Next Article in Journal
Systematic Development of Vanadium Catalysts for Sustainable Epoxidation of Small Alkenes and Allylic Alcohols
Next Article in Special Issue
Synthesis, Structural Characterization, Cytotoxicity, and Protein/DNA Binding Properties of Pyridoxylidene-Aminoguanidine-Metal (Fe, Co, Zn, Cu) Complexes
Previous Article in Journal
The Reduction in the Mitochondrial Membrane Potential in Aging: The Role of the Mitochondrial Permeability Transition Pore
Previous Article in Special Issue
The Effect of Metal Ions (Fe, Co, Ni, and Cu) on the Molecular-Structural, Protein Binding, and Cytotoxic Properties of Metal Pyridoxal-Thiosemicarbazone Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Lung Microbiome in COPD and Lung Cancer: Exploring the Potential of Metal-Based Drugs

by
Megan O’Shaughnessy
1,*,
Orla Sheils
1,2 and
Anne-Marie Baird
1
1
School of Medicine, Trinity Translational Medicine Institute, Trinity College Dublin, D08 W9RT Dublin, Ireland
2
Department of Histopathology and Morbid Anatomy, Trinity Translational Medicine Institute, St. James’s Hospital, D08 RX0X Dublin, Ireland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(15), 12296; https://doi.org/10.3390/ijms241512296
Submission received: 30 June 2023 / Revised: 28 July 2023 / Accepted: 29 July 2023 / Published: 1 August 2023
(This article belongs to the Special Issue Metal-Based Complexes in Cancer 2.0)

Abstract

:
Chronic obstructive pulmonary disease (COPD) and lung cancer 17 are two of the most prevalent and debilitating respiratory diseases worldwide, both associated with high morbidity and mortality rates. As major global health concerns, they impose a substantial burden on patients, healthcare systems, and society at large. Despite their distinct aetiologies, lung cancer and COPD share common risk factors, clinical features, and pathological pathways, which have spurred increasing research interest in their co-occurrence. One area of particular interest is the role of the lung microbiome in the development and progression of these diseases, including the transition from COPD to lung cancer. Exploring novel therapeutic strategies, such as metal-based drugs, offers a potential avenue for targeting the microbiome in these diseases to improve patient outcomes. This review aims to provide an overview of the current understanding of the lung microbiome, with a particular emphasis on COPD and lung cancer, and to discuss the potential of metal-based drugs as a therapeutic strategy for these conditions, specifically concerning targeting the microbiome.

Graphical Abstract

1. Introduction

The human microbiome has gained significant attention due to its influential role in health and disease [1,2]. While there has been extensive research on gut microbiota, the microbiome of the human respiratory tract has received less attention, mainly because the lungs were historically considered sterile [3]. The advent of modern culture-independent methods and next-generation sequencing (NGS) techniques has heralded a paradigm shift in our comprehension of the microbial landscape within the lung [4,5]. This realisation has sparked a burgeoning field of research, unveiling fascinating insights into the complex interplay between the lung microbiome, the immune system, and host physiology [6,7]. The lung microbiome exhibits remarkable plasticity, dynamically shifting in response to a myriad of factors, including age, environmental exposure, lifestyle choices, medications, and genetic predisposition [4,7,8,9]. Evidence from human studies has shown that alterations in the lung microbiome have been implicated in several pathological pulmonary conditions, including chronic obstructive pulmonary disease (COPD) and lung cancer [10,11,12].
In 2019, 212.3 million cases of COPD were reported globally, with COPD accounting for 3.3 million deaths [13]. The World Health Organization (WHO) predicted that COPD would be the third leading cause of death by 2030. However, it has already reached that landmark and is now expected to be the most lethal respiratory disease within the same time frame [14]. Lung cancer is the second most commonly diagnosed cancer and is responsible for more cancer-related deaths globally than any other cancer type [15]. The overall five-year survival rate is approximately 20%, which is partly due to the majority of people presenting with advanced disease. There is a high incidence of lung cancer among people with COPD, and this association transcends the established connection that both conditions have with a smoking history. For instance, several studies have reported a heightened risk of lung cancer in people with COPD, independent of age or tobacco use [16,17,18,19,20,21,22,23]. The data suggest that COPD in those who smoke have a two to five-fold higher risk of developing lung cancer compared to non-smokers, with the overall survival rate in those with COPD and lung cancer significantly lower than those with lung cancer without COPD [17,24,25]. Park et al. [23] undertook a large national cohort study in South Korea that investigated the incidence rate of lung cancer in never-smokers with COPD. In comparison to never-smokers without COPD, the hazard ratios for lung cancer in never-smokers with COPD, ever-smokers without COPD, and ever-smokers with COPD were 2.67, 1.97, and 6.19, respectively, indicating that COPD is a strong independent risk factor of lung cancer, irrespective of smoking status [23]. Reported mechanisms linking COPD and lung cancer include genetic factors, epigenetic changes, non-coding RNAs, oxidative stress, immune environment, and sex differences [17,26,27,28,29]. However, emerging evidence also suggests that dysregulation in the lung microbiome may play a key role in the pathogenesis of COPD and the transition to lung cancer [7,12,30]. In relation to the microbiome, commonly reported mechanisms are the chronic inflammatory environment in COPD, fuelled by an imbalanced lung microbiota, which can create a pro-oncogenic environment promoting lung cancer through oxidative stress, genetic mutation, and DNA damage [26]. Dysregulated lung microbiota can also modulate the immune response, with certain bacteria inducing pro-inflammatory cytokines, potentially fostering lung cancer development [31]. The precise pathogenic mechanisms through which the microbiome mediates the progression of lung cancer remain largely elusive, although they are considered multifaceted and involve bacterial toxins such as lipopolysaccharide (LPS) and the release of inflammatory cytokines by immune cells [32]. For instance, the LPS of H. pylori can trigger the production of pro-inflammatory agents, including tumour necrosis factor (TNF), interleukin (IL)-1, and IL-6 [33]. These inflammation-promoting mediators are implicated in the progression of chronic lung diseases such as COPD as a precursor to lung cancer [27]. Understanding the role of the lung microbiome in the pathogenesis and progression of COPD to lung cancer may also open new avenues for therapeutic interventions, such as metal-based drugs.
The use of metals and metal-based drugs in medicine has a longstanding history, with some of the earliest examples dating back to ancient civilisations [34,35]. The potential of metal-based drugs to exhibit unique biochemical properties, largely due to the diverse redox states and coordination geometries of metal ions, allows for a wide range of potential applications, including antimicrobial agents [36,37,38,39,40], cancer therapeutics [41,42,43], and modulators of the microbiome [44,45,46]. Moreover, their ability to engage in multiple simultaneous interactions with biomolecules increases their therapeutic effectiveness and versatility [36,47]. Recent research has shown that metal-based drugs can manipulate microbial populations, both in vitro and in vivo, by selectively inhibiting growth or suppressing virulence in a multi-modal fashion [37,38,48]. For instance, gold-based drugs have demonstrated antimicrobial effects and can modulate gut microbiota [49,50,51]. In the context of the lung, the use of metal-based drugs as microbiome modulators is still in its infancy, but could offer a new avenue in treating respiratory diseases where chronic infections and dysregulation of the lung microbiome play a critical role.

2. The Healthy Lung Microbiome

The concept of the lung microbiome has been radically altered with the understanding that the human lung hosts a complex ecosystem comprised of bacteria, fungi, and viruses, which actively participate in maintaining respiratory health by contributing to immune modulation, pathogen displacement, and metabolic contributions [52,53]. The composition of the respiratory microbiome is transient and determined by continuous microbial immigration (through microaspiration, inhalation, and direct mucosal spread), elimination (by the immune system and mucociliary clearance), and replication [3]. The notion of a healthy lung microbiome refers to a state in which a multitude of beneficial microorganisms coexist in harmony, promoting an immune environment that is neither too reactive nor too lax, providing robustness against invading pathogens, and supporting the crucial function of the lungs [5,54]. Under homeostatic conditions, the predominant bacterial phyla in the human lung are Bacteroidota (45–50%, with the genus Prevotella), Bacillota (30–35%, including the genera Streptococcus and Veillonella), Pseudomonadota (10–15%, with the genera Haemophilus and Neisseria), Actinomycetota (5%, represented by the genus Corynebacterium), and Fusobacteriota (5%, with the genus Fusobacterium) [7,55]. The diversity of the microbiome, which is a significant factor in maintaining respiratory health, is reflected by its richness (total number of different microbial species) and its evenness (relative abundance or proportion of different species) [3,56]. In a diseased state, alterations in the structure and local microenvironment of the lung, such as changes in mucosal pH, oxygen gradients, nutrient availability, inflammation, and host defence, foster proliferation of pathogenic microbes and consequently lead to shifts in the microbiome composition [5,57,58]. These disruptions, known as dysbiosis, can have detrimental effects on lung health. Dysbiosis can result from various factors, including environmental influences (exposure to harmful pollutants or smoking), antibiotic use, respiratory infections, lifestyle, or genetics, and has been associated with several pulmonary diseases, including COPD and lung cancer [7,58]. Understanding the factors that contribute to dysbiosis and developing strategies to restore a healthy lung microbiome through targeted approaches could offer novel therapeutic avenues for these conditions.

3. The Lung Microbiome in Lung Disease

Numerous studies have distinguished the lung microbiome in human health and disease, in which a shift of the microbiome is associated with diseases and key clinical parameters, such as severity, exacerbation, phenotype, endotype, inflammation, and mortality [57,59,60,61]. Dysbiosis, characterised by a decrease in microbial diversity and a shift in community composition, is observed in various respiratory disorders, including asthma, cystic fibrosis, idiopathic pulmonary fibrosis (IPF), tuberculosis, COPD, and lung cancer [62,63,64,65,66,67]. Diversity is a measure of the evenness and richness of a microbial community, which can be measured within a biological sample (α-diversity) or between samples (β-diversity) [68,69]. Lower bacterial diversity has been linked to disease progression, although it is unknown whether microbial dysbiosis is a cause or effect of the disease [2,7,70]. Dysbiosis may have a causative role in lung diseases by upregulating inflammatory signals (such as nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), Ras, IL-17, and phosphoinositide 3-kinase (PI3K)) [67,71,72,73,74] or by suppressing the production of TNF and interferon-gamma (IFNγ) in response to pathogen presence in the lower respiratory tract [58,75]. The following sections will focus specifically on the lung microbiome in COPD and lung cancer.

