Coronaviruses have been responsible for multiple challenging global pandemics, including coronavirus disease 2019 (COVID-19), which is caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Papain-like protease (PLpro), one of two cysteine proteases responsible for the maturation and infectivity of SARS-CoV-2, processes and liberates functional proteins from the viral polyproteins and cleaves ubiquitin and ISG15 modifications to inhibit innate immune sensing. Consequently, PLpro is an attractive target for developing COVID-19 therapies. PLpro contains a zinc-finger domain important for substrate binding and structural stability. However, the impact of metal ions on the activity and biophysical properties of SARS-CoV-2 PLpro has not been comprehensively studied. Here, we assessed the impacts of metal ions on the catalytic activity of PLpro. Zinc had the largest inhibitory effect on PLpro, followed by manganese. Calcium, magnesium, and iron had smaller or no effects on PLpro activity. EDTA at a concentration of 0.5 mM was essential for PLpro activity, likely by chelating trace metals that inhibit PLpro. IC50 values for ZnCl2, ZnSO4, and MnCl2 of 0.42 ± 0.02 mM, 0.35 ± 0.01 mM, and 2.6 ± 0.3 mM were obtained in the presence of 0.5 mM EDTA; in the absence of EDTA, the estimated IC50 of ZnCl2 was 14 µM. Tryptophan intrinsic fluorescence analysis confirmed the binding of zinc and manganese to PLpro, and differential scanning calorimetry revealed that zinc but not manganese reduced ΔHcal of PLpro. The results of this study provide a reference for further work targeting PLpro to prevent and treat COVID-19.

Since its discovery in late 2019, coronavirus disease 19 (COVID-19) has infected over a half billion people and caused more than six million deaths [1]. Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is the novel coronavirus responsible for COVID-19 [2,3]. Human coronaviruses cause illnesses ranging from the common cold to highly pathogenic respiratory diseases, including respiratory disease outbreaks in 2002 and 2012 linked to SARS-CoV and Middle East respiratory syndrome coronavirus (MERS-CoV), respectively [4–6]. Due to the high transmission efficiency of SARS-CoV-2, the COVID-19 outbreak quickly evolved from an epidemic to a pandemic in early 2020 and continues to affect lives worldwide [7].

SARS-CoV-2 is an enveloped virus belonging to the genus β-coronavirus and has a positive-sense, single-stranded RNA genome of ∼30 kb [8]. The genome of SARS-CoV-2 has high similarity with those of other β-coronaviruses and shares ∼80% of its genome with SARS-CoV and 50% with MERS-CoV. After the coronavirus spike (S) protein specifically binds to the outside of the host cell, the viral genomic RNA enters the cell and is translated via two open reading frames (ORFs), ORF1a and ORF1b, to produce polyproteins containing non-structural proteins (nsps) [6,8]. The polyproteins translated from ORF1a and ORF1b, pp1a and pp1ab, are cleaved by the cysteine proteases 3-chymotrypsin-like cysteine protease (3CLpro) and papain-like protease (PLpro), which are encoded by nsp5 and nsp3, respectively. PLpro cleaves sites between nsp1 and 2, nsp2 and 3, and nsp3 and 4, while 3CLpro processes the other 11 sites between the remaining nsps, resulting in a total of 16 nsps [6,8–10]. These nsps then assemble to form the viral replication and transcription complex on the host cell membrane.

In addition to its role in viral maturation, PLpro inhibits the host immune response, further increasing its attractiveness as a drug target [6,10]. PLpros from various coronaviruses have been shown to antagonize the type I interferon (IFN) pathway by inactivating different components of the pathway [11–13]. For example, PLpro inhibits the phosphorylation and nuclear translocation of IRF-3 to disrupt the activation of type I IFN responses [12,13]. In addition, PLpro is a deubiquitinating enzyme that cleaves the LXGG sequence found in the C-terminal tail of ubiquitin and the ubiquitin-like modification ISG15 [13,14]. PLpro cleaves both types of modifications to inhibit innate immune sensing [9]. PLpro's ability to remove ubiquitin and ISG15 modifications impedes the host's ability to tag and degrade foreign proteins, including viral proteins [14,15]. Targeting conserved coronavirus proteins like SARS-CoV-2 PLpro, which shares structural and functional similarities with the PLpros of other human coronaviruses such as SARS-CoV and MERS-CoV, has the potential to treat multiple coronavirus strains [16,17]. To design effective treatment strategies targeting PLpro, an investigation of the biochemical and biophysical properties of PLpro is required to identify key similarities and differences between coronaviruses.

Like other human deubiquitinating enzymes belonging to the ubiquitin-specific protease family, including SARS-CoV PLpro, the tertiary structure of SARS-CoV-2 PLpro resembles a hand with four domains: the fingers, palm, thumb, and Ub1 binding domains [9,18]. The catalytic site is located between the palm and thumb domains and comprises the conserved catalytic residues Cys111, His272, and Asp286 [18,19]. The Ub1 binding domain binds the C-terminus of ubiquitin or ISG15, which contains an LXGG cleavage site that enters the catalytic site [19,20]. The catalytic site also cleaves viral proteins containing the LXGG sequence to facilitate the maturation of the viral replicase complex. The fingers domain cradles the ubiquitin in the palm domain once it binds to PLpro and is necessary for structural stability [19]. The fingers domain contains a labile zinc finger ribbon motif in which a Zn2+ ion is tetrahedrally complexed with four cysteines [21–23]. Substituting these zinc-binding cysteines with alanine results in loss of function, emphasizing the importance of this domain to the function and structure of PLpro [24].

Zinc finger domains are a common feature of antiviral proteins, and FDA-approved zinc-ejecting drugs that target this domain are currently on the market [21]. These zinc-ejectors disrupt the zinc-cysteine tetrahedral complex by binding to the cysteine, thus releasing the zinc. However, a molecular dynamics study of PLpro revealed that the zinc finger domain is highly flexible, suggesting that this region may be a poor drug target because it is too dynamic for small-molecule drugs to bind tightly [25]. Nonetheless, the zinc-ejector drugs disulfiram and ebselen destabilize PLpro, thereby reducing its catalytic activity, and could be used in tandem with other antiviral drugs as part of a multi-target approach to treat COVID-19 [21,26,27].

Despite the essential role of zinc in the structure of PLpro, studies of the effects of metal ions on the biochemical properties of PLpro have been limited to analyses of kinetics and thermodynamic stability, and none of these studies has examined the impact of metal chelators such as EDTA on catalytic function [5,28,29]. Here, we present a detailed analysis of the abilities of different metal ions, including zinc, magnesium, manganese, calcium, and iron, to bind and inhibit PLpro. We also analyzed the impact of EDTA on PLpro activity and divalent metal ion binding. The results revealed that zinc had the most significant inhibitory effect on PLpro, followed by manganese. The presence of EDTA was essential for PLpro activity, likely by chelating trace metal ions that inhibit PLpro. The binding of zinc and manganese to PLpro was confirmed by tryptophan intrinsic fluorescence analysis; however, differential scanning calorimetry revealed that zinc, but not manganese, reduced ΔHcal of PLpro. The results of this study provide a durable framework for future efforts to target PLpro to successfully prevent and treat SARS-CoV-2 infection.

The effects of metal ions and EDTA on the catalytic activity of PLpro

The effects of five metal ions on the catalytic activity of SARS-CoV-2 were evaluated: zinc, magnesium, manganese, calcium, and iron. The activity of PLpro was determined by assaying the proteolytic digestion of the fluorogenic peptide substrate CBZ-LRGG-AMC, where CBZ (carbobenzoxy) is a quencher for the AMC (7-amino-4-methyl coumarin) fluorescent probe [29,30]. Following previous studies, PLpro activity was measured in 50 mM HEPES pH 7.5, 0.5 mM EDTA, and 0.5 mM TCEP at a fixed peptide substrate concentration of 135 µM [31–33]. PLpro had very little activity in 50 mM HEPES pH 7.5 and 0.5 mM TCEP, whereas activity recovered upon the addition of EDTA. Titration with different concentrations of EDTA indicated that 0.5 mM EDTA was optimal for PLpro activity, consistent with a previous report [33]. Therefore, 0.5 mM EDTA was included in all PLpro assays.

In the initial screen, chloride salts of all metals plus zinc sulfate were used at a concentration of 1 mM. Adding 1 mM ZnCl2 or ZnSO4 resulted in almost complete inhibition of PLpro, whereas 1 mM MnCl2 or CaCl2 reduced PLpro activity by ∼40% (Figure 1a). Neither 1 mM MgCl2 nor 1 mM FeCl2 affected the activity of PLpro significantly (Figure 1a). The effects of zinc, manganese, and calcium on SARS-CoV-2 PLpro activity were further investigated by determining their IC50s. The calculated IC50s of the metal ions are dependent on the concentrations of PLpro and the peptide substrate, which were 0.5 µM and 135 µM in the assay, respectively. Zinc exhibited the strongest inhibitory effect on the enzymatic activity of PLpro (Figure 1b,c). The calculated IC50 values for ZnCl2, ZnSO4, and MnCl2 were 0.42 ± 0.02 mM, 0.35 ± 0.01 mM, and 2.6 ± 0.3 mM, respectively (Figure 1d). Although 1 mM CaCl2 reduced PLpro activity by ∼40% in the initial screen, increasing the calcium concentration to 100 mM only reduced PLpro activity by ∼70%; therefore, an IC50 value could not be calculated for calcium.

Effect of metal ions on the activity of PLpro.

Figure 1.
Effect of metal ions on the activity of PLpro.

(a) Bar plot of the percent inhibition of PLpro in the presence of zinc sulfate and chloride salts of zinc, magnesium, manganese, calcium, and iron at a metal ion concentration of 1 mM. The PLpro assay was performed in 50 mM HEPES pH 7.5, 0.5 mM EDTA, and 0.5 mM TCEP at a fixed peptide substrate concentration of 135 µM and a PLpro enzyme concentration of 0.5 µM. The percent inhibition was determined by comparing the rate in the presence and absence of metal ions. (bd) Determination of the IC50 values of the metal ions that reduced the catalytic activity of PLpro in the initial screen. The IC50 of calcium could not be determined because 50% inhibition was not reached at a calcium chloride concentration of 100 mM. Data points are means ± S.D. of triplicate determinations.

Figure 1.
Effect of metal ions on the activity of PLpro.

(a) Bar plot of the percent inhibition of PLpro in the presence of zinc sulfate and chloride salts of zinc, magnesium, manganese, calcium, and iron at a metal ion concentration of 1 mM. The PLpro assay was performed in 50 mM HEPES pH 7.5, 0.5 mM EDTA, and 0.5 mM TCEP at a fixed peptide substrate concentration of 135 µM and a PLpro enzyme concentration of 0.5 µM. The percent inhibition was determined by comparing the rate in the presence and absence of metal ions. (bd) Determination of the IC50 values of the metal ions that reduced the catalytic activity of PLpro in the initial screen. The IC50 of calcium could not be determined because 50% inhibition was not reached at a calcium chloride concentration of 100 mM. Data points are means ± S.D. of triplicate determinations.

