.P.
,
Voronkova
,
M.A.
,
Webb
,
B.A.
,
Lynch
,
E.M.
,
Kollman
,
J.M
, et al
. (
2021
)
Selective activation of PFKL suppresses the phagocytic oxidative burst
.
Cell
184
,
4480
4494
https://doi.org/10.1016/j.cell.2021.07.004
72.
Lynch
,
E.M.
,
Hansen
,
H.
,
Salay
,
L.
,
Cooper
,
M.
,
Timr
,
S.
,
Kollman
,
J.M.
et al
. (
2024
)
Structural basis for allosteric regulation of human phosphofructokinase-1
.
Nat. Commun.
15
, 7323 https://doi.org/10.1038/s41467-024-51808-6
73.
Martínez-Costa
,
O.H.
,
Sánchez
,
V.
,
Lázaro
,
A.
,
Hernández
,
E.D.
,
Tornheim
,
K.
and
Aragón
,
J.J
. (
2012
)
Distinct functional roles of the two terminal halves of eukaryotic phosphofructokinase
.
Biochem. J.
445
,
213
218
https://doi.org/10.1042/BJ20120173
74.
Rizzo
,
S.C.
and
Eckel
,
R.E
. (
1966
)
Control of glycolysis in human erythrocytes by inorganic phosphate and sulfate
.
Am. J. Physiol.
211
,
429
436
https://doi.org/10.1152/ajplegacy.1966.211.2.429
75.
Yugi
,
K.
,
Kubota
,
H.
,
Toyoshima
,
Y.
,
Noguchi
,
R.
,
Kawata
,
K.
,
Komori
,
Y
, et al
. (
2014
)
Reconstruction of insulin signal flow from phosphoproteome and metabolome data
.
Cell Rep.
8
,
1171
1183
https://doi.org/10.1016/j.celrep.2014.07.021
76.
Cronin
,
C.N.
and
Tipton
,
K.F
. (
1985
)
Purification and regulatory properties of phosphofructokinase from Trypanosoma (Trypanozoon) brucei brucei
.
Biochem. J.
227
,
113
124
https://doi.org/10.1042/bj2270113
77.
Martinez-Oyanedel
,
J.
,
McNae
,
I.W.
,
Nowicki
,
M.W.
,
Keillor
,
J.W.
,
Michels
,
P.A.M.
,
Fothergill-Gilmore
,
L.A.
et al
. (
2007
)
The first crystal structure of phosphofructokinase from a eukaryote: Trypanosoma brucei
.
J. Mol. Biol.
366
,
1185
1198
https://doi.org/10.1016/j.jmb.2006.10.019
78.
Michels
,
P.A.M.
,
Chevalier
,
N.
,
Opperdoes
,
F.R.
,
Rider
,
M.H.
and
Rigden
,
D.J
. (
1997
)
The glycosomal ATP‐dependent phosphofructokinase of Trypanosoma Brucei must have evolved from an ancestral pyrophosphate‐dependent enzyme
.
Eur. J. Biochem.
250
,
698
704
https://doi.org/10.1111/j.1432-1033.1997.00698.x
79.
Fernandes
,
P.M.
,
Kinkead
,
J.
,
McNae
,
I.W.
,
Vásquez‐Valdivieso
,
M.
,
Wear
,
M.A.
,
Michels
,
P.A.M.
et al
. (
2020
)
Kinetic and structural studies of Trypanosoma and Leishmania phosphofructokinases show evolutionary divergence and identify AMP as a switch regulating glycolysis versus gluconeogenesis
.
FEBS J.
287
,
2847
2861
https://doi.org/10.1111/febs.15177
80.
Verlinde
,
C.L.
,
Hannaert
,
V.
,
Blonski
,
C.
,
Willson
,
M.
,
Périé
,
J.J.
,
Fothergill-Gilmore
,
L.A.
et al
. (
2001
)
Glycolysis as a target for the design of new anti-trypanosome drugs
.
Drug Resist. Updat.
4
,
50
65
https://doi.org/10.1054/drup.2000.0177
81.
McNae
,
I.W.
,
Kinkead
,
J.
,
Malik
,
D.
,
Yen
,
L.H.
,
Walker
,
M.K.
,
Swain
,
C.
et al
. (
2021
)
Fast acting allosteric phosphofructokinase inhibitors block trypanosome glycolysis and cure acute African trypanosomiasis in mice
.
Nat. Commun.
12
,
1052
https://doi.org/10.1038/s41467-021-21273-6
82.
Chi
,
A.
and
Kemp
,
R.G
. (
2000
)
The primordial high energy compound: ATP or inorganic pyrophosphate?