3.1. Chronic Obstructive Pulmonary Disease (COPD)

COPD is a chronic inflammatory lung condition characterised by persistent respiratory symptoms and progressive airflow restriction. Predominantly triggered by long-term exposure to harmful pollutants, such as cigarette smoke and environmental toxins, it clinically presents with dyspnoea, chronic cough, sputum production, and wheezing [76,77]. The disease comprises two primary phenotypes: chronic bronchitis, hallmarked by a chronic productive cough, and emphysema, typified by alveolar wall destruction over time. COPD, as a progressive and potentially fatal condition, can significantly deteriorate quality of life due to recurrent exacerbations and a decline in pulmonary function [77,78,79]. Although there is no definitive cure at present, the management of symptoms and decelerating disease progression are pivotal. Recent research has begun to shed light on the role of the lung microbiome in the pathogenesis and progression of COPD, offering a new perspective on this debilitating disease [80,81,82]. Gram-negative pathogenic bacteria tied to COPD (Haemophilus spp., Moraxella, Pseudomonas) possess a notably higher potential to stimulate an immune response compared to Gram-negative commensal bacteria (Prevotella spp.) [83]. Poor oral hygiene has been identified as a risk factor for inflammatory lung conditions through microaspiration of oral commensals, such as Veillonella and Prevotella, which have been associated with increased TH17 lymphocytes within the lung [71,84]. Several studies have reported the composition of the airway microbiome in COPD and found a shift in microbiome diversity with a decrease in Bacillota (Firmicutes) and Bacteroidota and an increase in Pseudomonadota, particularly the genus Haemophilus, which positively correlated with IL-8 present in sputum [8,80,85,86,87,88]. This shift in the microbiome composition, particularly the elevation in Pseudomonadota, has been associated with greater emphysema, and increased immune cell infiltration leading to chronic inflammation, airway remodelling, and exacerbations [88,89,90,91,92]. Exacerbations, defined as acute worsening of respiratory symptoms, significantly contribute to the morbidity and mortality associated with COPD. These episodes are often triggered by bacterial or viral infections and treated with antibiotics and corticosteroids, critical elements in the standard therapeutic approach [93]. Antibiotic-mediated perturbation of the gut microbiome has been widely reported to be associated with numerous infectious and autoimmune diseases of the gastrointestinal tract [94,95,96,97]. Although less extensively studied, antibiotic use has been reported to cause alterations in the lung microbiome, which can negatively impact the ecological balance of microbial communities within the lung and potentially escalate disease progression and exacerbation severity [56,98]. Moreover, dysbiosis may potentiate bacterial resistance, creating challenges for future antimicrobial treatment [99,100]. Thus, while antibiotics and corticosteroids are essential for managing COPD exacerbations, their impact on the lung microbiome warrants careful consideration within therapeutic strategies [101].
The lung microbiome in COPD has been found to be significantly associated with bacterial biomass, lymphocyte proportion, TH17 immune response, exacerbation frequency, and resistance to antimicrobial therapy [102]. Initial investigations into the microbiome of patients with stable COPD have demonstrated a significant correlation between the presence of pathogenic bacteria, such as Haemophilus influenzae, Streptococcus pneumoniae, Moraxella catarrhalis, Staphylococcus aureus, Pseudomonas aeruginosa, and Enterobacterales. As COPD progresses, chronic inflammation impairs the innate immune response within the lung, which in turn creates a favourable environment for an increase in bacterial burden. Numerous studies have shown that in moderate to severe COPD (Global Initiative for COPD (GOLD) 2–4), there is an enrichment of Gammaproteobacteria (Haemophilus and Moraxella spp.) in bronchiole lavage and lung tissue samples [80,103,104]. Erb-Downward et al. [80] found that the microbiomes of patients with moderate or severe COPD had lower bacterial diversity scores than healthy smokers and non-smokers. They identified a core COPD lung microbiome that included Pseudomonas, Streptococcus, Prevotella, Fusobacterium, Haemophilus, Veillonella, and Porphyromonas [80]. Interestingly, in patients with very advanced COPD, there were significant differences in the microbiomes at adjacent lung sites, suggesting heterogeneity within the individual lung microbiome. In a similar vein, Sze et al. [105] evaluated the microbiomes in lung tissues taken from patients with severe COPD (GOLD 4) at the time of lung transplantation and found an increase in bacterial diversity in patients with severe COPD compared to non-smokers, smokers, and patients with CF, with a notable increase in the phylum Bacillota (Firmicutes), specifically Lactobacillus [105]. These findings highlight the dynamic nature of the lung microbiome in COPD patients and the differences across sample types. In a longitudinal observational study analysing the microbiome of sputum of clinically stable COPD patients, reduced microbiome diversity was observed associated with Pseudomonadota (predominantly Haemophilus) dominance, which was associated with neutrophil-associated protein profiles and an increased risk of mortality [89]. In a recent cohort study, longitudinal sputum samples were taken from COPD patients during acute exacerbation (AECOPD), which found significant positive correlations between the abundance of Pseudomonas and TNF, the abundance of Klebsiella, and the percentage of eosinophils. Furthermore, the study identified four clusters of COPD based on the respiratory microbiome, with the AECOPD-related cluster characterised by the enrichment of Pseudomonas and Haemophilus and a high level of TNF [88]. Therefore, these patients might benefit from targeted antibacterial agents, which may aid in alleviating inflammation. In contrast, patients with a diverse microbiome profile, including Veillonella and Prevotella, exhibited a more dynamic microbiome over time and showed elevations of IL-17A within sputum and serum [106]. It was observed that these patients had greater microbiome shifts during exacerbations and, therefore, would profit from an anti-inflammatory therapeutic strategy. This growing body of evidence underscores the intricate relationship between the lung microbiome and COPD, shedding light on the potential of microbiota-targeted interventions to improve the long-term prognosis of COPD. However, a comprehensive understanding of these relationships warrants further in-depth investigations, including longitudinal studies to track microbiome changes over time and interventional studies to test the causality of observed associations.

3.2. Lung Cancer

Lung cancer, a heterogeneous group of malignancies arising in the lung parenchyma or bronchi, is the leading cause of cancer-related deaths globally [107]. The disease is classified into non-small cell lung cancer (NSCLC), which constitutes about 85% of cases, and small cell lung cancer (SCLC), which represents the remaining 15%. The three main subtypes of NSCLC are adenocarcinoma (40%), squamous cell carcinoma (25–30%), and large cell carcinoma (5–10%). Lung cancer is typically asymptomatic in the early stages, with clinical manifestations such as persistent cough, chest pain, and haemoptysis emerging as the disease progresses [108,109,110]. The predominant contributors to lung cancer include exposure to tobacco smoke, environmental toxins, carcinogens, persistent airway inflammation instigated by pathogenic infections, as well as fibrosis and scarring resulting from co-existing lung diseases [29,111]. Despite advancements in therapeutic interventions, lung cancer continues to be a primary cause of death related to cancer worldwide due to its late-stage diagnosis and high recurrence rates, even in those with early-stage disease. Comprehensive strategies for early detection and treatment are crucial to reducing the global burden of the disease.
The intricate relationship between the lung microbiome and lung cancer has begun to be elucidated. Research has shown associations between specific bacterial phyla, family, genera, and species and the progression of lung cancer. Among these are Actinomycetota, Bacteroidota, Pseudomonadota, Bacillota (Firmicutes) [112], Capnocytophaga, Neisseria, and Selenomonas [113], Thermus, Legionella, Megasphaera, Veillonella [114], Cyanobacteria [115], Acidovorax temporans [33] and Helicobacter pylori [116]. Dysbiosis of the microbiome is mainly manifested by the decrease in symbiotic bacteria and the increase in pathogenic bacteria, and then inducement of carcinogenesis at multiple levels, including metabolism alteration, inflammation, and altered immune response [117,118]. Several studies have pointed towards an increased abundance of certain genera, such as Veillonella, Streptococcus, and Prevotella, in lung cancer patients compared to healthy individuals, which have also been observed in patients with COPD [119,120]. For example, a study by Yu et al. [121] reported that the microbiota of lung tumour tissue showed significantly lower α-diversity compared to non-malignant lung tissue samples, with higher levels of Veillonella. Moreover, it was reported that bacterial composition correlated with cancer stage, with a higher abundance of Thurmus in advanced-stage (IIIB, IV) patients, and Legionella in patients with metastases [121]. Studies have also indicated an enrichment of Veillonella and Megasphaera in the bronchoalveolar lavage fluid (BALF) of patients diagnosed with NSCLC stages II-IV (72% adenocarcinoma and 28% squamous cell carcinoma), suggesting these genera could serve as biomarkers to predict disease progression [113]. Complementary findings by Huang et al. [122] support this observation in the bronchial washing fluid of patients with lung adenocarcinoma compared to those with squamous cell lung carcinoma. Gomes et al. [114] reported that lung cancer microbiota was enriched in Pseudomonadota and more diverse in squamous cell carcinoma than adenocarcinoma, particularly in males and heavier smokers, suggesting a potential link between these bacteria and the presence of other risk factors. Najafi et al. [123] found that the relative abundance of several bacterial taxa, including Actinomycetota phylum, Corynebacteriaceae, and Halomonadaceae families, and Corynebacterium, Lachnoanaerobaculum, and Halomonas genera, is significantly decreased in lung tumour tissues of lung cancer patients in comparison with matched tumour-adjacent normal tissues. The microbiota within the airway plays a distinct role in the onset and progression of lung cancer. Research has indicated that mice, either germ-free or treated with antibiotics, show considerable resistance against lung cancer development, even with Kras mutation and p53 loss [124]. Other studies have identified differences in the gut microbiota between patients with lung cancer and healthy controls, with an increased abundance of Enterococcus and decreased levels of the phylum Actinomycetota and genus Bifidobacterium [125]. The “gut-lung axis” is an emerging concept linking the state of the gut microbiota to respiratory health outcomes [126,127]. Simultaneously, it was observed that depleting the microbiota or inhibiting γδ T cells or their downstream effector molecules all effectively suppressed the growth of lung adenocarcinoma in a genetically engineered mouse model driven by an activating point mutation of Kras and loss of p53 [128]. The role of the resident microbiome in the progression of lung cancer was further examined in a nested case control study encompassing 4336 lung cancer subjects and 10,000 matched controls aged 40–84 years [129]. The study sought to establish any correlation between antibiotic usage and the risk of lung cancer. Remarkably, subjects who received 10 or more antibiotic courses presented a relative lung cancer risk of 2.52 (95% CI, 2.25–2.83) compared with controls who had not received antibiotics. The elevated relative risk could potentially be attributed to the inflammatory conditions induced by frequent infections and consequent alterations in the lung microbiomes among those administered antibiotics [129].
The proposed theory is that dysbiosis or microbial imbalance might propel carcinogenesis through three channels: (i) disruption of immune equilibrium, (ii) chronic inflammation instigation, and (iii) activation of cancer-causing pathways [130,131,132,133,134]. Firstly, dysbiosis has the potential to disrupt the lung immune system’s fundamental stimulation, and depletion of microbial diversity impairs the initial activation of antigen-presenting cells, thereby inhibiting their response to tumour antigens [2]. Conversely, bacterial overgrowth can lead to an overstimulation of the immune system and unchecked proliferation of IL-17-producing CD4+ helper T (TH17) cells, mediators in lung tumorigenesis [135]. Secondly, dysbiosis incites chronic inflammation via the release of DNA-damaging metabolites and genotoxins from commensal organisms. Inflammatory cells activated by dysbiosis can also release reactive oxygen (ROS) and nitrogen (RNS) species, promoting carcinogenesis and angiogenesis [136,137]. Lastly, several studies have highlighted that some species within the microbiota can directly stimulate cancer-causing pathways. For instance, Apopa et al. [115] demonstrated an increase in PARP1 in NSCLC tissues in the presence of the cyanobacteria toxin microcystin. Likewise, research by Tsay et al. [67] connected Streptococcus and Veillonella to the stimulation of the PI3K and extracellular signal-regulated protein kinase (ERK) pathways involved in the disease. Ochoa et al. [138] found that exposure of the airway to smoke particulates and nontypeable H. influenzae (NTHi) promoted lung cancer cell proliferation by release of IL-6 and TNF, which further activated the STAT3 and NF-κB pathways in the airway epithelium. Interestingly, it was demonstrated that IL-6 blockade significantly inhibited lung cancer promotion, tumour cell-intrinsic STAT3 activation, tumour cell proliferation, and angiogenesis markers [139]. As previously mentioned, TH17 cell-mediated inflammation has been identified as playing a critical role in lung tumorigenesis [140]. Jungnickel et al. [73] indicated that the epithelial cytokine IL-17C mediates the tumour-promoting effect of bacteria, such as NTHi, through neutrophilic inflammation. There has been growing awareness of the importance of NTHi in the pathophysiology of COPD, and COPD-like airway inflammation induced by NTHi provides a tumour microenvironment that favours cancer promotion and progression [141,142,143]. Thus, NTHi may act as a bridge between COPD and lung cancer. Despite these encouraging findings associating specific microbes with carcinogenesis, distinguishing between microbes genuinely inducing cancer-causing pathways and those opportunistically colonising the tumour microenvironment remains a challenge. However, given the impact of dysregulated lung microbiomes on diseases such as COPD and lung cancer, the exploration of innovative therapeutics is warranted. Metal-based drugs offer a promising avenue with their potential to modulate the microbiome, alleviate inflammation, and directly target malignant cells.