Close modal

Since the presence of EDTA in the enzyme assay buffer increased the activity of PLpro, the effect of EDTA on the IC50 of zinc was investigated. The IC50 values for zinc ion reported above were slightly lower than the EDTA concentration of 0.5 mM included in the PLpro enzyme assay buffer. To estimate the inhibitory effect of zinc in the absence of EDTA, the IC50 of ZnCl2 was determined at different concentrations of EDTA: 0.1 mM, 0.5 mM, 1.0 mM, 2.5 mM, 5.0 mM and 10.0 mM EDTA (Figure 2a,b). The IC50 of zinc increased linearly as the concentration of EDTA increased (Figure 2c). Extrapolation of the linear fit of the IC50 of ZnCl2 to the y-intercept gave an estimated IC50 value of 14 µM ZnCl2 in the absence of EDTA (Figure 2c).

Effect of EDTA on the IC50s of metal ions.

Figure 2.
Effect of EDTA on the IC50s of metal ions.

(a and b) The IC50 profiles of ZnCl2 against PLpro were determined in the presence of 0.1 mM, 1.0 mM, 2.5 mM, 5.0 mM, and 10.0 mM EDTA. The peptide substrate concentration and reaction conditions were similar to those in Figure 1. The IC50 was determined at the mid-transition point, giving values of 0.11 ± 0.02 mM, 0.40 ± 0.02 mM, 0.67 ± 0.01 mM, 1.80 ± 0.02 mM, 2.90 ± 0.01 mM, and 7.50 ± 0.15 mM at EDTA concentrations of 0.1 mM, 1 mM, 2.5 mM, 5 mM, and 10 mM, respectively. (c) The plot of the IC50 of ZnCl2 versus EDTA concentration was linear. Extrapolation of the IC50 value to 0 mM EDTA from the linear fit gave a value of 14 µM for ZnCl2. Data points are means ± S.D. of triplicate determinations.

Figure 2.
Effect of EDTA on the IC50s of metal ions.

(a and b) The IC50 profiles of ZnCl2 against PLpro were determined in the presence of 0.1 mM, 1.0 mM, 2.5 mM, 5.0 mM, and 10.0 mM EDTA. The peptide substrate concentration and reaction conditions were similar to those in Figure 1. The IC50 was determined at the mid-transition point, giving values of 0.11 ± 0.02 mM, 0.40 ± 0.02 mM, 0.67 ± 0.01 mM, 1.80 ± 0.02 mM, 2.90 ± 0.01 mM, and 7.50 ± 0.15 mM at EDTA concentrations of 0.1 mM, 1 mM, 2.5 mM, 5 mM, and 10 mM, respectively. (c) The plot of the IC50 of ZnCl2 versus EDTA concentration was linear. Extrapolation of the IC50 value to 0 mM EDTA from the linear fit gave a value of 14 µM for ZnCl2. Data points are means ± S.D. of triplicate determinations.

Close modal

Pattern of PLpro inhibition by metal ions

To further investigate the inhibitory effects of zinc and manganese on PLpro activity, initial velocity patterns were obtained in the presence of varying concentrations of the peptide substrate and different fixed concentrations of the metal salts in the presence of 0.5 mM EDTA. The concentration of the peptide substrate was varied at 55, 110, 250, and 800 µM in the presence of 1.0, 1.5, 2.5, or 6.0 mM MnCl2 (Figure 3). Manganese yielded a mixed inhibition pattern against PLpro with an inhibition constant (Ki) of 1.5 ± 0.4 mM. A similar inhibition study was performed on the zinc salts; however, due to the sharp inhibition window observed during the IC50 calculations, mechanistic characterization of zinc inhibition of PLpro was not experimentally feasible, and we could not accurately estimate the inhibition patterns.

Inhibition pattern of PLpro by manganese.

Figure 3.
Inhibition pattern of PLpro by manganese.

(a) Michaelis–Menten fit graph and (b) Lineweaver-Burk graph of the non-competitive inhibition of PLpro by MnCl2. Points are experimental data, and lines represent the best fit of the data. The assay was performed in 50 mM HEPES pH 7.5, 0.5 mM EDTA, and 0.5 mM TCEP. The concentration of the peptide substrate was varied at 55 µM, 75 µM, 250 µM, and 800 µM. The lines represent increasing concentrations of manganese chloride at 1 mM, 1.5 mM, 2.5 mM, and 6 mM. The calculated Ki for manganese chloride is 1.5 ± 0.4 mM. Data points are means ± S.D. of triplicate determinations.

Figure 3.
Inhibition pattern of PLpro by manganese.

(a) Michaelis–Menten fit graph and (b) Lineweaver-Burk graph of the non-competitive inhibition of PLpro by MnCl2. Points are experimental data, and lines represent the best fit of the data. The assay was performed in 50 mM HEPES pH 7.5, 0.5 mM EDTA, and 0.5 mM TCEP. The concentration of the peptide substrate was varied at 55 µM, 75 µM, 250 µM, and 800 µM. The lines represent increasing concentrations of manganese chloride at 1 mM, 1.5 mM, 2.5 mM, and 6 mM. The calculated Ki for manganese chloride is 1.5 ± 0.4 mM. Data points are means ± S.D. of triplicate determinations.

Close modal

Evaluation of the binding affinity of PLpro for metal ions based on tryptophan fluorescence

SARS-CoV-2 PLpro contains two tryptophan residues (Trp-96 and Trp-109) and 24 tyrosine residues. The two tryptophan residues are located near the active site of PLpro, while the tyrosine residues are spread throughout the protease structure. To estimate the binding affinity of PLpro for metal ions, the change in tryptophan fluorescence upon titration with ZnSO4, ZnCl2, or MnCl2 was measured. The tryptophan emission scans of PLpro were collected at emission wavelengths (λem) from 300 nm to 450 nm using an excitation wavelength (λex) of 290 nm to minimize tyrosine emission (Figure 4a–c). Metal ion titration resulted in changes in fluorescence intensity in the emission spectra in a metal concentration-dependent manner; however, the emission profiles for zinc and manganese did not show a notable blueshift or redshift. In all spectra, the maximum tryptophan fluorescence intensity occurred at an emission wavelength (λmax) of 342 nm (Figure 4a–c).

Tryptophan fluorescence measurement for the determination of metal ion binding affinity.

Figure 4.
Tryptophan fluorescence measurement for the determination of metal ion binding affinity.

(ac) Intrinsic tryptophan fluorescence scans of PLpro in the presence of increasing concentrations of metal ions with excitation at 290 nm. The emission of PLpro initially increased with the addition of ZnSO4, ZnCl2, and MnCl2. Upon titration with higher concentrations of the metal ions, quenching of the tryptophan fluorescence signal was observed. (d and e) Change in fluorescence intensity at the maximum emission wavelength (λem) of 342 nm as a function of metal ion concentration. The data were fit to a one-site specific binding model to determine the KD values for ZnSO4, ZnCl2, and MnCl2.

Figure 4.
Tryptophan fluorescence measurement for the determination of metal ion binding affinity.

(ac) Intrinsic tryptophan fluorescence scans of PLpro in the presence of increasing concentrations of metal ions with excitation at 290 nm. The emission of PLpro initially increased with the addition of ZnSO4, ZnCl2, and MnCl2. Upon titration with higher concentrations of the metal ions, quenching of the tryptophan fluorescence signal was observed. (d and e) Change in fluorescence intensity at the maximum emission wavelength (λem) of 342 nm as a function of metal ion concentration. The data were fit to a one-site specific binding model to determine the KD values for ZnSO4, ZnCl2, and MnCl2.

Close modal

After adding 2 µM ZnSO4 or ZnCl2, the intensity of the intrinsic tryptophan fluorescence signal of PLpro increased sharply by 20% and 60%, respectively, compared with the enzyme in the absence of zinc (Figure 4a,b). Further increasing the zinc concentration beyond 2 µM resulted in quenching of fluorescence below the fluorescence intensity of the free enzyme. Similarly, titration of PLpro with MnCl2 initially increased the emission signal, with a maximum increase in 30% at 50 µM MnCl2 (Figure 4c). Further increasing the manganese concentration to 5 mM resulted in quenching, although the fluorescence intensity remained higher than that of the free enzyme (Figure 4c), in contrast with the titration of PLpro with high concentrations of ZnCl2 and ZnSO4.

To estimate the binding affinity constant (KD) of metal ions to PLpro, the change in the tryptophan fluorescence signal was plotted as a function of the metal ion concentration, and the graphs were fit to a one-site binding model using GraphPad Prism 9 (Figures 4d–f). The effect of protein dilution upon the addition of the metal ion was considered, and a background fluorescence emission scan was obtained in parallel with the ion titration scans by adding buffer only. The reduction in fluorescence intensity due to the addition of buffer was insignificant compared with the quenching of the PLpro signal due to metal ion titration. As expected, the KD values of ZnSO4 (0.48 ± 0.21 mM) and ZnCl2 (0.27 ± 0.17 mM) were 10-fold lower than that of MnCl2 (2.74 ± 0.60 mM), indicating a higher binding affinity of PLpro for zinc compared with manganese.

Effects of metal ions and EDTA on the thermodynamic stability of PLpro

The thermodynamic stability of PLpro was first assessed in the absence of EDTA by DSC in the presence and absence of 0.25 mM or 0.5 mM ZnSO4, ZnCl2, or MnCl2 in 50 mM HEPES pH 7.0 and 0.5 mM TCEP (Figure 5a,b). The temperature was ramped from 15°C to 75°C at a scan rate of 1°C/min to acquire the thermal unfolding transitions. The melting temperature (Tm) was calculated at the apex of the melting peak, and the calorimetric enthalpy (ΔHcal) was determined from the area under the thermographic peak. The DSC scan of PLpro in the absence of metal salts exhibited a two-phase transition with a late shoulder peak (Figure 5a). The overall shape of the thermogram was the same in the presence of ZnSO4, except that the shoulder peak was less prominent (Figure 5a). Tm decreased from 44.2 ± 0.1°C in the absence of ZnSO4 to 35.3 ± 0.4°C at 0.25 mM ZnSO4 and 34.3 ± 0.3°C at 0.5 mM ZnSO4 (Figure 5d). ΔHcal decreased slightly from 1476 ± 113 kJ/mol in the absence of ZnSO4 to 1302 ± 30 kJ/mol at 0.25 mM ZnSO4 (Figure 5f) but exhibited a larger drop to 540 ± 14 kJ/mol at 0.5 mM ZnSO4 (Figure 5f). Adding 0.25 mM ZnSO4 marginally decreased ΔHcal by 174 kJ/mol, while a larger decrease in 936 kJ/mol was observed upon the addition of 0.5 mM ZnSO4.

Effect of metal ions on the thermodynamic stability of PLpro.

Figure 5.
Effect of metal ions on the thermodynamic stability of PLpro.