J. Biol. Chem.
275
,
35677
35679
https://doi.org/10.1074/jbc.C000581200
83.
Fothergill-Gilmore
,
L.A.
and
Michels
,
P.A
. (
1993
)
Evolution of glycolysis
.
Prog. Biophys. Mol. Biol.
59
,
105
235
https://doi.org/10.1016/0079-6107(93)90001-z
84.
Cross
,
P.J.
,
Allison
,
T.M.
,
Dobson
,
R.C.J.
,
Jameson
,
G.B.
and
Parker
,
E.J
. (
2013
)
Engineering allosteric control to an unregulated enzyme by transfer of a regulatory domain
.
Proc. Natl. Acad. Sci. U.S.A.
110
,
2111
2116
https://doi.org/10.1073/pnas.1217923110
85.
Alves
,
A.M.
,
Meijer
,
W.G.
,
Vrijbloed
,
J.W.
and
Dijkhuizen
,
L
. (
1996
)
Characterization and phylogeny of the PFP gene of Amycolatopsis methanolica encoding PPᵢ-dependent phosphofructokinase
.
J. Bacteriol.
178
,
149
155
https://doi.org/10.1128/jb.178.1.149-155.1996
86.
Ding
,
Y.H.
,
Ronimus
,
R.S.
and
Morgan
,
H.W
. (
2000
)
Sequencing, cloning, and high-level expression of the PFP gene, encoding a PPᵢ-dependent phosphofructokinase from the extremely thermophilic eubacterium Dictyoglomus thermophilum
.
J. Bacteriol.
182
,
4661
4666
https://doi.org/10.1128/JB.182.16.4661-4666.2000
87.
Craig
,
J.M.
,
Kumar
,
S.
and
Hedges
,
S.B
. (
2023
)
The origin of eukaryotes and rise in complexity were synchronous with the rise in oxygen
.
Front. Bioinform.
3
, 1233281 https://doi.org/10.3389/fbinf.2023.1233281
88.
Brewster
,
J.L.
,
Finn
,
T.J.
,
Ramirez
,
M.A.
and
Patrick
,
W.M
. (
2016
)
Whither life? Conjectures on the future evolution of biochemistry
.
Biol. Lett.
12
, 20160269 https://doi.org/10.1098/rsbl.2016.0269
89.
Kalapos
,
M.P.
and
Bari
,
L
. (
2022
)
Hidden biochemical fossils reveal an evolutionary trajectory for glycolysis in the prebiotic era
.
FEBS Lett.
596
,
1955
1968
https://doi.org/10.1002/1873-3468.14408
90.
Lyons
,
T.W.
,
Reinhard
,
C.T.
and
Planavsky
,
N.J
. (
2014
)
The rise of oxygen in Earth’s early ocean and atmosphere
.
Nature
506
,
307
315
https://doi.org/10.1038/nature13068
91.
Kornberg
,
A
. (
1995
)
Inorganic polyphosphate: toward making a forgotten polymer unforgettable
.
J. Bacteriol.
177
,
491
496
https://doi.org/10.1128/jb.177.3.491-496.1995
92.
Copley
,
S.D
. (
2020
)
Evolution of new enzymes by gene duplication and divergence
.
FEBS J.
287
,
1262
1283
https://doi.org/10.1111/febs.15299
93.
Yčas
,
M
. (
1974
)
On earlier states of the biochemical system
.
J. Theor. Biol.
44
,
145
160
https://doi.org/10.1016/s0022-5193(74)80035-4
94.
Jensen
,
R.A
. (
1976
)
Enzyme recruitment in evolution of new function
.
Annu. Rev. Microbiol.
30
,
409
425
https://doi.org/10.1146/annurev.mi.30.100176.002205
95.
Ronimus
,
R.S.
and
Morgan
,
H.W
. (
2003
)
Distribution and phylogenies of enzymes of the Embden-Meyerhof-Parnas pathway from archaea and hyperthermophilic bacteria support a gluconeogenic origin of metabolism
.
Archaea
1
,
199
221
https://doi.org/10.1155/2003/162593
96.
Vickers
,
C.J.
,
Fraga
,
D.
and
Patrick
,
W.M
. (
2021
)
Quantifying the taxonomic bias in enzymology
.
Protein Sci.
30
,
914
921
https://doi.org/10.1002/pro.4041
97.
Parks
,
D.H.
,
Chuvochina
,
M.
,
Rinke
,
C.
,
Mussig
,
A.J.