4. Novel Therapeutic Strategies: Metal-Based Drugs

Metal-based remedies have played a crucial role in medicine throughout history. Ancient Egyptians discovered the therapeutic potential of gold (Au) salts, and alchemists mixed powdered Au into water to ease the aching of limbs, which is one of the primal references to arthritis [144]. Chinese medicine has long established the antiseptic competence of arsenic (As) [145], and the use of silver (Ag) in wound management can be traced back to the 17th century, during which silver nitrate (AgNO3) was administered to treat ulcers [146]. Paul Ehrlich and his co-workers developed the first successful metalloid complex in the 1900s, an As-based therapeutic agent named salvarsan. This effectively cleared a syphilis infection for which no prior treatment was available [147,148]. Nevertheless, research into metal-based coordination complexes was not stimulated until the fortuitous discovery of a platinum (Pt) complex known as cisplatin by Barnett Rosenberg [149]. Since then, cisplatin has been extensively studied as a chemotherapeutic drug and is broadly administered to various cancers, including ovarian, lung, head and neck, testicular, and bladder [150,151]. It is particularly effective in testicular germ cell tumours, leading to complete remission in approximately 70–80% of treated patients [152], and combination treatment with radiotherapy is more successful than radiotherapy alone in cohorts of NSCLC [153]. Cisplatin is structurally a coordination compound with square planar geometry and exerts its antitumour activity via intra-cellular hydrolysis and subsequent covalent binding to DNA-forming adducts, triggering apoptosis [154]. Despite its triumph, cisplatin has several disadvantages, including toxicity and drug resistance [155,156,157]. Since the discovery of cisplatin, a range of metal-containing drugs have been approved to treat various conditions, including cancer, rheumatoid arthritis, anaemia, iron overload, bipolar disorder, and gastrointestinal disorders [42] as presented in Table 1.
The realm of metal-based drugs also holds substantial promise in the treatment of lung cancer, offering novel mechanisms of action that may help overcome the limitations of traditional therapies [43,158,159,160]. The unique properties of metals, bring forth new avenues for attacking malignant cells while mitigating side effects associated with conventional cancer treatment [161]. Despite significant advances in personalised medicine and the development of therapies specifically targeting driver mutations, Pt-based chemotherapy remains a viable option in the management of NSCLC [162]. Cisplatin, the pioneer in this class, has been extensively used as a first-line therapy for NSCLC, and carboplatin, a less nephrotoxic and neurotoxic derivative of cisplatin, is frequently administered in conjunction with other agents as part of combined chemotherapeutic protocols for NSCLC [163,164]. A significant clinical challenge in the treatment of NSCLC patients with Pt agents, in particular cisplatin, is intrinsic and acquired resistance, which necessitates alternative metal-based drugs [165]. Ruthenium (Ru) complexes, such as KP1019 and NAMI-A, have shown potential in clinical studies in patients with NSCLC [166,167]. Ru(III)-based complexes can mimic iron, allowing them to be taken up by cells where Ru(III) can be reduced to Ru(II) in the hypoxic environments typical of solid tumours, which then irreversibly bind to DNA and proteins, inducing cytotoxic effects [168]. Another emerging class of metal-based drugs involves Au(I) complexes, such as auranofin, which have been shown to exhibit anticancer activity due to their ability to inhibit the selenoenzyme thioredoxin reductase (TrxR) [169]. A phase II clinical trial which assessed auranofin with sirolimus in patients with advanced or recurrent NSCLC found the combination arrested and regressed tumours [170]. Copper (Cu) complexes, such as Casiopeinas, have demonstrated anticancer properties through mechanisms such as DNA damage, ROS generation, and topoisomerase II inhibition [171]. Tetrathiomolybdate, a copper chelator, was assessed in a phase I clinical trial for NSCLC; however, it was found to be more effective in breast cancer targeting the copper-dependent components of the tumour microenvironment [172]. Additionally, there is growing interest in gallium (Ga) compounds due to their ability to disrupt the iron-dependent processes critical for cancer cell growth. For instance, gallium nitrate has shown potential in combating NSCLC [173]. While the application of metal-based drugs in the treatment of lung disease exhibits promise, there are several limiting factors that need to be addressed [174]. One major limitation lies in the potential adverse effects and systemic toxicity associated with these therapeutics. Such compounds often display indiscriminate toxicity towards both cancerous and healthy cells, leading to a host of side effects, including toxicity, and myelosuppression, which can severely compromise patient health and limit the dosage that can be safely administered [175,176]. Furthermore, the issue of drug resistance often surfaces with prolonged use of these agents, mirroring the challenges faced with more traditional chemotherapeutic agents [177]. The pharmacokinetics of metal-based drugs can also present limitations; poor bioavailability and non-specific biodistribution can affect drug efficacy and increase the potential for systemic toxicity [178,179]. The future of metal-based drugs in lung cancer treatment will depend on concerted efforts in research, clinical trials, and the translation of findings to clinical practice.

5. Metal Drugs as Microbiome Modulators

The current antimicrobial clinical pipeline is inadequate to treat mounting infections caused by multidrug-resistant pathogens [180]. Of the 12 new antibacterial agents approved for clinical use since 2017, the WHO reported that only one compound, cefiderocol, meets their innovation criteria (absence of known cross-resistance, new target, a new mode of action, or new class) that also has activity against all three critical priority pathogens [181]. Cefiderocol is a siderophore-conjugated cephalosporin that promotes the formation of chelated complexes with ferric iron and facilitates siderophore-like transport across the outer membrane of Gram-negative bacteria using iron transport systems accumulating in the periplasmic space [182,183]. This has highlighted the essential need to investigate ‘non-traditional’ approaches to antibacterial therapy that explore different avenues compared to ‘traditional’ organic molecules that target pathogens through already established targets [184,185]. These agents can prevent or treat bacteria through several modes of action, including directly or indirectly inhibiting growth, dampening virulence, truncating, or removing biofilm, alleviating resistance, restoring the natural microbiome, or boosting the immune system to clear or manage infections [184,186,187].
Metal-bearing drugs can adopt a range of coordination geometries and redox states, allowing for more significant chemical variations when compared with purely organic antibiotics, with different and potentially multi-modal mechanisms of action [40,179,188]. For instance, a recent study by the Community for Open Antimicrobial Drug Discovery (CO-ADD), a global free open-access screening initiative, discussed metal complexes’ enhanced activity profile [189]. The group evaluated 906 individual metal compounds within their database, from d-block elements, against critical ESKAPE bacteria and fungi, and found an impressive success rate of the metal compounds (9.9%) in comparison to solely organic molecules (0.87%). From this panel of metal complexes, 88 demonstrated activity (minimum inhibitory concentration (MIC) ≤ 16 µg/mL or 10 µM) against one of their tested strains (58 against fungi and 30 against bacteria) while also being tolerated by mammalian cells (CC50 > 32 µg/mL or >20 µM against human embryonic kidney cell line) and not demonstrating haemolytic activity (HC10 > 32 µg/mL or >20 µM). Only 14 of these metal complexes showed activity (MIC ≤ 32 µg/mL) against Gram-negative bacteria, including pathogenic bacteria tied to COPD and lung cancer [83,114]. Overall, the group emphasised the potential therapeutic capabilities of metal compounds due to the extent of possible modes of action, with broader coverage of three-dimensional chemical space than their organic counterparts [189].
The diverse antimicrobial mechanisms of metal-based drugs offer a promising avenue of exploration for the treatment of respiratory diseases, where dysbiosis plays a significant role, including COPD and lung cancer. The subsequent sections will focus on notable discoveries of metal-based compounds that have been studied as antimicrobial agents and exhibit the characteristics of microbiome modulators. However, it is in no way exhaustive. The coordination of metals to organic ligands such as 1,10-phenanthroline (phen) will also be discussed, as metal-phen complexes are emerging as tangible alternatives to the traditional antibiotic, with some studies reporting targeted inhibition and suppression of virulence as opposed to indiscriminate toxicity [190]. This presents an exciting new frontier for future research and therapeutic strategies for these prevalent and challenging diseases.

5.1. Bismuth (Bi)

Bi compounds have been utilised for many years to treat gastrointestinal disorders, including H. pylori infections, which are commonly associated with gastritis, and peptic ulcer disease, and are a well-established risk factor associated with the development of gastric mucosa-associated lymphoid tissue (MALT) lymphoma and gastric cancer [191,192,193]. In fact, due to the acidic environment of the stomach, it was also historically thought to be a sterile organ until the landmark discovery of H. pylori infection and its association with gastric disease [194]. H. pylori is a Gram-negative, microaerophilic bacteria which colonises the gastric mucosa of over half the world’s population, making it one of the most widespread bacterial infections [195]. Although 80% of H. pylori-infected individuals remain asymptomatic, some develop chronic gastritis, peptic ulcers, and eventually gastric cancer. Gastric cancer is the fourth leading cause of cancer-related deaths globally, and H. pylori infection accounts for nearly 90% of all non-cardia gastric adenocarcinomas [196]. Eradication of H. pylori infection can reduce the risk of gastric cancer development, especially if treated early before the onset of precancerous lesions [197]. The standard treatment for H. pylori infection is a combination of proton pump inhibitors (PPIs) and antibiotics (clarithromycin, amoxicillin, or metronidazole). The emergence of antibiotic resistance has led to a decline in the effectiveness of these regimens [198]. However, the synergistic effect between Bi salts and antibiotics has been observed, making Bi-containing quadruple therapy a recommended first-line treatment in areas with a high prevalence of antibiotic resistance [199]. Bi has a multi-targeted mode of antimicrobial activity by disrupting the bacterial cell envelope, interfering with enzyme function, inhibiting bacterial protein synthesis, and disrupting nickel homeostasis [40,200]. Nickel is essential for the survival and pathogenesis of H. pylori, as it regulates nickel acquisition, storage, delivery, and efflux via the synthesis of various metalloproteins/chaperones [201]. Bi drugs can interfere with nickel homeostasis by binding to nickel-associated proteins that play a critical role in urease and [Ni,Fe]-hydrogenase maturation, leading to the inhibition of enzyme activity. It has been widely reported that in addition to antimicrobial activity, Bi compounds exhibit anti-inflammatory and gastroprotective properties, which contribute to their effectiveness in treating these conditions and highlight their potential as a microbiome modulator [202,203].