(ac) DSC thermal scans of PLpro in the absence (black) and presence of 0.25 mM (green) or 0.5 mM (blue) zinc sulfate, zinc chloride, or manganese chloride. The temperature was increased from 20°C to 85°C at scanning rate of 1°C/min in buffer containing 50 mM Hepes pH 7.0 and 0.5 TCEP. (d and e) Bar plots of the Tm1 and Tm2 values calculated from the midpoints of the thermal transitions of the DSC scans. The addition of 0.25 mM metal salt reduced Tm compared with its absence (black); however, further increasing the concentration to 0.50 mM (blue) did not change the Tm value of PLpro. (f) Bar plots of ΔHcal of PLpro at different concentrations of zinc and manganese salts. The addition of 0.25 mM (green) zinc slightly decreased the ΔHcal value of PLpro, whereas manganese did not affect ΔHcal. Increasing the concentration of salt to 0.50 mM (blue) greatly reduced ΔHcal, indicating a change in the overall bonding interactions of the protein structure. Increasing the concentration of zinc decreased the ΔHcal of PLpro more than increasing the concentration of manganese to 0.5 mM. Data points are means ± S.D. of triplicate determinations.

Figure 5.
Effect of metal ions on the thermodynamic stability of PLpro.

(ac) DSC thermal scans of PLpro in the absence (black) and presence of 0.25 mM (green) or 0.5 mM (blue) zinc sulfate, zinc chloride, or manganese chloride. The temperature was increased from 20°C to 85°C at scanning rate of 1°C/min in buffer containing 50 mM Hepes pH 7.0 and 0.5 TCEP. (d and e) Bar plots of the Tm1 and Tm2 values calculated from the midpoints of the thermal transitions of the DSC scans. The addition of 0.25 mM metal salt reduced Tm compared with its absence (black); however, further increasing the concentration to 0.50 mM (blue) did not change the Tm value of PLpro. (f) Bar plots of ΔHcal of PLpro at different concentrations of zinc and manganese salts. The addition of 0.25 mM (green) zinc slightly decreased the ΔHcal value of PLpro, whereas manganese did not affect ΔHcal. Increasing the concentration of salt to 0.50 mM (blue) greatly reduced ΔHcal, indicating a change in the overall bonding interactions of the protein structure. Increasing the concentration of zinc decreased the ΔHcal of PLpro more than increasing the concentration of manganese to 0.5 mM. Data points are means ± S.D. of triplicate determinations.

Close modal

A similar thermodynamic destabilization effect was observed for ZnCl2, with a major two-phase transition (Figure 5a). The Tm values for the early and late thermographic transitions (Tm1 and Tm2) decreased by ∼10°C in the presence of 0.25 mM and 0.5 mM ZnCl2 (Figure 5d,e). Similar to the impact of ZnSO4, ΔHcal decreased slightly from 1476 ± 113 kJ/mol in the absence of ZnCl2 to 1149 ± 122 kJ/mol in the presence of 0.25 mM ZnCl2, a decrease in 327 kJ/mol (Figure 5f). The addition of 0.5 mM ZnCl2 resulted in an even larger drop to 634 ± 41 kJ/mol, a decrease in 842 kJ/mol.

The DSC thermograms of PLpro in the presence of MnCl2 also exhibited a two-phase transition. Adding 0.25 mM MnCl2 shifted the thermographic peak to lower temperatures but did not significantly affect the peak height (Figure 5c). The Tm1 and Tm2 values for the first and second transitions decreased by 13°C and 9°C, respectively, upon adding 0.25 mM MnCl2 (Figure 5d,e). The ΔHcal values were identical in the absence or presence of 0.25 mM MnCl2 (Figure 5f). Increasing the concentration of MnCl2 to 0.5 mM did not further decrease Tm1 and Tm2 compared with 0.25 mM MnCl2. However, a drop in ΔHcal of 490 kJ/mol was observed in the presence of 0.5 mM MnCl2, whereas no change in ΔHcal was observed upon adding 0.25 mM MnCl2.

Next, the impact of EDTA on the binding of zinc to PLpro was assessed. First, the thermodynamic stability of PLpro was assessed in the free enzyme state in the absence and presence of equal concentrations of 0.5 mM EDTA and 0.5 mM ZnSO4. Under the latter conditions, effectively none of the zinc ions were available for binding to PLpro. The overall shapes of the two DSC thermograms were similar, except that the small shoulder of the late transition that was present in the absence of EDTA disappeared in the presence of equal concentrations of EDTA and ZnSO4 (Figure 6a). Tm decreased by 2°C in the presence of equal concentrations of EDTA and ZnSO4 (Figure 6d), whereas ΔHcal was similar under the two conditions (Figure 6e).

Effect of EDTA on the thermodynamic stability of PLpro in the presence of zinc sulfate.

Figure 6.
Effect of EDTA on the thermodynamic stability of PLpro in the presence of zinc sulfate.

(a) DSC thermal scans of PLpro in the free enzyme (black) state. In the absence of EDTA, (solid line), the DSC scan of PLpro was collected for the free enzyme. In the presence of 0.5 mM EDTA (dashed line), an equivalent concentration of 0.5 mM ZnSO4 was added. The DSC scanning conditions and parameters were similar to those in Figure 3. (b) DSC scans of PLpro in the presence of 0.25 mM ZnSO4 state (green). In the absence of EDTA (solid line), 0.25 mM ZnSO4 was included; however, in the presence of 0.5 mM EDTA (dashed line), 0.75 mM ZnSO4 was added. (b) DSC scans of PLpro in the presence of 0.5 mM ZnSO4. In the absence of EDTA (solid blue line), 0.5 mM of ZnSO4 was included; however, in the presence of 0.5 mM EDTA (dashed red line), 1 mM ZnSO4 was added. (d) Bar plot of Tm values for PLpro at different concentrations of ZnSO4 in the presence (solid bars) and absence (checkered bars) of 0.5 mM EDTA. The presence of EDTA did not affect the Tm of PLpro when an equivalent concentration of ZnSO4 was present. Similar to the measurements in Figure 5, the presence of 0.25 mM ZnSO4 reduced the Tm of PLpro; however, increasing the concentration of ZnSO4 to 0.5 mM did not change the Tm of PLpro regardless of the presence or absence of EDTA. (e) Bar plots of ΔHcal of PLpro at different concentrations of ZnSO4 in the presence (solid bars) and absence (checkered bars) of 0.5 mM EDTA. The addition of EDTA did not alter the ΔHcal values of PLpro when an equivalent concentration of ZnSO4 was present. Data points are means ± S.D. of triplicate determinations.

Figure 6.
Effect of EDTA on the thermodynamic stability of PLpro in the presence of zinc sulfate.

(a) DSC thermal scans of PLpro in the free enzyme (black) state. In the absence of EDTA, (solid line), the DSC scan of PLpro was collected for the free enzyme. In the presence of 0.5 mM EDTA (dashed line), an equivalent concentration of 0.5 mM ZnSO4 was added. The DSC scanning conditions and parameters were similar to those in Figure 3. (b) DSC scans of PLpro in the presence of 0.25 mM ZnSO4 state (green). In the absence of EDTA (solid line), 0.25 mM ZnSO4 was included; however, in the presence of 0.5 mM EDTA (dashed line), 0.75 mM ZnSO4 was added. (b) DSC scans of PLpro in the presence of 0.5 mM ZnSO4. In the absence of EDTA (solid blue line), 0.5 mM of ZnSO4 was included; however, in the presence of 0.5 mM EDTA (dashed red line), 1 mM ZnSO4 was added. (d) Bar plot of Tm values for PLpro at different concentrations of ZnSO4 in the presence (solid bars) and absence (checkered bars) of 0.5 mM EDTA. The presence of EDTA did not affect the Tm of PLpro when an equivalent concentration of ZnSO4 was present. Similar to the measurements in Figure 5, the presence of 0.25 mM ZnSO4 reduced the Tm of PLpro; however, increasing the concentration of ZnSO4 to 0.5 mM did not change the Tm of PLpro regardless of the presence or absence of EDTA. (e) Bar plots of ΔHcal of PLpro at different concentrations of ZnSO4 in the presence (solid bars) and absence (checkered bars) of 0.5 mM EDTA. The addition of EDTA did not alter the ΔHcal values of PLpro when an equivalent concentration of ZnSO4 was present. Data points are means ± S.D. of triplicate determinations.

Close modal

Similar DSC analyses were performed to assess the effects of 0.25 mM and 0.5 mM ZnSO4 on the thermodynamic stability of PLpro in the presence or absence of EDTA. In the absence of EDTA, only 0.25 mM and 0.5 mM ZnSO4 were included. In contrast, 0.75 mM and 1.0 mM ZnSO4 were added in the presence of 0.5 mM EDTA to obtain unchelated zinc ion concentrations of 0.25 mM and 0.5 mM, respectively. The shapes of the DSC thermographic peaks for the 0.25 mM and 0.5 mM ZnSO4 states were similar in the presence or absence of 0.5 mM EDTA (Figure 6b,c). Tm was ∼34.5°C in the presence of 0.25 mM ZnSO4 regardless of the presence of 0.5 mM EDTA, corresponding to a decrease in 9°C compared with the Tm of free PLpro (Figure 6d). However, Tm did not decrease further upon increasing the ZnSO4 concentration to 0.5 mM. In addition, ΔHcal of PLpro decreased slightly upon adding 0.25 mM ZnSO4, with decreases of 174 kJ/mol and 283 kJ/mol in the absence or presence of 0.5 mM EDTA, respectively (Figure 6e). However, upon adding 0.5 mM ZnSO4, ΔHcal decreased significantly by 936 kJ/mol and 984 kJ/mol in the absence or presence of 0.5 mM EDTA, respectively.

Effects of metal ions on the structural integrity of PLpro

In the presence of metal ions, the Tm of PLpro decreased by ∼10°C, indicating a decrease in thermodynamic stability. To confirm that the inhibition of PLpro by zinc or manganese was not a result of denaturation of the protease, the effects of metal ions on the structural integrity of PLpro were assessed by far-UV CD spectroscopy. First, CD spectra of PLpro in the native and denatured states were acquired at 25°C and after incubation for 30 min at 80°C, respectively (Figure 7a). The CD spectra were collected from 200 nm to 260 nm, and each spectrum was the average of five CD scans. The CD spectrum of PLpro in the native state showed two ellipticity minima at 208 and 222 nm, consistent with mixed α-helical and β-sheet structures [5,34]. In the denatured state, a total loss of the secondary structural elements of PLpro was observed, as indicated by the reduction in its ellipticity.

Far-UV circular dichroism scans of PLpro.

Figure 7.
Far-UV circular dichroism scans of PLpro.

Far-UV CD spectra of PLpro were collected from 260 to 200 nm. (a) The native (blue) and denatured (red) states of PLpro were collected at 25°C and after incubation for 30 min at 80°C, respectively. (b and c) Far-UV CD spectra of native PLpro collected at 25°C in the absence (black) and presence of different concentrations of zinc and manganese metal ions. The concentration of ZnCl2 was varied at 0.5 mM (green), 1.0 mM (red), and 5.0 mM (green), while the concentration of MnCl2 was varied at 1.0 mM (green), 5.0 mM (red), and 10.0 mM (green). Each spectrum was corrected for the buffer background and is the average of five CD scans.

Figure 7.
Far-UV circular dichroism scans of PLpro.

Far-UV CD spectra of PLpro were collected from 260 to 200 nm. (a) The native (blue) and denatured (red) states of PLpro were collected at 25°C and after incubation for 30 min at 80°C, respectively. (b and c) Far-UV CD spectra of native PLpro collected at 25°C in the absence (black) and presence of different concentrations of zinc and manganese metal ions. The concentration of ZnCl2 was varied at 0.5 mM (green), 1.0 mM (red), and 5.0 mM (green), while the concentration of MnCl2 was varied at 1.0 mM (green), 5.0 mM (red), and 10.0 mM (green). Each spectrum was corrected for the buffer background and is the average of five CD scans.