,
Chaumeil
,
P.A.
and
Hugenholtz
,
P
. (
2022
)
GTDB: an ongoing census of bacterial and archaeal diversity through a phylogenetically consistent, rank normalized and complete genome-based taxonomy
.
Nucleic Acids Res.
50
,
D785
D794
https://doi.org/10.1093/nar/gkab776
98.
Voronkova
,
M.A.
,
Hansen
,
H.L.
,
Cooper
,
M.P.
,
Miller
,
J.
,
Sukumar
,
N.
,
Geldenhuys
,
W.J.
et al
. (
2023
)
Cancer-associated somatic mutations in human phosphofructokinase-1 reveal a critical electrostatic interaction for allosteric regulation of enzyme activity
.
Biochem. J.
480
,
1411
1427
https://doi.org/10.1042/BCJ20230207
99.
Yehorova
,
D.
,
Di Geronimo
,
B.
,
Robinson
,
M.
,
Kasson
,
P.M.
and
Kamerlin
,
S.C.L
. (
2024
)
Using residue interaction networks to understand protein function and evolution and to engineer new proteins
.
Curr. Opin. Struct. Biol.
89
,
102922
https://doi.org/10.1016/j.sbi.2024.102922
100.
Bernetti
,
M.
,
Bosio
,
S.
,
Bresciani
,
V.
,
Falchi
,
F.
and
Masetti
,
M
. (
2024
)
Probing allosteric communication with combined molecular dynamics simulations and network analysis
.
Curr. Opin. Struct. Biol.
86
,
102820
https://doi.org/10.1016/j.sbi.2024.102820
101.
Pillai
,
A.
,
Idris
,
A.
,
Philomin
,
A.
,
Weidle
,
C.
,
Skotheim
,
R.
,
Leung
,
P.J.Y.
et al
. (
2024
)
De novo design of allosterically switchable protein assemblies
.
Nature
632
,
911
920
https://doi.org/10.1038/s41586-024-07813-2
102.
Plaper
,
T.
,
Merljak
,
E.
,
Fink
,
T.
,
Satler
,
T.
,
Ljubetič
,
A.
,
Lainšček
,
D.
et al
. (
2024
)
Designed allosteric protein logic
.
Cell Discov.
10
,
8
https://doi.org/10.1038/s41421-023-00635-y
103.
Hannah
,
W.B.
,
Derks
,
T.G.J.
,
Drumm
,
M.L.
,
Grünert
,
S.C.
,
Kishnani
,
P.S.
and
Vissing
,
J
. (
2023
)
Glycogen storage diseases
.
Nat. Rev. Dis. Primers
9
,
46
https://doi.org/10.1038/s41572-023-00456-z
104.
Musumeci
,
O.
,
Bruno
,
C.
,
Mongini
,
T.
,
Rodolico
,
C.
,
Aguennouz
,
M.
,
Barca
,
E.
et al
. (
2012
)
Clinical features and new molecular findings in muscle phosphofructokinase deficiency (GSD type VII)
.
Neuromuscul. Disord.
22
,
325
330
https://doi.org/10.1016/j.nmd.2011.10.022
105.
Campos
,
M.
and
Albrecht
,
L.V
. (
2023
)
Hitting the sweet spot: how glucose metabolism is orchestrated in space and time by phosphofructokinase-1
.
Cancers
16
,
16
https://doi.org/10.3390/cancers16010016
106.
Li
,
L.
,
Li
,
L.
,
Li
,
W.
,
Chen
,
T.
,
Zou
,
B.
,
Zhao
,
L
, et al
. (
2018
)
TAp73-induced phosphofructokinase-1 transcription promotes the Warburg effect and enhances cell proliferation
.
Nat. Commun.
9
,
4683
https://doi.org/10.1038/s41467-018-07127-8
107.
Feng
,
Y.
,
Zhang
,
Y.
,
Cai
,
Y.
,
Liu
,
R.
,
Lu
,
M.
,
Li
,
T.
et al
. (
2020
)
A20 targets PFKL and glycolysis to inhibit the progression of hepatocellular carcinoma
.
Cell Death Dis.
11
,
89
https://doi.org/10.1038/s41419-020-2278-6
108.
Zheng
,
C.
,
Yu
,
X.
,
Liang
,
Y.
,
Zhu
,
Y.
,
He
,
Y.
,
Liao
,
L.
et al
. (
2022
)
Targeting PFKL with penfluridol inhibits glycolysis and suppresses esophageal cancer tumorigenesis in an AMPK/FOXO3a/BIM-dependent manner
.
Acta Pharm. Sin. B
12
,
1271
1287
https://doi.org/10.1016/j.apsb.2021.09.007
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).

Supplementary data