5.2. Gold (Au)

Auranofin, an Au-based compound initially developed and approved for rheumatoid arthritis, has recently been recognised as a promising microbiome modulator. The antimicrobial potency of auranofin has been demonstrated against a broad spectrum of pathogenic bacteria, including antibiotic-resistant strains, such as Clostridioides difficile, M. tuberculosis, methicillin-resistant S. aureus (MRSA), and vancomycin-resistant Enterococcus (VRE), both in vitro and in vivo [49,204,205,206]. The modus operandi of its antimicrobial action is its ability to disrupt the redox balance within bacteria by inhibiting the TrxR enzyme, an essential component of their antioxidant defence mechanism [205]. Additionally, auranofin has been shown to have anti-virulence properties, reducing the production of key virulence factors, including proteases, lipase, and haemagglutinin [207]. Interestingly, auranofin’s influence on the microbiome extends beyond its antimicrobial properties. The inherent anti-inflammatory properties of the drug also hold implications for microbiome modulation. It achieves this by curtailing the expression of pro-inflammatory cytokine IL-6 via the inhibition of the NF-κB-IL-6-STAT3 signalling cascade, a critical pathway involved in the pathogenesis of various inflammatory diseases [208]. By mitigating local inflammation, auranofin maintains the integrity of the host barrier, thereby fostering a more conducive environment for beneficial bacteria. For instance, in murine models, auranofin treatment led to decreased inflammation and a shift towards a more balanced microbiota, characterised by an increase in anti-inflammatory bacterial strains such as Faecalibacterium prausnitzii [209,210,211]. Thus, auranofin presents a dual action microbiome modulator, harnessing both antimicrobial and anti-inflammatory capabilities. This novel mechanism of action offers a promising avenue for therapeutic intervention in conditions characterised by microbial imbalance, although further research is warranted to fully elucidate its potential clinical utility. Moreover, recent studies have focused on auranofins’ potential asan anticancer agent, showing its efficacy against various cancers, including lung cancer (IC50 < 2 μM, for A549) [169,212,213,214], while also enhancing Ibrutinib (tyrosine kinase inhibitor) activity in EGFR-mutant lung adenocarcinoma [215]. Mimicking its antimicrobial mechanism of action, auranofin is a selective inhibitor of TrxR, triggering increased production of ROS and activating the p38 mitotic activated protein kinase (p38 MAPK) [216]. It has also been reported that auranofin can inhibit proteasome-associated deubiquitinases (DUB), deregulating the ubiquitin-proteasome system (UPS) [217]. The proven effectiveness of auranofin in cancer management has sparked interest among pharmaceutical chemists in exploring other Au(I) complexes for their potential therapeutic roles. The future of Au-based drugs, thus, presents an intriguing avenue for cancer treatment, in particular cancer that arises from microbiome dysbiosis.

5.3. Silver (Ag)

Silver has long been known for its potent antimicrobial properties, both historically and in modern times. A variety of medical products containing silver are available, such as bandages, ointments, and catheters in the form of nanocrystalline silver (including silver nanoparticles and colloidal silver), silver nitrate, and silver sulfadiazine (a complex formed with the antibiotic sulfadiazine) [218]. Ag(I) compounds have well-documented multi-modal properties that exhibit broad-spectrum activity against a wide range of bacterial species, including Gram-positive and Gram-negative bacteria, as well as fungi and viruses [219,220]. The toxicity of Ag(I) compounds primarily stems from the release of ions that interact with the cell envelope and destabilise the membrane [221,222], coupled with nucleic acids and proteins disrupting replication and synthesis [223] and inhibiting metabolic pathways [199,224]. Although Ag(I) is generally not regarded as redox-active, the generation of ROS is also attributed to its antibacterial activity [35,48,219]. However, it is thought that the production of ROS indirectly occurs through the perturbation of the respiratory electron transfer chain [225], Fenton chemistry following destabilisation of Fe-S clusters or displacement of Fe [226] and inhibition of anti-ROS defences by thiol–Ag bond formation [227]. Studies have also shown that Ag(I) often exhibits synergistic effects when combined with a range of antibiotics, such as β-lactams [228,229,230], aminoglycosides [48,230,231,232] and fluoroquinolones [229,230], and tetracyclines [48], against both planktonic and biofilm forms. While the direct antimicrobial effects of Ag(I) are well documented, its role as a microbiome modulator has been an area of growing interest in recent years and has produced contradictory results thus far. An in vivo study evaluating changes in the populations of intestinal-microbiota and intestinal-mucosal gene expression in rats after oral administration of Ag nanoparticles (AgNP) (9, 18, and 36 mg/kg body weight/day) and silver acetate (100, 200, and 400 mg/kg body weight/day). The results indicate that exposure to AgNP prompted size- and dose-dependent changes to ileal mucosal microbial populations, as well as intestinal gene expression, and induced an apparent shift in the gut microbiota toward greater proportions of Gram-negative bacteria [233]. In contrast, another study found non-significant alterations in the Bacillota (Firmicutes) and Bacteroidota populations with no toxicological effects on rats that received AgNPs orally by gavage for 28 days [234]. Various studies have reported that the shape of AgNPs can influence their impact on gut microbiota. For example, cubic AgNPs have been shown to reduce the abundance of Christensenellaceae, Clostridium spp., Bacteroides uniformis, and Coprococcus eutectic [235], while spherical AgNPs decrease the presence of Oscillospira spp., Dehalobacterium spp., Peptococcaceae, Corynebacterium spp., and Aggregatibacter pneumotropica populations [236]. Ag(I) complexes have also been studied for their potential use as anticancer agents due to their unique properties and interaction with cellular components, such as DNA and proteins, leading to disruption of essential biological processes [237]. For example, Ag(I) complexes with N-heterocyclic carbene ligands have been studied for their anticancer activities against a range of cancer cell lines, including lung cancer [238].

6. 1,10-Phenanthroline and Its Metal Complexes

The coordination of a metal ion to a biologically active ligand can serve to facilitate the uptake of the non-lipophilic metal across the lipophilic cell envelope or to act synergistically with the metal, such that the combined toxic effects of the metal and the active ligand will exert enhanced and targeted activity in the problematic cell [35,189]. 1,10-Phenanthroline (phen) is a heterocyclic and chelating bidentate ligand for metal ions, which has played an important role in advancing coordination chemistry. The ideally placed nitrogen atoms, as seen in Figure 1, have rigid planar structure, hydrophobic, π-electron-deficient heteroaromatic, and acidic properties, allowing the ligand to assist in the stabilisation of a variety of metal complexes in various oxidation states [239,240]. Phen has long shown antibacterial effects in an in vitro setting, demonstrating excellent bacteriostatic activity on Gram-positive and Gram-negative species of pathogenic bacteria [241,242]. This antimicrobial action has been attributed to the chelating capabilities of phen and the sequestering nature of metal ions [243]. Thus, it selectively disturbs the essential metal cellular metabolism interference with metal acquisition and bioavailability for crucial reactions impeding the microbial nutrition, growth, virulence, and pathogenesis of a variety of microorganisms, including Leishmania spp., Aspergillus spp., Candida albicans, Fonsecaea pedrosoi, and Streptococcus agalactiae [244,245]. Accordingly, metal chelators, such as phen, have been investigated for their therapeutic potential against microbial infections, as metals such as Fe, Cu, and zinc (Zn) play an important role in cellular homeostasis [190]. Moreover, it was shown that the sequestered metals produce a metal-phen complex, and the emerging complex drives the toxic effects [246]. Investigations into the in vitro antibacterial activity of phen derivatives (Figure 1), such as 3,4,7,8-tetramethyl-1,10-phenanthroline, 5-nitro-1,10-phenanthroline, 1,10-phenanthroline-5,6-dione, 2,9-dimethyl-1,10-phenanthroline, and various others have also been undertaken [247]. A significant increase in biocidal activity was achieved when the various ligands were coordinated with Cu(II) ions, with the 2,9-dimethyl derivative being the most active against S. aureus and Escherichia coli.
Many anticancer metal complexes with cytotoxicity have been reported based on phen bidentate ligands [248,249,250,251,252]. Compounds with the general formula [Cu(L-dipeptide)(phen)]·nH2O have been screened for anticancer activity in lung (A549) cancer cells showing anticancer potencies in the micromolar concentration range [253]. Mixed Cu(II) phen-based complexes with the general formula Cu(N-N1)x(OH2)y(ClO4)z, where N-N1 = phen, were reported to be toxic towards the squamous cell carcinoma (SKMES-1) cell line ranging from picomolar to micromolar in terms of their IC50 [254,255]. The mechanism by which Cu(II)-phen complexes exert their toxicity is reported to be through strong DNA binding activity and by inducing oxidative stress through mitochondrial dysfunction initiating apoptosis [249,250]. A series of Au(III) bearing phen and derivative 2,2′-bipyridine (bipy) scaffolds have shown promising anticancer activity against the A549 cell line, which exhibited IC50 values that were significantly lower than those of cisplatin control, possibly inducing cell death via a TrxR-mediated mechanism [256]. The coordination complexes [AuIII(5-chloro-phen)Cl2]PF6) and [AuIII(bipy)Cl2]PF6 displayed potent cytotoxicity in the A549 cell line, potentially through the inhibition of the water and glycerol channel aquaglyceroporin-3 (AQP3), which play crucial roles in cell apoptosis, proliferation, and migration and therefore have been proposed as new drug targets for cancer treatment [257]. Ru-phen complexes, such as [Ru(phen)2dppz]2+ (where dppz=dipyrido[3,2-a:2′,3′-c]phenazine), has been shown to intercalate with DNA and induce photocleavage, leading to significant cytotoxicity in various cancer cell lines, including lung cancer [258]. The octahedral complex [Ru(phen)3]2+ can intercalate and unwind DNA as effectively as ethidium bromide [259]. Silver(I) complex Ag(Phen)2(CH3COO)· H2O exhibits anticancer properties against lung adenocarcinoma (A549) cells through ROS production, which in turn induces a change in the mitochondrial membrane potential [260]. Overall, the exact mechanism of action of these metal-phen complexes in the context of lung cancer still requires further exploration. More comprehensive in vitro and in vivo studies, followed by clinical trials, are necessary to fully understand their potential.
Metal complexes containing phen, and its derivatives, have also been reported in the literature for their antifungal [244,261,262,263,264], antiparasitic [265,266,267,268], antiviral [269,270,271,272], and antibacterial [273,274,275,276,277] competence. The work of polypyridyl metal complexes was pioneered by Francis Dwyer and co-workers when they published a landmark study outlining the biological activity of dicationic Ru, Fe, nickel (Ni), cobalt (Co), Cu, Zn, calcium (Ca), and manganese (Mn) chelates containing phen and its derivatives [278]. Their work established the in vitro toxicity of the metal-phen chelates against M. tuberculosis, S. aureus, S. pneumonia, Clostridium perfringens, E. coli, and Proteus vulgaris (presented in descending order of activity), while the metal-free phen demonstrated dampened effects. Furthermore, the toxicity exerted was independent of the metal present, except for M. tuberculosis. They also identified that the bacteria and fungi did not develop any significant resistance to the selected chelates. In vivo bacterial infection treatment studies using mice and guinea pig models showed that metal-phen chelates were clinically useful as topical antimicrobials. However, the selected compounds were chemotherapeutically ineffective when administered intravenously due to rapid clearance from the bloodstream [278]. Although the results were promising, interest in polypyridyl metal complexes as potential antimicrobial chemotherapeutics was void in the pharmaceutical sector, possibly due to the vast array of antimicrobials in the pipeline or the immensely lower incidence of resistance at the time. In modern times, transition metal-based compounds have had a revival of interest as possible alternatives or adjuvants to the current armamentarium of antimicrobial agents that cannot contend with the rapid emergence of resistant microbes worldwide [35].