Close modal

Next, far-UV CD spectra of PLpro were collected in the presence of different concentrations of ZnCl2 or MnCl2 at 25°C (Figure 7b,c). The addition of either metal ion induced structural changes in SARS-CoV-2 PLpro, but the folded state was maintained. The addition of 0.5 mM and 1.0 mM ZnCl2 increased the negative ellipticity of PLpro at 208 nm by 11% and 3%, respectively, compared with the enzyme in the absence of zinc (Figure 7b). The increase in the negative ellipticity of PLpro in the presence of zinc indicates an enhancement of its secondary structural content. However, the CD spectrum of PLpro in the presence of 5.0 mM ZnCl2 was similar to that of PLpro in the absence of metal ions. Thus, depending on its concentration, the presence of zinc either enhanced or had no effect on the structural integrity of PLpro.

Similarly, the addition of manganese increased the negative ellipticity of PLpro by 6%, 16%, and 13% in the presence of 1.0 mM, 5.0 mM, and 10.0 mM MnCl2, respectively (Figure 7c). Thus, both manganese and zinc enhanced the structural integrity of PLpro. The metal ion concentrations used in CD spectroscopy were higher than the KD values of 0.27 mM and 2.74 mM for ZnCl2 and MnCl2 from tryptophan fluorescence measurements, respectively; therefore, the observed inhibition of PLpro by zinc and manganese reflects the loss of enzymatic activity rather than denaturation of the protease.

Interactome of PLpro with host proteins and the relevance of Zn2+ and Mn2+

We considered only hub proteins and those human proteins directly connected to the hub proteins in the interactome analysis. Hub proteins have the most connections with other human proteins in the network and best represent the overall functional relevance of the proteins in the network. Human proteins associated with SARS-CoV-2 PLpro were classified as Level 1 or Level 2 (Figure 8). Level 1 human proteins are in the immediate vicinity of or directly connected to SARS-CoV-2 PLpro in the interactome network, whereas Level 2 human proteins are indirectly connected to PLpro through the Level 1 proteins (Figure 8).

Protein clusters highlighting interactions between PLpro and host core proteins with level 1 and level 2 interactions.

Figure 8.
Protein clusters highlighting interactions between PLpro and host core proteins with level 1 and level 2 interactions.

The large circles correspond to the central proteins (Hub-bottleneck — black hand pointing, bottleneck — double asterisks). The first interactions of PLpro are highlighted with wide grey connectors. The smaller circles correspond to second-degree interactions with PLpro (second PLpro neighborhood). The yellow triangles represent Zn2+ and Mn2+. The dashed black lines represent the inhibitory activity of Zn2+ and Mn2+ on PLpro. The collective functional annotation of each protein cluster in the interactome for both level 1 and level 2 interactions between PLpro and host proteins is noted. The figure was constructed by integrating centrality, cluster, and functional enrichment analyses using the Cytoscape program [35]. Edits were performed in CorelDRAW®.

Figure 8.
Protein clusters highlighting interactions between PLpro and host core proteins with level 1 and level 2 interactions.

The large circles correspond to the central proteins (Hub-bottleneck — black hand pointing, bottleneck — double asterisks). The first interactions of PLpro are highlighted with wide grey connectors. The smaller circles correspond to second-degree interactions with PLpro (second PLpro neighborhood). The yellow triangles represent Zn2+ and Mn2+. The dashed black lines represent the inhibitory activity of Zn2+ and Mn2+ on PLpro. The collective functional annotation of each protein cluster in the interactome for both level 1 and level 2 interactions between PLpro and host proteins is noted. The figure was constructed by integrating centrality, cluster, and functional enrichment analyses using the Cytoscape program [35]. Edits were performed in CorelDRAW®.

Close modal

Twenty-seven host proteins directly interact with PLpro (Level 1). The functions of these host proteins include transcription, ribosomal function, actin-cytoskeleton, mitochondrial function-ADP ribosylation, cell cycle, RNA processing, endoplasmic reticulum and function, phospholipid and fatty acid synthesis, ubiquitination, histone function, chromatin remodeling and nucleosome function, zinc binding (TRIM13), and antiviral immunity. All these processes are highly relevant to the interaction between the SARS-CoV-2 virus and the human host. Among these Level 1 proteins, 4 are hub-bottleneck proteins (RPL30, NOB1, NDUFS1, RPL21), and 23 are bottleneck proteins (SRPR, FDFT1, ACSL3, EMC8, NDC1, STIM1, STIM2, ARPC1B, HIST1H1B, AGPAT2, STT3A, DDRGK1, EMC1, TRIM13, CCNB2, ARL6IP5, LSM1, ESPL1, SLC27A4, AAAS, EMC7, CAMLG, USP3). Bottleneck proteins function as connections between functional clusters and have high betweenness centrality; i.e. many short paths in the network pass through these points. Thus, bottleneck proteins provide key connections in the network and are more likely to be essential proteins than hub proteins [36]. Hub-bottleneck proteins are hub proteins that are directly connected to bottleneck proteins; they are frequently central proteins that link multiple complexes or peripheral components of central complexes [36].

Previous studies have identified RPL8, RPSA, RPL12, EEF1A1, RPL6, RPL15, RPS11, RPS16, RPL11, and RPS15A [37–42] as the top 10 Level 1 hub proteins interacting with SARS-CoV-2 PLpro [43]; KEGG pathway analysis by Ghosh et al. [43] showed that these proteins are involved in ribosome function and RNA transport. Based on the high degree of interaction of the hub proteins in the present analysis, we can add three hub-bottleneck proteins — RPL30, NOB1 and RPL21 — to the proteins reported by Ghosh et al. The networks of these three hub bottleneck proteins appear to be involved in translational elongation and its regulation, ribonucleoprotein complex biogenesis, intracellular protein transport RNA splicing and ribosome assembly (Figure 8). In addition, we identified a major cluster connected to NDUFS1, which is directly connected to PLpro (Figure 8). NDUFS1 is a multi-subunit NADH:ubiquinone oxidoreductase (complex I; EC 1.6.5.3) and is the first enzyme complex in the mitochondrial electron transport chain. All of the proteins in NDUFS1's network appear to be involved in the mitochondrial electron transport chain, chromatin reorganization, regulation of macrophage differentiation and histone h4-r3 methylation, which are highly relevant to SARS-CoV-2 infection.

We also analyzed the interactions of manganese and zinc with host proteins that interact with SARS-CoV-2 PLpro. Both clearly interact with host metalloproteinases, proteins that are functionally relevant to antiviral and innate immunity processes, in addition to host proteins involved in the regulation of the influx and efflux of metal ions from cells (e.g. CASP3, SOD-1, HSP90AA1, and NOS-3). Metal ion regulation, which is integral to immunity against viral infection, is subverted by SARS-CoV-2 infection. Of particular note is the STRING network of the bottleneck protein CCNB2 [44,45], which comprises 10 proteins, including CCNA2, CDK2, CDC25A and BUB1B. These four proteins are mainly involved in cell cycle function and control of G1/S and G2/M transition phases of the cell cycle. They also show a close interaction with Mn2+ (Figure 8). Manganese is an essential nutrient for bacteria and viruses and an essential cofactor for numerous cellular processes. When viruses infect a vertebrate host, a hostile environment with minimal bioavailable manganese is created via sequestration of manganese at the focus of infection. To combat this host strategy, pathogens can express high-affinity manganese importers that are vital in inducing virulence [46]. The host–pathogen interface is largely a battle for manganese, which shapes the pathogenesis of infection. Thus, the network of CCNB2 is highly relevant during the infectious process induced by SARS-CoV-2, and its interaction with PLpro and manganese is not coincidence but is functionally relevant to SARS-CoV-2 infection.

Several targets have been identified for the development of antivirals against SARS-CoV-2. The most promising are the viral proteases. Both PLpro and the main protease (Mpro) are vital in the processing of the viral polyproteins (pp1a and pp1ab), which is a prerequisite for the maturation of the non-structural viral proteins [47,48]. PLpro, which originates from the membrane anchored multi-domain protein nsp3 [15], also suppresses antiviral innate immunity by reversing ubiquitination and ISGylation [9,24,49–52]. This dual functionality of PLpro makes it an attractive antiviral drug target [19,53,54]. Moreover, PLpro is highly conserved among coronaviruses, with 82% and 30% amino acid sequence identity of SARS-CoV-2 PLpro with SARS-CoV PLpro and MERS PLpro, respectively, suggesting that drugs targeting PLpro could be used to prevent and treat a broad range of coronavirus infections.

Multiple cysteine residues are essential for the function and stability of PLpro. In addition to the catalytic Cys111 residue, Cys189, Cys192, Cys223 and Cys226 in the finger domain of PLpro tetrahedrally coordinate a zinc cation that is essential for structural stability and protease activity [9,24]. Consistent with this structural role of zinc, zinc-ejector drugs reduce PLpro activity. However, zinc has also been shown to inhibit SARS-CoV-1 and MERS PLpro [28,34]. Zinc has been utilized clinically to reduce the symptoms of COVID-19 and enhance antiviral immunity against SARS-CoV-2, and patients with severe COVID-19 symptoms have been shown to possess low zinc levels [55–57]. Moreover, zinc possesses antiviral properties against various coronaviruses, hepatitis C virus, and HIV [56,58,59]. Beyond PLpro, zinc inhibits both SARS-CoV-1 and equine arterivirus RNA polymerase [60], and an IC50 of zinc of 50 µM has been reported for SARS-CoV-1 RNA-dependent RNA polymerase. Overall, these findings suggests that zinc in the right form and quantity can inhibit different biological targets essential for viral replication, including RNA synthesis, topoisomerase, and polyprotein processing [59].

The effects of zinc and other metal ions on SARS-CoV-2 PLpro have not been comprehensively investigated. To address this gap, this study investigated the effects of various divalent metal cations on the activity and biophysical properties of PLpro, as a mechanistic and structural understanding of the effects of metal ions is necessary to develop effective antiviral drug combinations.

Zinc and manganese inhibit PLpro activity

In addition to zinc, the effects of manganese, magnesium, iron, and calcium, which are commonly found in the body and bind to proteins in coordination complexes, were investigated [61]. Neither magnesium nor iron impacted PLpro activity, and the inhibitory effect of calcium was too weak to warrant further investigation. The IC50 of zinc was ∼0.4 mM when 0.5 mM EDTA was included in the assay. This value is much higher than the IC50 of 1.3 µM reported for SARS-CoV PLpro [28,34]. However, this discrepancy may reflect the absence of EDTA in the assays of SARS-CoV PLpro. In the presence of increasing concentrations of EDTA, the IC50 of zinc increased, indicating that zinc inhibits PLpro only after saturating the zinc-EDTA interaction. The extrapolation of the linear fit of the IC50 of zinc as a function of EDTA concentration yielded an estimated IC50 of zinc of 14 µM in the absence of EDTA (Figure 5c). This analysis indicates a strong inhibitory effect of zinc on SARS-CoV-2 PLpro.