7. Mechanisms of Metal-Phen Complexes

Research into the possible mechanisms by which promising phen-based complexes exert their toxic effects has been carried out within in vitro models, including mammalian, fungal, and bacterial cells, and in a range of in vivo biological models. McCann et al. [279] proposed modes of antifungal action by phen and its derivatives to be (i) the dysfunction and disruption of the cell membrane along with the withdrawal of the cytoplasmic membrane, (ii) drug-induced circumvention on the action of cell division (budding), (iii) damage to the functional mechanisms of the mitochondria, (iv) chelation or sequestering of trace metal ions which inhibits glycosylphosphatidylinositol (GPI) biosynthesis, (v) rupturing of internal organelles along with the enlargement of the nucleus, and (vi) the degrading of nuclear DNA [279]. The following sections discuss the potential antibacterial capability of metal-phen complexes and their mechanism of action.

7.1. The Bacterial Cell Envelope and Activity of Metal-Phen Complexes

The cell wall and outer cell membrane are considered a significant obstacle in developing novel antibacterial agents that are effective against Gram-negative bacteria, as they must be lipophilic and able to penetrate the outer membrane. Contradictory to this, Ag(I) has well-documented bactericidal properties. In its cationic form, Ag(I), it is oligodynamic and displays a broad spectrum of activity that is dependent upon the slow release of biologically active ions thought to bind to the bacterial cell surface and interfere with growth by inhibiting transport across the cell wall and membrane [35,219]. Being charged entities, free metal ions require protein- or ionophore-mediated transport to cross a lipid bilayer, as these transporter proteins encase the cations in a hydrophobic sleeve to enable its passing. The complexation of metals to a hydrophobic chelating ligand such as phen can enact the same process, enabling their penetration through the bacterial cell envelope, presenting the chelate to its target biomolecule to inhibit cell growth or initiate cell death [280]. Cationic metal-phen chelates can be bacteriostatic [278] or bactericidal [281] towards many Gram-positive bacteria, including S. aureus and S. pyogenes and C. perfringens. However, they do not exhibit the same potency against Gram-negative bacteria. The lipophilicity of a complex is taken as a good measure of its ability to pass into the cell by diffusion, and in some cases, increased lipophilicity correlates with antimicrobial potency [282]. Dwyer et al. [283] were the first to identify this relationship when working with mononuclear Ru(II) and the phen complex, [Ru(phen)3]2+. Originally, this complex was shown to be inactive against all tested bacterial strains; however, when methyl substituents were incorporated into the phen ligand, [Ru(3,4,7,8-Me4phen)3]2+, the non-polar surface interaction was increased. This corresponded to an increase in activity against all examined bacteria, especially Gram-positive and acid-fast bacteria [283]. There has been renewed interest in the antimicrobial activity of polypyridylruthenium(II) complexes over the past decade. Crowley et al. [284] have used a series of 2-(1-R-1H-1,2,3-triazol-4-yl)pyridine “click” ligands (R-pytri) as functionalised analogues of 2,2′-bipyridine (bpy) and phen chelators to synthesise a diverse range of ruthenium complexes ([Ru(R-pytri)3](PF6)2), and examined their antimicrobial activities. Some complexes demonstrated moderate activity against Gram-positive strains, but this was not maintained when examined against Gram-negative bacteria. Ru(II) complexes [Ru(L)2amtp]2+ (L = bpy) [285] and [Ru(bpy)2L] (L = p-tFMPIP) [286] were also active against Gram-positive bacteria, particularly S. pneumoniae which is a problematic pathogen for COPD patients. The mode of action was reported to be interference with iron acquisition systems in S. pneumoniae cells [285], oxidative stress and membrane damage [286]. Moreover, the Ru(II) complexes were not toxic towards human bronchial epithelial cells [286] or A549 cell line [285]. Keene, Collins, and co-workers have extensively researched Ru(II) complexes and their potential as antimicrobial agents. They developed kinetically inert mono-, di-, tri-, and tetra-nuclear polypyridylruthenium(II) complexes, whereby the Ru(II) metal centres are bridged by flexible and lipophilic bis [4(4′-methyl-2,2′-bipyridyl)]-1),n-alkane ligand (bbn) and are collectively termed ‘Rubbn’ complexes, where n = the number of methyl groups in the bbn linker chain (n = 2, 5, 7, 10, 12, 14 or 16). The antibacterial activity of this series of complexes is correlated with increasing lipophilicity through increased length of the bbn chain and only slight differences were observed with enantiomeric forms of the complexes [285]. The dinuclear ruthenium complexes ‘Rubbn’ have been the most extensively studied and have produced exciting results. The Rubbn complexes enter bacteria in an energy-dependent manner as they significantly depolarise and permeabilise the cellular membrane [286]. There was preferential uptake of Rubbn in prokaryotes compared to cancerous cells, which was suggested to result from the greater presence of negatively charged components in the bacterial envelope [287]. Rubb16 was found to be the most active with selective toxicity towards bacteria. This complex condensed ribosomes when they existed as polysomes by preferentially binding to RNA over DNA in vivo, offering the interruption of protein synthesis in actively growing bacterial cells as a potential mode of action [288].
The corresponding tri- and tetra-nuclear complexes ‘Rubbn-tri’ and ‘Rubbn-tetra’ were the more active, demonstrating MIC’s four times their dinuclear analogues [289]. Although the level of cellular uptake in Gram-negative bacteria was similar to that of Gram-positive bacteria, there was significantly less activity in the former species. The authors suggested that this was a result of the inherent resistance of Gram-negative bacteria to inert polpyridylruthenium(II) complexes, particularly P. aeruginosa, in which the outer membrane permeability is 10- to 100-fold lower than, for example, that of E. coli [290]. However, while the antibacterial activity increased as the ruthenium centres and the length of the alkyl chain in the bbn ligand increased, the toxicity towards eukaryotic cells reduced selectivity [291].

7.2. DNA as an Antibacterial Target for Metal-Phen Complexes

The design of agents capable of controlled nucleic acid cleavage is of great importance, and since the initial work of Sigman et al. [292], there has been considerable attraction to artificial metallonucleases. The copper-bis-phenanthroline complex, [Cu(phen)2]2+, induced oxidative cleavage of DNA in the presence of a reductant, which is unusual for complexes containing phen, as this compound as a singular agent usually intercalates with DNA. DNA as a target offers a fresh avenue for potential antibacterial agents, as it has been relatively unexplored thus far [293]. A large number of publications have reported the enhanced antibacterial activity of quinolone/fluoroquinolone antibiotics containing metal(II)-phen complexes (metal = Cu(II), Ni(II), Co(II), Mn(II)) [294]. One example is the combination of Levofloxacin (lvx) with Cu(II), forming the binary complex, [Cu(lvx)]2+ that significantly increases DNA binding but is not stable at a physiological pH. However, the addition of phen as an N-donor forms a very stable ternary complex [Cu(lvx)(phen)]2+ [295]. Cu(II)/phenanthroline/fluoroquinolone complexes have demonstrated intercalation with DNA, exhibiting nuclease activity, and they are taken into the cell differently from that of the free fluoroquinolone drug [296]. Furthermore, when tested against a mutant strain of E. coli lacking porins, it was identified that the higher the hydrophobicity of the complexes, the higher the need for porins for their uptake [297]. Marmion et al. [298] rationally developed a family of metallo-antibiotics with the general formula [Cu(N,N)(CipA)Cl], where N,N represents a phenanthrene ligand and CipA is a derivative of fluoroquinolone antibiotic ciprofloxacin. The complexes were active against susceptible and resistant S. aureus bacteria, which were identified in the lungs of COPD patients during exacerbation. They appear to intercalate DNA via minor groove interactions and mediate DNA damage by generating ROS with superoxide and hydroxyl free radicals playing crucial roles in DNA strand scission [298]. Molecular docking analysis prompted the synthesis of derivatives [Cu(N,N)(CipHA)]NO3, where CipHA represents a hydroxamic acid of ciprofloxacin. Proteomic analysis of S. aureus exposed to two lead complexes [Cu(phen)(CipHA)]NO3 and [Cu(DPPZ)(CipHA)]NO3 (where DPPZ = dipyridophenazine) suggests that proteins involved in virulence, pathogenesis, and the synthesis of nucleotides and DNA repair mechanisms are most affected [299]. Metal-phen complexes without the addition of the quinolone ligand have also demonstrated moderate antibacterial activity, with DNA binding or nuclease activity as the proposed antibacterial mechanism of action. The metals include Ag(I) [300], Cu(I) [301], Cu(II) [302], Zn(II) [303], Pt(II) [304], Mn(II) [305], and Fe(III) [306].

7.3. The Activity of Metal-Phen Complexes on Biofilms

Bacterial biofilm communities differ from planktonic bacteria in various ways, such as growth rate, gene expression, and protein expression [307]. This is due to biofilms creating an altered microenvironment with higher osmolarity, nutrient scarcity, and increased cell density of heterogeneous bacterial communities [308,309]. Bacteria usually reside in biofilms, and biofilm-residing bacteria can be resilient to the immune system, antibiotics, and other treatments [307,310,311]. Such biofilms enable bacteria to persist in the lower respiratory tract, which can exacerbate the disease and complicate the treatment of patients with COPD and lung cancer [312,313,314]. Therefore, agents that can navigate the difficult terrain of a biofilm to the bacteria embedded within or dissociate the extracellular matrix to expose the bacteria are important.
Although there are few reports of metal-phen complexes with anti-biofilm activity, there have been some advances in the development of novel compounds. A range of Cu(II) complexes, [Cu(IL)(AL)]2+ (where IL represents functionalised phens and AL represents 1S,2S-1,2-diaminoethane or 1R,2R-diaminocyclohexane), were tested on a number of bacterial strains [315]. Although the metal complexes generated higher MICs (2–32 µg/mL) than the control antibiotic vancomycin (MIC = 0.25 µg/mL), they showed significant activity against S. aureus biofilms and were better at removing biofilms than vancomycin. For example, 100 µg/mL of vancomycin, which is 400-fold the MIC, reduced biofilm biomass by just 44%, whereas 25 µg/mL of [Cu(5,6-dimethyl-phen)(SS-dach)]2+ (equivalent to 3-fold the MIC) reduced the biofilm by 68% in only 2 h. This decrease in biomass was similar in all Cu complexes and is metal-dependent, as replacing the centre with Pt(II) or Pd(II) removed both the antibacterial and anti-biofilm action. Therefore, this suggests that the potential mechanism of action in both planktonic and biofilm cells is related to the nuclease activity of Cu, as neither Pt nor Pd possesses this capability, particularly given that the extracellular matrix contains a considerable quantity of nucleic acids [315]. Similarly, Mn(II), Cu(II), and Ag(I) complexes incorporating phen and 3,6,9-trioxaundecanedioate (tddaH2) (Figure 2) showed enhanced activity profiles when tested against clinical isolates of P. aeruginosa [232], a bacteria frequently reported to be problematic to both COPD and lung cancer patients. The results showed that the metal-tdda-phen complexes could prevent biofilm formation, in relation to mass and cellular viability, to a greater capacity than gentamicin across the clinical strains and disturb mature biofilm. This was supported by reducing the separate biofilm components examined, suggesting extracellular DNA (eDNA) and extracellular polysaccharides as potential molecular targets [232]. The ability to act on P. aeruginosa clinical isolates synergistically with gentamicin on mature biofilms prompted in vivo studies using G. mellonella larvae [316]. Mn-tdda-phen and Ag-tdda-phen were able to clear a P. aeruginosa infection at concentrations that are non-toxic towards G. mellonella larvae in a multi-modal fashion by acting directly on the bacteria in addition to stimulating both the cellular (hemocytes) and humoral (immune-related peptides, specifically transferrin and inducible metalloproteinase inhibitor) immune response of the larvae. The amalgamation of metal-tdda-phen complexes and gentamicin further intensified this response at lower concentrations, clearing a P. aeruginosa infection that was previously resistant to gentamicin alone [316]. The same complexes have also been reported to have antitubercular activity and were highly toxic towards the NSCLC (A549) cell line [274]. Gandra et al. [261] reported on the antifungal capabilities of metal-phen complexes against isolates of three species that make up the highly resistant Candida haemulonii species complex, an emerging opportunistic pathogen that has been reported to be problematic in COPD and lung cancer [317,318]. Mn(II)- and Ag(I)-phen chelates could conserve antifungal activity at concentrations that were reasonably non-toxic toward G. mellonella. Most notable was the Mn-tdda-phen complex, as it induced the lowest mortality rate while reducing the fungal burden on infected larvae and could also affect the virulence of C. haemulonii [262]. Across all studies, the inclusion of phen was paramount to the potency of the complexes, with the addition of tddaH2 enhancing their water solubility and mode of action in various microbial cells, which are problematic in chronic lung diseases, particularly highlighting COPD and lung cancer. A series of complexes incorporating phen and cyanoguanidine (cnge) have been reported with the general formula M(II)/phen/cnge (where M = Cd, Cu, or Zn) [319]. The cadmium complex [Cd(phen)2(SO4)H2O]cnge· 5H2O possesses enhanced activity across the assessed bacterial and fungal pathogen in comparison to its metal derivatives. This prompted subsequent anti-biofilm analysis against P. aeruginosa laboratory strain ATCC 27853, in which 0.5 x MIC (93.5 µg/mL) of the metal complex inhibited approximately 40% of biofilm formation. Moreover, there was a reduction in the acute toxicity of the phen ligand when it was incorporated into the Cd(II)/phen/cnge complex within a crustacean model, Artemia salina.
A derivative of phen, 1,10-phenanthroline-5,6-dione (phendione, Figure 1), has also been investigated for its antibacterial and anti-biofilm capabilities. This compound contains an o-quinoid moiety and forms strong complexes with transition metal ions at the N-N terminal, with predominance toward Zn(II) and to a lesser extent for Fe(II), Ca(II), Cu(II), Co(II), and Mn(II) [279]. Tay et al. [320] reported the MIC and minimum bactericidal concentration (MBC) values of phendione for Enterococcus faecalis as 2 μg/mL and 16 μg/mL, respectively, relating its activity to its ability to sequester Zn(II) from metalloenzymes. In order to kill E. faecalis cells embedded in a biofilm structure, an MIC four times that required to kill planktonic bacteria was required. However, the biofilm was eradicated at this concentration. Although the authors could not explain the mechanism by which phendione eradicates E. faecalis biofilm, they speculated that it may weaken the extracellular polymeric substances of the biofilm by disrupting Zn(II) balance. The metal-free phendione and coordinated Cu(II) and Ag(I) complexes, [Ag(phendione)2]ClO4 (Ag-phendione) and [Cu(phendione)3](ClO4)4H2O (Cu-phendione) (Figure 3) have been extensively studied across a variety of microbial cells [321]. The metal-phendione complexes were able to inhibit the growth of the Phialophora verrucosa [263,264], Pseudallescheria boydii [244], Trichomonas vaginalis [266] and Candida albicans [279]. Viganor et al. [273] investigated the effect of metal-phendione complexes on planktonic and biofilm-growing P. aeruginosa. The compounds presented the following potency against susceptible and resistant planktonic cells: Cu-phendione (7.76 µM) > Ag-phendione (14.05 µM) > phendione (31.15 µM) > phen (579.28 µM). It was also discovered that the compounds could disrupt a mature biofilm in a dose-dependent manner, especially Ag-phendione (IC50 = 9.39 µM) and Cu-phendione (IC50 = 10.16 µM). The metal-phendione complexes were reported to cause a significant reduction in the expression of the metalloenzyme Elastase B produced by P. aeruginosa at a gene and mature protein production level [322], therefore suggesting this as a potential therapeutic target. Furthermore, the complexes also offer protective action for lung epithelial cells. Metal-phendione complexes can interact with double-stranded DNA and promote oxidative damage, suggesting multiple mechanisms of action in P. aeruginosa [323]. The same activity sequence of the test complexes (Cu-phendione > Ag-phendione > phendione) was maintained when assessed in both planktonic- and biofilm-forming cells of MDR Acinetobacter baumannii [276] and Klebsiella pneumoniae [277] clinical isolates. The combination of either Cu-phendione or Ag-phendione with meropenem had synergistic activities, according to their fractional inhibitory concentration, against Klebsiella pneumoniae carbapenemase (KPC)-producing K. pneumoniae clinical isolates. Moreover, the combination of metal complex and meropenem restored the antibiotic’s efficacy (in terms of MIC) in 87% of phenotypically resistant strains [277]. The metal-phendione complexes also had low toxicity in G. mellonella larvae [262,264] and mice [243], reinforcing that these compounds may represent a novel antimicrobial agent with potential therapeutic applications across a variety of pathogens that predominate in chronic lung disorders.

8. Conclusions and Future Perspectives

The field of lung microbiome research is rapidly expanding, but faces numerous challenges [7,324,325,326]. The low microbial biomass and high host contamination make it challenging to use metagenomics and metatranscriptomics to understand the haemostatic and pro-oncogenic functions of the microbiome. Secondly, chronic lung diseases are heterogeneous, making it necessary to untangle the complex relationships between microbiome, disease phenotypes, and endotypes. This understanding is crucial for determining the role of the microbiome in the development and progression of these diseases, including the transition to cancer. Thirdly, unlike the well-characterised gut microbiome, little is known about the specific mediators produced by the lung microbiome and their functions. A systems biology approach is required to study the interaction between airway microbes and host in diseased states. There needs to be a standard procedure for manipulating airway microbiota in animal studies, which is crucial for assessing its functional impacts. While NGS is powerful, culturing microbes from the respiratory tract is essential for translational research to gain a comprehensive understanding of the pro-oncogenic pathways that are being activated. Overall, these challenges hinder our understanding of the lung microbiome and its potential implications for health and disease progression.
Metal-based drugs, including metal-phen complexes, have emerged as promising candidates for managing chronic lung diseases and related microbial complications [38,200,327,328]. These drugs have demonstrated significant antimicrobial, anti-biofilm, and anti-virulence activity, targeting multiple pathways in pathogenic bacteria, distinct from traditional antibiotics. The potential of these complexes to act against antibiotic-resistant strains and their synergy with existing antibiotics could also help tackle the pressing issue of antibiotic resistance. While these avenues are promising, it is essential to note that the development of metal-based drugs as therapies for microbiome modulation is in its nascent stages. Moreover, the link between microbiome modulation and cancer prevention remains a complex issue that requires more research [118,120,329,330]. While it is known that chronic inflammation can contribute to cancer development and that bacterial imbalances in the lungs can drive this inflammation, it is not yet entirely clear how effectively and consistently altering the microbiome can reduce cancer risk. Nevertheless, the versatility of metal drugs and metal-phen complexes and their potential for further chemical modification opens new avenues for developing effective antimicrobial and anticancer agents. While it is promising that metal complexes are currently under evaluation in clinical trials for other indications, it is crucial to substantiate whether these metal-based drugs can effectively safeguard humans against pathogenic bacteria. Furthermore, it is critical to consider that while metal-based drugs may successfully kill or inhibit the growth of pathogenic bacteria, they may also impact beneficial bacteria within the lung microbiome, potentially leading to unintended consequences.
The future of research in the lung microbiome requires innovative experimental and analytical strategies to overcome these challenges. Longitudinal, interventional, and mechanistic studies are needed to determine causality. These studies aim to address fundamental scientific questions about the lung microbiome, such as its baseline status in healthy individuals, how it responds to environmental factors, its roles in different types of respiratory diseases, and how its dysregulation can advance disease to cancer. Researchers also hope to determine whether the microbiome can be used as a marker for the diagnosis, phenotyping, and prognosis of respiratory diseases. Additionally, they seek to understand the topographic structure and spatial dynamics of microbial communities in the lung and the interactions and influences between respiratory bacteria, fungi, and viruses on the host immune system. Identifying key microbial metabolites that regulate host inflammation or other processes in the respiratory tract is of interest. Ultimately, the goal is to use the airway microbiome as a biomarker and therapeutic target for precision medicine in respiratory and broader human diseases.
To conclude, the potential of metal-based drugs to serve as novel therapeutics for COPD and lung cancer is compelling. By deepening our understanding of their interaction with the diverse bacterial communities within these conditions and through rigorous clinical testing, we can pave the way for innovative treatments that could significantly improve patient outcomes.

Author Contributions

Conceptualisation—M.O. and A.-M.B., writing—original draft preparation, M.O.; writing—review and editing, A.-M.B. and O.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Irish Research Council under the Government of Ireland Postdoctoral Fellowship Programme (M.O.), grant number GOIPD/2022/522, and through the Trinity College Dublin Med seed funding programme (M.O.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable—no new experimental data were created.

Acknowledgments

The authors would like to dedicate this paper to Thomas Glennon.

Conflicts of Interest

A.-M.B. has received honoraria from Roche and AstraZeneca. A.-M.B. is president of Lung Cancer Europe (LuCE), which has received support from Amgen, AstraZeneca, Bayer, Blueprint Medicines, Bristol Myers Squibb, Boehringer Ingelheim, Daiichi-Sankyo, Lilly, Merck, MSD, Novartis, Pfizer, Regeneron, Roche, Sanofi, Takeda, Janssen, Novocure, and ThermoFisher.