Our study is the first to report an inhibitory effect of manganese on SARS-CoV-2 PLpro; a previous screen of divalent metal ions found that SARS-CoV PLpro was not inhibited by manganese [28]. The use of manganese along with other metal ions and vitamins to relieve the symptoms of COVID-19 has been proposed [62]. The IC50 of manganese for PLpro was 2.6 mM, higher than that for zinc. In addition, manganese exhibited a mixed inhibition pattern, indicating that manganese metal ions bind to either free enzyme or the enzyme-substrate complex at a site other than the active site. Unfortunately, a similar analysis of the inhibition pattern of zinc was experimentally unfeasible due to the steep slopes of the IC50 curves for zinc and its narrow inhibition range.

In this study, adding EDTA was essential for PLpro activity, and a concentration of 0.5 mM EDTA was found to be optimal, consistent with previous work [33]. The requirement for EDTA can be explained by the observed inhibition of PLpro by zinc and manganese; slight traces of zinc (or other metals) in the purified enzyme or buffer are sufficient to inhibit SARS-CoV-2 PLpro. Upon increasing the zinc concentration beyond the EDTA concentration, PLpro activity dropped sharply. As a result, the acquired IC50 values were lower than the concentration of EDTA. The true IC50s of ZnSO4 and ZnCl2 are lower than the apparent IC50s determined in the presence of 0.5 mM EDTA reported here of 0.35 ± 0.01 mM and 0.42 ± 0.02 mm, respectively; we estimated an IC50 for ZnCl2 of 14 µM in the absence of EDTA.

The binding affinity of PLpro for zinc and manganese was assessed by measuring intrinsic tryptophan fluorescence as a function of metal ion concentration. Consistent with the effects of zinc and manganese on activity, the KD values for ZnSO4 (KD, 0.48 ± 0.21 mM) and ZnCl2 (KD, 0.27 ± 0.17 mM) were 10-fold lower than that for MnCl2 (KD, 2.74 ± 0.60 mM), indicating higher binding affinity of PLpro for zinc compared with manganese. Increasing the concentration of zinc or manganese quenched intrinsic tryptophan fluorescence without shifting λmax, implying that the electronic environment around the tryptophan residues of PLpro did not change enough to shift the emission signal of tryptophan [63–65]. Since the zinc finger domain is integral to the structural stability of PLpro, we assume that no zinc replacement occurred in this domain in the intrinsic fluorescence experiments.

Metal binding induces conformational changes and reduces the thermodynamic stability of PLpro

Due to the large number of cysteine residues and the chemical nature of the surface of PLpro, this protease is capable of binding a variety of ions, including non-structural zinc ions [60]. In the present study, far-UV CD spectroscopy confirmed that the addition of zinc or manganese induced conformational changes in SARS-CoV-2 PLpro while preserving a significant amount of secondary structure (Figure 7b,c). A large structural shift in the zinc-binding domain is also evident when the crystal structure of SARS-CoV-2 PLpro is compared with that of MERS PLpro [59,60]. The confirmational changes observed by CD spectroscopy upon the addition of zinc or manganese reflect the formation of a new inhibited state of SARS-CoV-2 PLpro, which may explain the molecular mechanism of PLpro inhibition by zinc and manganese.

Overall, adding zinc or manganese decreased the thermodynamic stability of PLpro and its melting point as assessed by DSC. However, a pronounced change in stability was observed only upon increasing the concentration of metal ions to 0.5 mM. The substantial decrease in ΔHcal of PLpro with increasing ZnSO4 concentration but no change in Tm indicates that zinc alters the protein bonding characteristics of PLpro. Proteins with hydrophobic character exhibit negative exothermic peaks on DSC thermograms, whereas polar proteins yield positive endothermic peaks [66,67]. The decrease in the ΔHcal of PLpro implies changes in protein bonding behavior from hydrogen-bonding and polar to hydrophobic interactions, indicating that high zinc concentrations alter the structural characteristics of SARS-CoV-2 PLpro.

Intrinsic relationship between metal ions and host–virus interactions

Studying the centrality of proteins in interaction networks is crucial to identify essential elements of the biosystem. Perturbations (deletion, overexpression, etc.) of these elements can have a large impact on the biosystem. Proteins with high network connectivity, i.e. hubs, are considered more important for the system than those with low connectivity. There are many proteins in the Zn2+ and Mn2+ networks, and in this study, we only selected hub proteins encompassing PLpro, Zn2+ and Mn2+ for analysis and discussion. For instance, both Zn2+ and Mn2+ interact with HSP90AA1 (heat shock protein 90 kDa alpha, class A, member 1), also known as HSP90alpha, a member of the large family of HSP molecular chaperones (Figure 8). HSPs are involved in responses to diverse sources of stress that disrupt protein conformation, such as ethanol, arsenite, cadmium, zinc, copper, mercury, sulfhydryl reagents, calcium ionophores, steroid hormones, chelating agents, and viruses [68].

Zn2+ directly influences apoptotic regulators, especially the caspase family of enzymes, which also indirectly control the decline in intracellular Zn2+ below a critical threshold. Low intracellular Zn2+ levels may not only trigger pathways leading to caspase activation but may also facilitate the process by which caspases are activated by viruses — an epigenetic process that viruses use to exploit the host antiviral immunity. Studies of airway epithelial cells have shown that Zn2+ colocalizes with the precursor form of caspase-3, mitochondria and microtubules, suggesting that Zn2+ is critically positioned to control apoptosis. In the interactome, Zn2+ is connected to caspase-3, a protein subverted by SARS-CoV-2 PLpro [69]. Another member of the Zn2+ and PLpro network is superoxide dismutase (SOD1). SOD1 is abundant in cells throughout the body and binds copper and zinc to break down toxic charged oxygen molecules called superoxide radicals. Plasma proteomics analyses have revealed that SOD1 is a possible predictor of COVID-19 progression [70].

The Zn2+ and PLpro network also includes nitric oxide synthase (NOS-3), which produces nitric oxide (NO). NO plays important roles in the neuronal, muscular, cardiovascular, and immune systems and in viral infections. Zinc is an important structural element of NOS and inhibits its catalytic activity. Zinc down-regulates the mRNA and protein expression of inducible nitric oxide synthase (iNOS) and decreases cytokine-mediated activation of the iNOS promoter. Zinc-mediated regulation of iNOS expression in endothelial cells is mainly attributed to the inhibition of NF-κB transactivation activity. Decreases in iNOS activity are observed during SARS-CoV-2 infection, and COX-2 expression seems to diminish NF-kB and AP-1 activation [71,72]. Decreased bioavailability of NO and increased inflammation are hallmarks of endothelial dysfunction and are prominent in SARS-CoV-2 infection (especially during acute respiratory distress syndrome (ARDS)), which is thought to be an endothelial disease. During severe SARS-CoV-2 infection and ARDS, a cytokine storm accompanied by decreases in NO, IL-6, MCP-1, TNF-alpha, IL-1B and IFN-gamma production in macrophages is observed [72]. Clinical studies of SARS-CoV-2 infection have shown that NO inhibits virus entry into host cells, viral replication, the host immune response, and subsequent thromboembolic complications. Therefore, restoring NO bioavailability may hold potential as a preventive or early treatment option for COVID-19 [73].

The interactome is consistent with the fact that zinc acts at multiple levels in cells to combat SARS-CoV-2, which makes it difficult to define a single cellular or biochemical mechanism of action. Recent studies have shown that different zinc salts possess antiviral activity against SARS-CoV-2 [74,75]. The antiviral activity of zinc reflects a series of effects on viral binding, penetration, uncoating, and replication [76]. As a membrane stabilizer, zinc directly inhibits coronaviral entry, polyprotein processing, and RNA-dependent RNA polymerase (RdRP) activity [77], consistent with our interactome analysis. Moreover, zinc is involved in the production of interleukin-12 (IL-12) and interferon (IFN) by stimulating macrophages. In particular, IL-12 activates natural killer and cytotoxic T cells to effectively combat the virus. The interactome data linking Zn2+ and Mn2+ with the positive regulation of macrophage differentiation strongly align with these effects.

Finally, Mn2+ not only inhibits PLpro but also directly interacts with CCNB2, which is associated with the regulation of ubiquitin-protein ligase activity involved in the mitotic cell cycle and is highly relevant in epigenetic modulation and host antiviral activity in response to SARS-CoV-2. The importance of CCNB2 and its functional network [44,45], which includes 10 proteins, is further supported by the fact that four of those proteins — CCNA2, CDK2, CDC25A and BUB1B — have close interactions with Mn2+ in the interactome (Figure 8). Interestingly, all 4 proteins control the G1/S and G2/M transition phases of the cell cycle and depend strongly on manganese availability [46]. Their interaction with PLpro further emphasizes that SARS-CoV-2 subverts host cell cycle regulation through a battle for manganese — an essential nutrient for a number of cellular processes and host immunity. This finally cements the biological basis of metal ions (mainly Zn2+ and Mn2+) and their intrinsic role in viral infections, along with their therapeutic use in drug formulations in conjunction with antiviral agents.

Conclusion

Zinc is widely used to control viral infections. The present study demonstrates that Zn2+ and Mn2+ suppress PLpro activity and thus may provide synergistic support in treating COVID-19 if used as supplements along with antiviral drugs. As the study here has shown strong evidence in favor of the biochemical activity of metal ions, we believe that future therapeutic strategies would benefit from combining Zn2+ and Mn2+ with antiviral drugs in effectively controlling SARS-CoV-2 infection.

Expression and purification of PLpro

The recombinant PLpro gene with an N-terminal Hisx6 tag was introduced into the pET-28b (+) bacterial expression vector by GenScript, Inc. (Piscataway, NJ). The vector was used to transform E. coli BL21-CodonPlus-RIL (Stratagene) for protein expression. The inoculated culture (3–6 L) was grown at 30 °C in lysogeny broth (LB) with 100 mg/L kanamycin, 50 mg/L chloramphenicol, and 1 µM ZnCl2 until the absorbance at 600 nm reached 0.5. The temperature was then reduced to 15°C, and protein expression was induced overnight (14–16 h) by adding 1 mM IPTG. The cells were pelleted by centrifugation at 12 000×g and 4°C for 10 min in an Avanti J26-XPI centrifuge (Beckman Coulter Inc.). The cell pellets were homogenized in lysis buffer (25 mM Tris, 150 mM NaCl, 5 mM imidazole, 3 mM β-mercaptoethanol (βME), and 0.1% protein inhibitor cocktail (Sigma–Aldrich: P8849) before sonication on ice. The cell lysate was then centrifuged at 40 000×g and 4°C for 45 min. The supernatant was loaded onto a nickel column pre-equilibrated with binding buffer (25 mM Tris, 150 mM NaCl, 5 mM imidazole, and 3 mM βME). The column was washed with binding buffer supplemented with 25 mM imidazole, followed by elution with binding buffer supplemented with 300 mM imidazole. The nickel column fractions containing PLpro were pooled and loaded onto a HiLoad Superdex 200 size-exclusion column (GE Healthcare) on an ÄKTA pure 25 chromatography system (Cytiva, U.S.A.). The gel filtration column was pre-equilibrated with 20 mM HEPES pH 7.5, 150 mM NaCl, and 0.5 mM tris(2-carboxyethyl)phosphine (TCEP). The PLpro-containing fractions were collected and concentrated to ∼150 µM. The sample concentration and purity were determined by the Bradford assay and SDS–PAGE, respectively.

Enzymatic assay and inhibition studies

PLpro enzymatic activity and inhibition analyses were performed using a Cytation 5 multi-mode microplate reader (BIOTEK Instruments, Winooski, VT). The peptide substrate contained the PLpro cleavage site Leu-Arg-Gly-Gly flanked by a fluorescent AMC (7-amino-4-methyl coumarin) tag and corresponding quencher CBZ (carbobenzoxy) (i.e. Z-LRGG-AMC). The proteolytic reaction was initiated by adding 0.5 µM PLpro enzyme to the peptide in buffer (50 mM HEPES pH 7.5, 0.5 mM EDTA, and 0.5 mM TCEP) containing 2% DMSO to enhance the solubility of the peptide [78]. The PLpro reaction was monitored in a 96-well microplate at 25 °C for 10 min at excitation and emission wavelengths of 340 nm and 487 nm, respectively. Cleavage of the peptide substrate by PLpro resulted in an increase in the AMC fluorescent signal.

The inhibitory effects of zinc, manganese, and calcium on the proteolytic activity of PLpro were investigated by determining their half maximal inhibitory concentrations (IC50s). The IC50s of the different metal ions were determined at a fixed peptide substrate concentration of 135 µM by serially diluting the metal by a factor of 1.25 from 0 µM to 600 µM for ZnCl2 and ZnSO4 and from 0 mM to 10 mM for CaCl2 and MnCl2. The IC50 data were fit to a four-parameter dose-response (variable slope) model in GraphPad Prism 9 (GraphPad, San Diego, CA).

For the inhibition studies, the peptide substrate concentration was varied at 45 µM, 55 µM, 75 µM, 110 µM, 250 µM, and 800 µM at different fixed metal ion concentrations around their IC50 values ranging from 300 µM to 480 µM for zinc salts and 2 mM to 10 mM for MnCl2. The inhibition data for MnCl2 were fit to a mixed-inhibition Michaelis–Menten equation using SigmaPlot (Systat Software, Inc., San Jose, CA):
v=Vmax(Km[S])(1+[I]Ki)+(1+[I]Ki)
where v and Vmax are the measured initial rate and maximum rate, respectively, [S] and [I] are the substrate and inhibitor concentrations, respectively, Km is the Michaelis constant for the substrate, and Ki is the inhibition constant.

Tryptophan fluorescence spectroscopy (TFS) scans of PLpro

TFS measurements of PLpro were performed in a quartz cuvette in a fluorescence spectrometer (HORIBA Fluorometer4, HORIBA Instruments Incorporated, Edison, NJ) with excitation and emission pathlengths of 4 mm and 10 mm, respectively. The emission spectra were acquired from 300 nm to 450 nm with slit width of 1 nm at an excitation wavelength of 290 nm. The temperature was maintained at 25°C, and the PLpro protease concentration was 10 µM in 50 mM HEPES pH 7.5, and 0.5 mM TCEP. TFS scans were collected upon titrating different concentrations of ZnCl2 and ZnSO4 in two ranges: from 0 µM to 2 µM and from 2 µM to 1000 µM. The MnCl2 concentration was varied from 5 µM to 5000 µM. The change in the tryptophan fluorescence signal was plotted as a function of metal ion concentration to determine the binding affinity constant (KD) of metal ions to PLpro. The plots were fit to a one-site binding model using GraphPad Prism 9. In parallel with the ion titration scans, TFS scans were obtained by adding buffer only to account for the effect of protein dilution upon addition of the metal ion. All measurements were performed in triplicate.

Differential scanning calorimetry (DSC)

The thermodynamic stability of PLpro was assessed by DSC in the presence or absence of metal ions as well as EDTA using a Nano DSC instrument (TA Instruments) at 0.25 mM and 0.5 mM ZnSO4, ZnCl2, and MnCl2. The instrument was calibrated using chicken egg white lysozyme, a known external Nano DSC standard that is part of the TA Instruments test kit (602198.901). The DSC scans were acquired at a PLpro concentration of 30 µM in buffer containing 50 mM HEPES pH 7.5 and 0.5 mM TCEP in the presence and absence of 0.5 mM EDTA. The protein samples were heated from 15°C to 75°C at a scan rate of 1°C/min and 3 atm pressure. The DSC thermogram was corrected by subtracting the baseline and converted to a plot of excess heat capacity (Cp) as a function of temperature. A corresponding buffer baseline was collected for each measurement by applying a second heating scan on the protein sample. The melting temperature (Tm) was determined at the maximum temperature of the thermal transition, and the calorimetric enthalpy (ΔHcal) of the transition was estimated from the area under the thermal transition using Nano Analyzer software (TA Instruments). DSC measurements were performed in duplicate.

Far-UV circular dichroism (CD) spectroscopy

CD spectra of PLpro were collected in 10 mM Tris pH 7.5. Phosphate buffer was not used because precipitation of the enzyme was observed when PLpro was incubated with phosphate buffer. Since zinc phosphate is water insoluble, the low stability of PLpro in phosphate buffer may be due to instability of the structural zinc in the zinc-binding domain, which has been shown to be important for the structural stability and protease activity of PLpro [9,24]. Scans were collected in a Chirascan CD spectrometer (Applied Photophysics) calibrated with aqueous camphor-10-sulfonic acid (CSA) from 200 nm to 260 nm at a scanning speed of 10 nm/sec and bandwidth of 1 nm using 30 µM PLpro in a 1 mm quartz cuvette. The temperature was maintained at 25°C and 80°C to acquire spectra of the native and denatured states of PLpro, respectively. All CD scans were corrected for buffer and substrate background, and five CD scans were averaged to obtain each spectrum.

Interactome analysis

Interactions between SARS-CoV-2 and Homo sapiens proteins were taken from the BIOGRID database. Since the STRING database has a more standardized scoring system [44], we used this database to identify high-confidence interactions (>900) between NSP3/PLpro and human proteins. This interactome (unpublished data) has been used previously by our group in studies of SARS-CoV-2 replication proteins.

The central proteins for the biosystem (hub, bottleneck, and hub-bottleneck) were determined based on betweenness and degree centrality using the betweenness and degree functions of the igraph package [79]. Hubs are proteins with high degree centrality, bottlenecks are proteins with high betweenness, and hub-bottlenecks are hub proteins that are directly connected to a bottleneck protein. Cluster analysis was also performed in the R environment using the fastgreedy community function [79,80]. The focus was on the cluster of PLpro-related proteins. In this cluster, only the first and second levels of interactions of PLpro with central proteins were highlighted. For this group of proteins, functional enrichment analysis was performed to identify biological processes related to central proteins interacting with PLpro. For this purpose, the plugin BINGO from the Cytoscape program was used with the hypergeometric test applying the Benjamini & Hochberg false discovery rate (FDR) correction test for multiple comparisons [35,81].

The authors declare that all data that support the findings of this study are available within the paper files.

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

This work was supported by New York University Abu Dhabi through research funds to the lab of Prof. Rabeh in addition to the COVID-19 Facilitator Research Fund (grant number: ADC05).

Wael M. Rabeh: Conceptualization, Formal analysis, Supervision, Funding acquisition, Methodology, Writing — original draft, Project administration, Writing — review and editing. Cameron Lee Shetler: Data curation, Formal analysis, Validation, Writing — original draft. Juliana C. Ferreira: Data curation. Thyago H. S. Cardoso: Data curation, Software. Edson M.A. Silva: Software, Formal analysis. Nitin K. Saksena: Data curation, Writing — original draft.

This research was partially carried out using the Core Technology Platforms resources at New York University Abu Dhabi.