References

  1. Aggarwal, N.; Kitano, S.; Puah, G.R.Y.; Kittelmann, S.; Hwang, I.Y.; Chang, M.W. Microbiome and Human Health: Current Understanding, Engineering, and Enabling Technologies. Chem. Rev. 2023, 123, 31–72. [Google Scholar] [CrossRef] [PubMed]
  2. Hou, K.; Wu, Z.X.; Chen, X.Y.; Wang, J.Q.; Zhang, D.; Xiao, C.; Zhu, D.; Koya, J.B.; Wei, L.; Li, J.; et al. Microbiota in health and diseases. Signal Transduct. Target. Ther. 2022, 7, 135. [Google Scholar] [CrossRef] [PubMed]
  3. Dickson, R.P.; Erb-Downward, J.R.; Martinez, F.J.; Huffnagle, G.B. The Microbiome and the Respiratory Tract. Annu. Rev. Physiol. 2016, 78, 481–504. [Google Scholar] [CrossRef] [Green Version]
  4. Wensel, C.R.; Pluznick, J.L.; Salzberg, S.L.; Sears, C.L. Next-generation sequencing: Insights to advance clinical investigations of the microbiome. J. Clin. Investig. 2022, 132, 154944. [Google Scholar] [CrossRef]
  5. Wypych, T.P.; Wickramasinghe, L.C.; Marsland, B.J. The influence of the microbiome on respiratory health. Nat. Immunol. 2019, 20, 1279–1290. [Google Scholar] [CrossRef]
  6. Kogut, M.H.; Lee, A.; Santin, E. Microbiome and pathogen interaction with the immune system. Poult. Sci. 2020, 99, 1906–1913. [Google Scholar] [CrossRef]
  7. Natalini, J.G.; Singh, S.; Segal, L.N. The dynamic lung microbiome in health and disease. Nat. Rev. Microbiol. 2022, 21, 222–235. [Google Scholar] [CrossRef] [PubMed]
  8. Garcia-Nuñez, M.; Millares, L.; Pomares, X.; Ferrari, R.; Pérez-Brocal, V.; Gallego, M.; Espasa, M.; Moya, A.; Monsó, E. Severity-related changes of bronchial microbiome in chronic obstructive pulmonary disease. J. Clin. Microbiol. 2014, 52, 4217–4223. [Google Scholar] [CrossRef] [Green Version]
  9. Leng, Q.; Holden, V.K.; Deepak, J.; Todd, N.W.; Jiang, F. Microbiota biomarkers for lung cancer. Diagnostics 2021, 11, 407. [Google Scholar] [CrossRef]
  10. Gunaydin, G.; Gedik, M.E.; Ayan, S. Photodynamic Therapy for the Treatment and Diagnosis of Cancer—A Review of the Current Clinical Status. Front. Chem. 2021, 9, 686303. [Google Scholar] [CrossRef] [PubMed]
  11. Perrone, F.; Belluomini, L.; Mazzotta, M.; Bianconi, M.; Di Noia, V.; Meacci, F.; Montrone, M.; Pignataro, D.; Prelaj, A.; Rinaldi, S.; et al. Exploring the role of respiratory microbiome in lung cancer: A systematic review. Crit. Rev. Oncol. Hematol. 2021, 164, 103404. [Google Scholar] [CrossRef] [PubMed]
  12. Russo, C.; Colaianni, V.; Ielo, G.; Valle, M.S.; Spicuzza, L.; Malaguarnera, L. Impact of Lung Microbiota on COPD. Biomedicines 2022, 10, 1337. [Google Scholar] [CrossRef] [PubMed]
  13. Safiri, S.; Carson-Chahhoud, K.; Noori, M.; Nejadghaderi, S.A.; Sullman, M.J.M.; Ahmadian Heris, J.; Ansarin, K.; Mansournia, M.A.; Collins, G.S.; Kolahi, A.A.; et al. Burden of chronic obstructive pulmonary disease and its attributable risk factors in 204 countries and territories, 1990–2019: Results from the Global Burden of Disease Study 2019. BMJ 2022, 378, e069679. [Google Scholar] [CrossRef] [PubMed]
  14. Adeloye, D.; Song, P.; Zhu, Y.; Campbell, H.; Sheikh, A.; Rudan, I. Global, regional, and national prevalence of, and risk factors for, chronic obstructive pulmonary disease (COPD) in 2019: A systematic review and modelling analysis. Lancet Respir. Med. 2022, 10, 447–458. [Google Scholar] [CrossRef]
  15. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA. Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  16. Hou, W.; Hu, S.; Li, C.; Ma, H.; Wang, Q.; Meng, G.; Guo, T.; Zhang, J. Cigarette Smoke Induced Lung Barrier Dysfunction, EMT, and Tissue Remodeling: A Possible Link between COPD and Lung Cancer. Biomed Res. Int. 2019, 2019, 2025636. [Google Scholar] [CrossRef]
  17. Parris, B.A.; O’Farrell, H.E.; Fong, K.M.; Yang, I.A. Chronic obstructive pulmonary disease (COPD) and lung cancer: Common pathways for pathogenesis. J. Thorac. Dis. 2019, 11, 2155–2172. [Google Scholar] [CrossRef]
  18. Zheng, Y.; Huang, Y.; Zheng, X.; Peng, J.; Chen, Y.; Yu, K.; Yang, Y.; Wang, X.; Yang, X.; Qian, J.; et al. Deaths from COPD in patients with cancer: A population-based study. Aging 2021, 13, 12641–12659. [Google Scholar] [CrossRef]
  19. Taucher, E.; Mykoliuk, I.; Lindenmann, J.; Smolle-Juettner, F.M. Implications of the Immune Landscape in COPD and Lung Cancer: Smoking Versus Other Causes. Front. Immunol. 2022, 13, 846605. [Google Scholar] [CrossRef]
  20. Wang, G.; Ma, A.; Zhang, L.; Guo, J.; Liu, Q.; Petersen, F.; Wang, Z.; Yu, X. Acute exacerbations of chronic obstructive pulmonary disease in a cohort of Chinese never smokers goes along with decreased risks of recurrent acute exacerbation, emphysema and comorbidity of lung cancer as well as decreased levels of circulating eosinophils and basophils. Front. Med. 2022, 9, 907893. [Google Scholar]
  21. de Alencar, V.T.L.; Figueiredo, A.B.; Corassa, M.; Gollob, K.J.; Cordeiro de Lima, V.C. Lung cancer in never smokers: Tumor immunology and challenges for immunotherapy. Front. Immunol. 2022, 13, 984349. [Google Scholar] [CrossRef]
  22. Zhao, G.; Li, X.; Lei, S.; Zhao, H.; Zhang, H.; Li, J. Prevalence of lung cancer in chronic obstructive pulmonary disease: A systematic review and meta-analysis. Front. Oncol. 2022, 12, 947981. [Google Scholar] [CrossRef]
  23. Park, H.Y.; Kang, D.; Shin, S.H.; Yoo, K.H.; Rhee, C.K.; Suh, G.Y.; Kim, H.; Shim, Y.M.; Guallar, E.; Cho, J.; et al. Chronic obstructive pulmonary disease and lung cancer incidence in never smokers: A cohort study. Thorax 2020, 75, 506–509. [Google Scholar] [CrossRef] [Green Version]
  24. Belkaid, Y.; Hand, T. Role of the Microbiota in Immunity and inflammation Yasmine. Early Hum. Dev. 2014, 157, 121–141. [Google Scholar]
  25. Ahn, S.V.; Lee, E.; Park, B.; Jung, J.H.; Park, J.E.; Sheen, S.S.; Park, K.J.; Hwang, S.C.; Park, J.B.; Park, H.S.; et al. Cancer development in patients with COPD: A retrospective analysis of the National Health Insurance Service-National Sample Cohort in Korea. BMC Pulm. Med. 2020, 20, 170. [Google Scholar] [CrossRef] [PubMed]
  26. Czarnecka-Chrebelska, K.H.; Mukherjee, D.; Maryanchik, S.V.; Rudzinska-Radecka, M. Biological and Genetic Mechanisms of COPD, Its Diagnosis, Treatment, and Relationship with Lung Cancer. Biomedicines 2023, 11, 448. [Google Scholar] [CrossRef] [PubMed]
  27. Forder, A.; Zhuang, R.; Souza, V.G.P.; Brockley, L.J.; Pewarchuk, M.E.; Telkar, N.; Stewart, G.L.; Benard, K.; Marshall, E.A.; Reis, P.P.; et al. Mechanisms Contributing to the Comorbidity of COPD and Lung Cancer. Int. J. Mol. Sci. 2023, 24, 2859. [Google Scholar] [CrossRef]
  28. Caramori, G.; Ruggeri, P.; Mumby, S.; Ieni, A.; Lo Bello, F.; Chaminka, V.; Donovan, C.; Andò, F.; Nucera, F.; Coppolino, I.; et al. Molecular links between COPD and lung cancer: New targets for drug discovery? Expert Opin. Ther. Targets 2019, 23, 539–553. [Google Scholar] [CrossRef] [PubMed]
  29. Durham, A.L.; Adcock, I.M. The relationship between COPD and lung cancer. Lung Cancer 2015, 90, 121–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Paudel, K.R.; Dharwal, V.; Patel, V.K.; Galvao, I.; Wadhwa, R.; Malyla, V.; Shen, S.S.; Budden, K.F.; Hansbro, N.G.; Vaughan, A.; et al. Role of Lung Microbiome in Innate Immune Response Associated With Chronic Lung Diseases. Front. Med. 2020, 7, 554. [Google Scholar] [CrossRef]
  31. Bou Zerdan, M.; Kassab, J.; Meouchy, P.; Haroun, E.; Nehme, R.; Bou Zerdan, M.; Fahed, G.; Petrosino, M.; Dutta, D.; Graziano, S. The Lung Microbiota and Lung Cancer: A Growing Relationship. Cancers 2022, 14, 4813. [Google Scholar] [CrossRef] [PubMed]
  32. Zhao, Y.; Liu, Y.; Li, S.; Peng, Z.; Liu, X.; Chen, J.; Zheng, X. Role of lung and gut microbiota on lung cancer pathogenesis. J. Cancer Res. Clin. Oncol. 2021, 147, 2177–2186. [Google Scholar] [CrossRef]
  33. Ito, N.; Tsujimoto, H.; Ueno, H.; Xie, Q.; Shinomiya, N. Helicobacter pylori-Mediated Immunity and Signaling Transduction in Gastric Cancer. J. Clin. Med. 2020, 9, 3699. [Google Scholar] [CrossRef] [PubMed]
  34. Mukherjee, A.; Sadler, P.J. Metals in Medicine: Therapeutic Agents. In Wiley Encyclopedia of Chemical Biology; Wiley: Hoboken, NJ, USA, 2009; pp. 1–47. [Google Scholar]
  35. Lemire, J.A.; Harrison, J.J.; Turner, R.J. Antimicrobial activity of metals: Mechanisms, molecular targets and applications. Nat. Rev. Microbiol. 2013, 11, 371–384. [Google Scholar] [CrossRef] [PubMed]
  36. Evans, A.; Kavanagh, K.A. Evaluation of metal-based antimicrobial compounds for the treatment of bacterial pathogens. J. Med. Microbiol. 2021, 70, 001363. [Google Scholar] [CrossRef] [PubMed]
  37. Turner, R.J. Metal-based antimicrobial strategies. Microb. Biotechnol. 2017, 10, 1062–1065. [Google Scholar] [CrossRef] [PubMed]
  38. Frei, A. Metal complexes, an untapped source of antibiotic potential? Antibiotics 2020, 9, 90. [Google Scholar] [CrossRef] [Green Version]