3CLpro

3-chymotrypsin-like cysteine protease

KD

binding affinity constant

ΔHcal

calorimetric enthalpy

CD

circular dichroism

COVID-19

coronavirus disease 2019

DSC

differential scanning calorimetry

λem

emission wavelength

λex

excitation wavelength

IC50

half maximal inhibitory concentration

Ki

inhibition constant

λmax

maximum tryptophan emission wavelength

Tm

melting temperature

MERS-CoV

Middle East respiratory syndrome coronavirus

nsps

non-structural proteins

PLpro

papain-like protease

ORFs

open reading frames

SARS-CoV

severe acute respiratory syndrome coronavirus

SARS-CoV-2

severe acute respiratory syndrome coronavirus 2

TFS

tryptophan fluorescence spectroscopy

1
WHO
. (
2022
)
World Health Organization Coronavirus (COVID-19) Dashboard
,
World Health Organization
,
Geneva, Switzerland
2
Zhou
,
P.
,
Yang
,
X.L.
,
Wang
,
X.G.
,
Hu
,
B.
,
Zhang
,
L.
,
Zhang
,
W.
et al (
2020
)
A pneumonia outbreak associated with a new coronavirus of probable bat origin
.
Nature
579
,
270
273
3
Zhu
,
N.
,
Zhang
,
D.
,
Wang
,
W.
,
Li
,
X.
,
Yang
,
B.
,
Song
,
J.
et al (
2020
)
A novel coronavirus from patients with pneumonia in China, 2019
.
N. Engl. J. Med.
382
,
727
733
4
Astuti
,
I.
and
Ysrafil,
(
2020
)
Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2): an overview of viral structure and host response
.
Diabetes Metab. Syndr.
14
,
407
412
5
Malik
,
A.
and
Alsenaidy
,
M.A.
(
2017
)
MERS-CoV papain-like protease (PLpro): expression, purification, and spectroscopic/thermodynamic characterization
.
3 Biotech.
7
,
100
6
V'Kovski
,
P.
,
Kratzel
,
A.
,
Steiner
,
S.
,
Stalder
,
H.
and
Thiel
,
V.
(
2021
)
Coronavirus biology and replication: implications for SARS-CoV-2
.
Nat. Rev. Microbiol.
19
,
155
170
7
Hu
,
B.
,
Guo
,
H.
,
Zhou
,
P.
and
Shi
,
Z.L.
(
2020
)
Characteristics of SARS-CoV-2 and COVID-19
.
Nat. Rev. Microbiol.
19
,
141
154
8
Kim
,
D.
,
Lee
,
J.-Y.
,
Yang
,
J.-S.
,
Kim
,
J.W.
,
Kim
,
V.N.
and
Chang
,
H.
(
2020
)
The architecture of SARS-CoV-2 transcriptome
.
Cell
181
,
914
921.e910
9
Klemm
,
T.
,
Ebert
,
G.
,
Calleja
,
D.J.
,
Allison
,
C.C.
,
Richardson
,
L.W.
,
Bernardini
,
J.P.
et al (
2020
)
Mechanism and inhibition of the papain-like protease, PLpro, of SARS-CoV-2
.
EMBO J.
39
,
e106275
10
Moustaqil
,
M.
,
Ollivier
,
E.
,
Chiu
,
H.P.
,
Van Tol
,
S.
,
Rudolffi-Soto
,
P.
,
Stevens
,
C.
et al (
2021
)
SARS-CoV-2 proteases PLpro and 3CLpro cleave IRF3 and critical modulators of inflammatory pathways (NLRP12 and TAB1): implications for disease presentation across species
.
Emerg. Microbes Infect.
10
,
178
195
11
Chen
,
X.
,
Yang
,
X.
,
Zheng
,
Y.
,
Yang
,
Y.
,
Xing
,
Y.
and
Chen
,
Z.
(
2014
)
SARS coronavirus papain-like protease inhibits the type I interferon signaling pathway through interaction with the STING-TRAF3-TBK1 complex
.
Protein Cell
5
,
369
381
12
Devaraj
,
S.G.
,
Wang
,
N.
,
Chen
,
Z.
,
Chen
,
Z.
,
Tseng
,
M.
,
Barretto
,
N.
et al (
2007
)
Regulation of IRF-3-dependent innate immunity by the papain-like protease domain of the severe acute respiratory syndrome coronavirus
.
J. Biol. Chem.
282
,
32208
32221
13
Yang
,
X.
,
Chen
,
X.
,
Bian
,
G.
,
Tu
,
J.
,
Xing
,
Y.
,
Wang
,
Y.
et al (
2014
)
Proteolytic processing, deubiquitinase and interferon antagonist activities of Middle East respiratory syndrome coronavirus papain-like protease
.
J. Gen. Virol.
95
,
614
626
14
Ratia
,
K.
,
Saikatendu
,
K.S.
,
Santarsiero
,
B.D.
,
Barretto
,
N.
,
Baker
,
S.C.
,
Stevens
,
R.C.
et al (
2006
)
Severe acute respiratory syndrome coronavirus papain-like protease: structure of a viral deubiquitinating enzyme
.
Proc. Natl Acad. Sci. U.S.A.
103
,
5717
5722
15
Ratia
,
K.
,
Pegan
,
S.
,
Takayama
,
J.
,
Sleeman
,
K.
,
Coughlin
,
M.
,
Baliji
,
S.
et al (
2008
)
A noncovalent class of papain-like protease/deubiquitinase inhibitors blocks SARS virus replication
.
Proc. Natl Acad. Sci. U.S.A.
105
,
16119
16124
16
Li
,
G.
and
De Clercq
,
E.
(
2020
)
Therapeutic options for the 2019 novel coronavirus (2019-nCoV)
.
Nat. Rev. Drug Discov.
19
,
149
150
17
Park
,
B.K.
,
Kim
,
J.
,
Park
,
S.
,
Kim
,
D.
,
Kim
,
M.
,
Baek
,
K.
et al (
2021
)
MERS-CoV and SARS-CoV-2 replication can be inhibited by targeting the interaction between the viral spike protein and the nucleocapsid protein
.
Theranostics
11
,
3853
3867
18
Shin
,
D.
,
Mukherjee
,
R.
,
Grewe
,
D.
,
Bojkova
,
D.
,
Baek
,
K.
,
Bhattacharya
,
A.
et al (
2020
)
Papain-like protease regulates SARS-CoV-2 viral spread and innate immunity
.
Nature
587
,
657
662
19
Baez-Santos
,
Y.M.
,
St John
,
S.E.
and
Mesecar
,
A.D.
(
2015
)
The SARS-coronavirus papain-like protease: structure, function and inhibition by designed antiviral compounds
.
Antivir. Res.
115
,
21
38
20
Gao
,
X.
,
Qin
,
B.
,
Chen
,
P.
,
Zhu
,
K.
,
Hou
,
P.
,
Wojdyla
,
J.A.
et al (
2020
)
Crystal structure of SARS-CoV-2 papain-like protease
.
Acta Pharm. Sin. B
11
,
237
245
21
Sargsyan
,
K.
,
Lin
,
C.-C.
,
Chen
,
T.
,
Grauffel
,
C.
,
Chen
,
Y.-P.
,
Yang
,
W.-Z.
et al (
2020
)
Multi-targeting of functional cysteines in multiple conserved SARS-CoV-2 domains by clinically safe Zn-ejectors
.
Chem. Sci.
11
,
9904
9909
22
Doboszewska
,
U.
,
Wlaź
,
P.
,
Nowak
,
G.
and
Młyniec
,
K.
(
2020
)
Targeting zinc metalloenzymes in coronavirus disease 2019
.
Br. J. Pharmacol.
177
,
4887
4898
23
Krishna
,
S.S.
,
Majumdar
,
I.
and
Grishin
,
N.V.
(
2003
)
Structural classification of zinc fingers: survey and summary
.
Nucleic Acids Res.
31
,
532
550
24
Barretto
,
N.
,
Jukneliene
,
D.
,
Ratia
,
K.
,
Chen
,
Z.
,
Mesecar
,
A.D.
and
Baker
,
S.C.
(
2005
)
The papain-like protease of severe acute respiratory syndrome coronavirus has deubiquitinating activity
.
J. Virol.
79
,
15189
15198
25
Bosken
,
Y.K.
,
Cholko
,
T.
,
Lou
,
Y.C.
,
Wu
,
K.P.
and
Chang
,
C.A.
(
2020
)
Insights into dynamics of inhibitor and ubiquitin-Like protein binding in SARS-CoV-2 papain-Like protease
.
Front. Mol. Biosci.
7
,
174
26
Ma
,
C.
,
Hu
,
Y.
,
Townsend
,
J.A.
,
Lagarias
,
P.I.
,
Marty
,
M.T.
,
Kolocouris
,
A.
et al (
2020
)
Ebselen, disulfiram, carmofur, PX-12, tideglusib, and shikonin are nonspecific promiscuous SARS-CoV-2 main protease inhibitors
.
ACS Pharmacol. Transl. Sci.
3
,
1265
1277
27
Lin
,
M.H.
,
Moses
,
D.C.
,
Hsieh
,
C.H.
,
Cheng
,
S.C.
,
Chen
,
Y.H.
,
Sun
,
C.Y.
et al (
2018
)
Disulfiram can inhibit MERS and SARS coronavirus papain-like proteases via different modes
.
Antiviral Res.
150
,
155
163
28
Han
,
Y.-S.
,
Chang
,
G.-G.
,
Juo
,
C.-G.
,
Lee
,
H.-J.
,
Yeh
,
S.-H.
,
Hsu
,
J.T.-A.
et al (
2005
)
Papain-like protease 2 (PLP2) from severe acute respiratory syndrome coronavirus (SARS-CoV): expression, purification, characterization, and inhibition
.
Biochemistry
44
,
10349
10359
29
Wang
,
L.
,
Hu
,
W.
and
Fan
,
C.
(
2020
)
Structural and biochemical characterization of SADS-CoV papain-like protease 2
.
Protein Sci.
29
,
1228
1241
30
Lindner
,
H.A.
,
Fotouhi-Ardakani
,
N.
,
Lytvyn
,
V.
,
Lachance
,
P.
,
Sulea
,
T.
and
Ménard
,
R.
(
2005
)
The papain-Like protease from the severe acute respiratory syndrome coronavirus Is a deubiquitinating enzyme
.
Virol. J.
79
,
15199
15208
31
Rut
,
W.
,
Lv
,
Z.
,
Zmudzinski
,
M.
,
Patchett
,
S.
,
Nayak
,
D.
,
Snipas
,
S.J.
et al (
2020
)
Activity profiling and crystal structures of inhibitor-bound SARS-CoV-2 papain-like protease: a framework for anti-COVID-19 drug design
.
Sci. Adv.
6
,
eabd4596
32
Chen
,
Z.
,
Cui
,
Q.
,
Cooper
,
L.
,
Zhang
,
P.
,
Lee
,
H.
,
Chen
,
Z.
et al (
2021
)
Ginkgolic acid and anacardic acid are specific covalent inhibitors of SARS-CoV-2 cysteine proteases
.
Cell Biosci.
11
,
45
33
Arya
,
R.
,
Prashar
,
V.
and
Kumar
,
M.
(
2022
)
Evaluating stability and activity of SARS-CoV-2 PLpro for high-throughput screening of inhibitors
.
Mol. Biotechnol.
64
,
1
8
34
Lin
,
M.-H.
,
Chuang
,
S.-J.
,
Chen
,
C.-C.
,
Cheng
,
S.-C.
,
Cheng
,
K.-W.
,
Lin
,
C.-H.
et al (
2014
)
Structural and functional characterization of MERS coronavirus papain-like protease
.
J. Biomed. Sci.
21
,
54
35
Shannon
,
P.
,
Markiel
,
A.
,
Ozier
,
O.
,
Baliga
,
N.S.
,
Wang
,
J.T.
,
Ramage
,
D.
et al (
2003
)
Cytoscape: a software environment for integrated models of biomolecular interaction networks
.
Genome Res.
13
,
2498
2504
36
Yu
,
H.
,
Kim
,
P.M.
,
Sprecher
,
E.
,
Trifonov
,
V.
and
Gerstein
,
M.
(
2007
)
The importance of bottlenecks in protein networks: correlation with gene essentiality and expression dynamics
.
PLoS Comput. Biol.
3
,
e59
37
Hu
,
L.
,
Wang
,
X.
,
Huang
,
Y.A.
,
Hu
,
P.
and
You
,
Z.H.
(
2021
)
A survey on computational models for predicting protein-protein interactions
.
Brief. Bioinform.
22
,
bbab036
38
Auwul
,
M.R.
,
Rahman
,
M.R.
,
Gov
,
E.
,
Shahjaman
,
M.
and
Moni
,
M.A.
(
2021
)
Bioinformatics and machine learning approach identifies potential drug targets and pathways in COVID-19
.
Brief. Bioinform.
22
,
bbab120
39
Gupta
,
P.
and
Mohanty
,
D.
(
2021
)
SMMPPI: a machine learning-based approach for prediction of modulators of protein-protein interactions and its application for identification of novel inhibitors for RBD:hACE2 interactions in SARS-CoV-2
.
Brief. Bioinform.
22
,
bbab111
40
Lian
,
X.
,
Yang
,
X.
,
Yang
,
S.
and
Zhang
,
Z.
(
2021
)
Current status and future perspectives of computational studies on human-virus protein-protein interactions
.
Brief. Bioinform.
22
,
bbab029
41
Ako-Adjei
,
D.
,
Fu
,
W.
,
Wallin
,
C.
,
Katz
,
K.S.
,
Song
,
G.
,
Darji
,
D.
et al (
2015
)
HIV-1, human interaction database: current status and new features
.
Nucleic Acids Res.
43
,
D566
D570
42
Kwofie
,
S.K.
,
Schaefer
,
U.
,
Sundararajan
,
V.S.
,
Bajic
,
V.B.
and
Christoffels
,
A.
(
2011
)
HCVpro: hepatitis C virus protein interaction database
.
Infect. Genet. Evol.
11
,
1971
1977
43
Ghosh
,
N.
,
Saha
,
I.
and
Sharma
,
N.
(
2021
)
Interactome of human and SARS-CoV-2 proteins to identify human hub proteins associated with comorbidities
.
Comput. Biol. Med.
138
,
104889
44
Szklarczyk
,
D.
,
Gable
,
A.L.
,
Lyon
,
D.
,
Junge
,
A.
,
Wyder
,
S.
,
Huerta-Cepas
,
J.
et al (
2019
)
STRING v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets
.
Nucleic Acids Res.
47
,
D607
D613
45
Szklarczyk
,
D.
,
Gable
,
A.L.
,
Nastou
,
K.C.
,
Lyon
,
D.
,
Kirsch
,
R.
,
Pyysalo
,
S.
et al (
2021
)
The STRING database in 2021: customizable protein-protein networks, and functional characterization of user-uploaded gene/measurement sets
.
Nucleic Acids Res.
49
,
D605
d612
46
Juttukonda
,
L. J.
, and
Skaar
,
E.P.
(
2017
) Manganese and Nutritional Immunity. In
Molecular, Genetic, and Nutritional Aspects of Major and Trace Minerals
(
Collins
,
J.
, ed.), pp.
377
387
,
Academic Press
,
Gainesville, USA
47
Lim
,
K.P.
,
Ng
,
L.F.
and
Liu
,
D.X.
(
2000
)
Identification of a novel cleavage activity of the first papain-like proteinase domain encoded by open Reading frame 1a of the coronavirus avian infectious bronchitis virus and characterization of the cleavage products
.
J. Virol.
74
,
1674
1685
48
Hilgenfeld
,
R.
(
2014
)
From SARS to MERS: crystallographic studies on coronaviral proteases enable antiviral drug design
.
FEBS J.
281
,
4085
4096
49
Békés
,
M.
,
Rut
,
W.
,
Kasperkiewicz
,
P.
,
Mulder
,
M.P.
,
Ovaa
,
H.
,
Drag
,
M.
et al (
2015
)
SARS hcov papain-like protease is a unique Lys48 linkage-specific di-distributive deubiquitinating enzyme
.
Biochem. J.
468
,
215
226
50
Békés
,
M.
,
van der Heden van Noort
,
G.J.
,
Ekkebus
,
R.
,
Ovaa
,
H.
,
Huang
,
T.T.
and
Lima
,
C.D.
(
2016
)
Recognition of Lys48-linked di-ubiquitin and deubiquitinating activities of the SARS coronavirus papain-like protease
.
Mol. Cell
62
,
572
585
51
Lindner
,
H.A.
,
Lytvyn
,
V.
,
Qi
,
H.
,
Lachance
,
P.
,
Ziomek
,
E.
and
Ménard
,
R.
(
2007
)
Selectivity in ISG15 and ubiquitin recognition by the SARS coronavirus papain-like protease
.
Arch. Biochem. Biophys.
466
,
8
14
52
Clementz
,
M.A.
,
Chen
,
Z.
,
Banach
,
B.S.
,
Wang
,
Y.
,
Sun
,
L.
,
Ratia
,
K.
et al (
2010
)
Deubiquitinating and interferon antagonism activities of coronavirus papain-like proteases
.
J. Virol.
84
,
4619
4629
53
Báez-Santos
,
Y.M.
,
Barraza
,
S.J.
,
Wilson
,
M.W.
,
Agius
,
M.P.
,
Mielech
,
A.M.
,
Davis
,
N.M.
et al (
2014
)
X-ray structural and biological evaluation of a series of potent and highly selective inhibitors of human coronavirus papain-like proteases
.
J. Med. Chem.
57
,
2393
2412
54
Lee
,
H.
,
Lei
,
H.
,
Santarsiero
,
B.D.
,
Gatuz
,
J.L.
,
Cao
,
S.
,
Rice
,
A.J.
, et al (
2015
)
Inhibitor recognition specificity of MERS-CoV papain-like protease may differ from that of SARS-CoV
.
ACS Chem. Biol.
10
,
1456
1465
55
Kumar
,
A.
,
Kubota
,
Y.
,
Chernov
,
M.
and
Kasuya
,
H.
(
2020
)
Potential role of zinc supplementation in prophylaxis and treatment of COVID-19
.
Med. Hypotheses
144
,
109848
56
Jothimani
,
D.
,
Kailasam
,
E.
,
Danielraj
,
S.
,
Nallathambi
,
B.
,
Ramachandran
,
H.
,
Sekar
,
P.
et al (
2020
)
COVID-19: poor outcomes in patients with zinc deficiency
.
Int. J. Infect. Dis.
100
,
343
349
57
Alexander
,
J.
,
Tinkov
,
A.
,
Strand
,
T.A.
,
Alehagen
,
U.
,
Skalny
,
A.
and
Aaseth
,
J.
(
2020
)
Early nutritional interventions with zinc, Selenium and vitamin D for raising anti-viral resistance against progressive COVID-19
.
Nutrients
12
,
2358
58
Barocas
,
J.A.
,
So-Armah
,
K.
,
Cheng
,
D.M.
,
Lioznov
,
D.
,
Baum
,
M.
,
Gallagher
,
K.
et al (
2019
)
Zinc deficiency and advanced liver fibrosis among HIV and hepatitis C co-infected anti-retroviral naïve persons with alcohol use in Russia
.
PLoS ONE
14
,
e0218852
59
Skalny
,
A.V.
,
Rink
,
L.
,
Ajsuvakova
,
O.P.
,
Aschner
,
M.
,
Gritsenko
,
V.A.
,
Alekseenko
,
S.I.
et al (
2020
)
Zinc and respiratory tract infections: perspectives for COVID-19 (Review)
.
Int. J. Mol. Med.
46
,
17
26
60
te Velthuis
,
A.J.W.
,
van den Worm
,
S.H.E.
,
Sims
,
A.C.
,
Baric
,
R.S.
,
Snijder
,
E.J.
and
van Hemert
,
M.J.
(
2010
)
Zn2+ inhibits coronavirus and arterivirus RNA polymerase activity in vitro and zinc ionophores block the replication of these viruses in cell culture
.
PLoS Pathog.
6
,
e1001176
61
Harding
,
M.M.
,
Nowicki
,
M.W.
and
Walkinshaw
,
M.D.
(
2010
)
Metals in protein structures: a review of their principal features
.
Crystallogr. Rev.
16
,
247
302
62
Hamada
,
A.M.
(
2020
)
Vitamins, omega-3, magnesium, manganese, and thyme can boost our immunity and protect against COVID-19
.
Eur. J. Biol. Res.
10
,
271
295
63
Vivian
,
J.T.
and
Callis
,
P.R.
(
2001
)
Mechanisms of tryptophan fluorescence shifts in proteins
.
Biophys. J.
80
,
2093
2109
64
Munishkina
,
L.A.
and
Fink
,
A.L.
(
2007
)
Fluorescence as a method to reveal structures and membrane-interactions of amyloidogenic proteins
.
Biochim. Biophys. Acta
1768
,
1862
1885
65
Sindrewicz
,
P.
,
Li
,
X.
,
Yates
,
E.A.
,
Turnbull
,
J.E.
,
Lian
,
L.-Y.
and
Yu
,
L.-G.
(
2019
)
Intrinsic tryptophan fluorescence spectroscopy reliably determines galectin-ligand interactions
.
Sci. Rep.
9
,
11851
66
Arntfield
,
S.D.
and
Murray
,
E.D.
(
1981
)
The influence of processing parameters on food protein functionality I. Differential scanning calorimetry as an indicator of protein denaturation
.
Can. Inst. Food Technol. J.
14
,
289
294
67
Nawaz
,
M.H.
,
Ferreira
,
J.C.
,
Nedyalkova
,
L.
,
Zhu
,
H.
,
Carrasco-Lopez
,
C.
,
Kirmizialtin
,
S.
et al (
2018
)
The catalytic inactivation of the N-half of human hexokinase 2 and structural and biochemical characterization of its mitochondrial conformation
.
Biosci. Rep.
38
,
BSR20171666
68
Lindquist
,
S.
(
1986
)
The heat-shock response
.
Annu. Rev. Biochem.
55
,
1151
1191
69
Truong-Tran
,
A.Q.
,
Carter
,
J.
,
Ruffin
,
R.E.
and
Zalewski
,
P.D.
(
2001
)
The role of zinc in caspase activation and apoptotic cell death
.
Biometals
14
,
315
330
70
Xu
,
B.
,
Lei
,
Y.
,
Ren
,
X.
,
Yin
,
F.
,
Wu
,
W.
,
Sun
,
Y.
et al (
2021
)
SOD1 is a possible predictor of COVID-19 progression as revealed by plasma proteomics
.
ACS Omega
6
,
16826
16836
71
Guimarães
,
L.M.F.
,
Rossini
,
C.V.T.
and
Lameu
,
C.
(
2021
)
Implications of SARS-Cov-2 infection on eNOS and iNOS activity: consequences for the respiratory and vascular systems
.
Nitric Oxide
111–112
,
64
71
72
Cortese-Krott
,
M.M.
,
Kulakov
,
L.
,
Opländer
,
C.
,
Kolb-Bachofen
,
V.
,
Kröncke
,
K.D.
and
Suschek
,
C.V.
(
2014
)
Zinc regulates iNOS-derived nitric oxide formation in endothelial cells
.
Redox Biol.
2
,
945
954
73
Kobayashi
,
J.
(
2021
)
Lifestyle-mediated nitric oxide boost to prevent SARS-CoV-2 infection: a perspective
.
Nitric Oxide
115
,
55
61
74
Tao
,
X.
,
Zhang
,
L.
,
Du
,
L.
,
Lu
,
K.
,
Zhao
,
Z.
,
Xie
,
Y.
et al (
2022
)
Inhibition of SARS-CoV-2 replication by zinc gluconate in combination with hinokitiol
.
J. Inorg. Biochem.
231
,
111777
75
Kladnik
,
J.
,
Dolinar
,
A.
,
Kljun
,
J.
,
Perea
,
D.
,
Grau-Expósito
,
J.
,
Genescà
,
M.
et al (
2022
)
Zinc pyrithione is a potent inhibitor of PLPro and cathepsin L enzymes with ex vivo inhibition of SARS-CoV-2 entry and replication
.
J. Enzyme Inhib. Med. Chem.
37
,
2158
2168
76
Read
,
S.A.
,
Obeid
,
S.
,
Ahlenstiel
,
C.
and
Ahlenstiel
,
G.
(
2019
)
The role of zinc in antiviral immunity
.
Adv. Nutr.
10
,
696
710
77
Rani
,
I.
,
Goyal
,
A.
,
Bhatnagar
,
M.
,
Manhas
,
S.
,
Goel
,
P.
,
Pal
,
A.
et al (
2021
)
Potential molecular mechanisms of zinc- and copper-mediated antiviral activity on COVID-19
.
Nutr. Res.
92
,
109
128
78
Ferreira
,
J.C.
,
Fadl
,
S.
,
Ilter
,
M.
,
Pekel
,
H.
,
Rezgui
,
R.
,
Sensoy
,
O.
et al (
2021
)
Dimethyl sulfoxide reduces the stability but enhances catalytic activity of the main SARS-CoV-2 protease 3CLpro
.
FASEB J.
35
,
e21774
79
Csárdi
,
G.
and
Nepusz
,
T.
(
2006
)
The igraph software package for complex network research
.
Int. J. Complex Syst.
1695
https://igraph.org
80
Clauset
,
A.
,
Newman
,
M.E.J.
and
Moore
,
C.
(
2004
)
Finding community structure in very large networks
.
Phys. Rev. E
70
,
066111
81
Maere
,
S.
,
Heymans
,
K.
and
Kuiper
,
M.
(
2005
)
BiNGO: a Cytoscape plugin to assess overrepresentation of Gene Ontology categories in Biological Networks
.
Bioinformatics
21
,
3448
3449
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-NC-ND).