  39. Abate, C.; Carnamucio, F.; Giuffrè, O.; Foti, C. Metal-Based Compounds in Antiviral Therapy. Biomolecules 2022, 12, 933. [Google Scholar] [CrossRef]
  40. Claudel, M.; Schwarte, J.V.; Fromm, K.M. New Antimicrobial Strategies Based on Metal Complexes. Chemistry 2020, 2, 849–899. [Google Scholar] [CrossRef]
  41. Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. Third row transition metals for the treatment of cancer. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2015, 373, 20140185. [Google Scholar] [CrossRef] [Green Version]
  42. Anthony, E.J.; Bolitho, E.M.; Bridgewater, H.E.; Carter, O.W.L.; Donnelly, J.M.; Imberti, C.; Lant, E.C.; Lermyte, F.; Needham, R.J.; Palau, M.; et al. Metallodrugs are unique: Opportunities and challenges of discovery and development. Chem. Sci. 2020, 11, 12888–12917. [Google Scholar] [CrossRef]
  43. Lucaciu, R.L.; Hangan, A.C.; Sevastre, B.; Oprean, L.S. Metallo-Drugs in Cancer Therapy: Past, Present and Future. Molecules 2022, 27, 6485. [Google Scholar] [CrossRef]
  44. Diaz-Ochoa, V.E.; Jellbauer, S.; Klaus, S.; Raffatellu, M. Transition metal ions at the crossroads of mucosal immunity and microbial pathogenesis. Front. Cell. Infect. Microbiol. 2014, 4, 2. [Google Scholar] [CrossRef] [Green Version]
  45. Fong, W.; Li, Q.; Yu, J. Gut microbiota modulation: A novel strategy for prevention and treatment of colorectal cancer. Oncogene 2020, 39, 4925–4943. [Google Scholar] [CrossRef] [PubMed]
  46. Zhu, W.; Spiga, L.; Winter, S. Transition metals and host-microbe interactions in the inflamed intestine. BioMetals 2019, 32, 369–384. [Google Scholar] [CrossRef] [PubMed]
  47. She, P.; Zhou, L.; Li, S.; Liu, Y.; Xu, L.; Chen, L.; Luo, Z.; Wu, Y. Synergistic microbicidal effect of auranofin and antibiotics against planktonic and biofilm-encased S. aureus and E. faecalis. Front. Microbiol. 2019, 10, 2453. [Google Scholar] [CrossRef] [PubMed]
  48. Zou, L.; Wang, J.; Gao, Y.; Ren, X.; Rottenberg, M.E.; Lu, J.; Holmgren, A. Synergistic antibacterial activity of silver with antibiotics correlating with the upregulation of the ROS production. Sci. Rep. 2018, 8, 11131. [Google Scholar] [CrossRef] [Green Version]
  49. Abutaleb, N.S.; Seleem, M.N. Auranofin, at clinically achievable dose, protects mice and prevents recurrence from Clostridioides difficile infection. Sci. Rep. 2020, 10, 7701. [Google Scholar] [CrossRef]
  50. Weersma, R.K.; Zhernakova, A.; Fu, J. Interaction between drugs and the gut microbiome. Gut 2020, 69, 1510–1519. [Google Scholar] [CrossRef]
  51. Conti, G.; D’Amico, F.; Fabbrini, M.; Brigidi, P.; Barone, M.; Turroni, S. Pharmacomicrobiomics in Anticancer Therapies: Why the Gut Microbiota Should Be Pointed Out. Genes 2023, 14, 55. [Google Scholar] [CrossRef]
  52. Sommariva, M.; Le Noci, V.; Bianchi, F.; Camelliti, S.; Balsari, A.; Tagliabue, E.; Sfondrini, L. The lung microbiota: Role in maintaining pulmonary immune homeostasis and its implications in cancer development and therapy. Cell. Mol. Life Sci. 2020, 77, 2739–2749. [Google Scholar] [CrossRef] [Green Version]
  53. Moffatt, M.F.; Cookson, W.O. The Lung Microbiome in Health and Respiratory Diseases. Clin. Pulm. Med. 2017, 17, 525–529. [Google Scholar] [CrossRef]
  54. Pattaroni, C.; Watzenboeck, M.L.; Schneidegger, S.; Kieser, S.; Wong, N.C.; Bernasconi, E.; Pernot, J.; Mercier, L.; Knapp, S.; Nicod, L.P.; et al. Early-Life Formation of the Microbial and Immunological Environment of the Human Airways. Cell Host Microbe 2018, 24, 857–865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Bassis, C.M.; Erb-Downward, J.R.; Dickson, R.P.; Freeman, C.M.; Schmidt, T.M.; Young, V.B.; Beck, J.M.; Curtis, J.L.; Huffnagle, G.B. Analysis of the upper respiratory tract microbiotas as the source of the lung and gastric microbiotas in healthy individuals. mBio 2015, 6, 00037. [Google Scholar] [CrossRef] [Green Version]
  56. Huffnagle, G.B.; Dickson, R.P.; Lukacs, N.W. The respiratory tract microbiome and lung inflammation: A two-way street. Mucosal Immunol. 2017, 10, 299–306. [Google Scholar] [CrossRef] [Green Version]
  57. Hilty, M.; Burke, C.; Pedro, H.; Cardenas, P.; Bush, A.; Bossley, C.; Davies, J.; Ervine, A.; Poulter, L.; Pachter, L.; et al. Disordered microbial communities in asthmatic airways. PLoS ONE 2010, 5, 8578. [Google Scholar] [CrossRef] [Green Version]
  58. Yang, D.; Xing, Y.; Song, X.; Qian, Y. The impact of lung microbiota dysbiosis on inflammation. Immunology 2020, 159, 156–166. [Google Scholar] [CrossRef] [Green Version]
  59. Whiteside, S.A.; McGinniss, J.E.; Collman, R.G. The lung microbiome: Progress and promise. J. Clin. Investig. 2021, 131, e150473. [Google Scholar] [CrossRef] [PubMed]
  60. Yagi, K.; Huffnagle, G.B.; Lukacs, N.W.; Asai, N. The lung microbiome during health and disease. Int. J. Mol. Sci. 2021, 22, 10872. [Google Scholar] [CrossRef] [PubMed]
  61. Yi, X.; Gao, J.; Wang, Z. The human lung microbiome—A hidden link between microbes and human health and diseases. iMeta 2022, 1, 33. [Google Scholar] [CrossRef]
  62. Sze, M.A.; Hogg, J.C.; Sin, D.D. Bacterial microbiome of lungs in COPD. Int. J. COPD 2014, 9, 229–238. [Google Scholar]
  63. Ramsheh, M.Y.; Haldar, K.; Esteve-Codina, A.; Purser, L.F.; Richardson, M.; Müller-Quernheim, J.; Greulich, T.; Nowinski, A.; Barta, I.; Stendardo, M.; et al. Lung microbiome composition and bronchial epithelial gene expression in patients with COPD versus healthy individuals: A bacterial 16S rRNA gene sequencing and host transcriptomic analysis. Lancet Microbe 2021, 2, 300–310. [Google Scholar] [CrossRef] [PubMed]
  64. Whelan, F.J.; Heirali, A.A.; Rossi, L.; Rabin, H.R.; Parkins, M.D.; Surette, M.G. Longitudinal sampling of the lung microbiota in individuals with cystic fibrosis. PLoS ONE 2017, 12, 0172811. [Google Scholar] [CrossRef] [Green Version]
  65. Cox, M.J.; Turek, E.M.; Hennessy, C.; Mirza, G.K.; James, P.L.; Coleman, M.; Jones, A.; Wilson, R.; Bilton, D.; Cookson, W.O.C.; et al. Longitudinal assessment of sputum microbiome by sequencing of the 16S rRNA gene in non-cystic fibrosis bronchiectasis patients. PLoS ONE 2017, 12, 0170622. [Google Scholar] [CrossRef] [Green Version]
  66. Hong, B.; Paulson, J.N.; Stine, O.C.; Weinstock, G.M.; Cervantes, J.L. Meta-analysis of the lung microbiota in pulmonary tuberculosis. Tuberculosis 2018, 109, 102–108. [Google Scholar] [CrossRef]
  67. Tsay, J.C.J.; Wu, B.G.; Badri, M.H.; Clemente, J.C.; Shen, N.; Meyn, P.; Li, Y.; Yie, T.A.; Lhakhang, T.; Olsen, E.; et al. Airway microbiota is associated with upregulation of the PI3K pathway in lung cancer. Am. J. Respir. Crit. Care Med. 2018, 198, 1188–1198. [Google Scholar] [CrossRef]
  68. Lozupone, C.A.; Hamady, M.; Kelley, S.T.; Knight, R. Quantitative and qualitative β diversity measures lead to different insights into factors that structure microbial communities. Appl. Environ. Microbiol. 2007, 73, 1576–1585. [Google Scholar] [CrossRef] [Green Version]
  69. Lozupone, C.A.; Knight, R. Species divergence and the measurement of microbial diversity. FEMS Microbiol. Rev. 2008, 32, 557–578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Kriss, M.; Hazleton, K.Z.; Nusbacherd, N.M.; Martine, C.G.; Lozuponed, C.A. Low Diversity Gut Microbiota Dysbiosis: Drivers, Functional Implications and Recovery. Curr Opin Microbiol 2018, 44, 34–40. [Google Scholar] [CrossRef]
  71. Segal, L.N.; Clemente, J.C.; Tsay, J.C.J.; Koralov, S.B.; Keller, B.C.; Wu, B.G.; Li, Y.; Shen, N.; Ghedin, E.; Morris, A.; et al. Enrichment of the lung microbiome with oral taxa is associated with lung inflammation of a Th17 phenotype. Nat. Microbiol. 2016, 1, 16031. [Google Scholar] [CrossRef] [Green Version]
  72. Gustafson, A.M.; Soldi, R.; Anderlind, C.; Scholand, M.B.; Qian, J.; Zhang, X.; Cooper, K.; Walker, D.; Mcwilliams, A.; Gang, L.; et al. Airway PI3K pathway activation is an early and reversible event in lung cancer development. Sci. Transl. Med. 2010, 2, 2625. [Google Scholar] [CrossRef] [Green Version]
  73. Jungnickel, C.; Schmidt, L.H.; Bittigkoffer, L.; Wolf, L.; Wolf, A.; Ritzmann, F.; Kamyschnikow, A.; Herr, C.; Menger, M.D.; Spieker, T.; et al. IL-17C mediates the recruitment of tumor-associated neutrophils and lung tumor growth. Oncogene 2017, 36, 4182–4190. [Google Scholar] [CrossRef] [PubMed]
  74. Ritzmann, F.; Lunding, L.P.; Bals, R.; Wegmann, M.; Beisswenger, C. IL-17 Cytokines and Chronic Lung Diseases. Cells 2022, 11, 2132. [Google Scholar] [CrossRef]
  75. Segal, L.N.; Clemente, J.C.; Li, Y.; Ruan, C.; Cao, J.; Danckers, M.; Morris, A.; Tapyrik, S.; Wu, B.G.; Diaz, P.; et al. Anaerobic Bacterial Fermentation Products Increase Tuberculosis Risk in Antiretroviral-Drug-Treated HIV Patients. Cell Host Microbe 2017, 21, 530–537. [Google Scholar] [CrossRef] [Green Version]
  76. Nazir, S.A.; Erbland, M.L. Chronic Obstructive Pulmonary Disease An Update on Diagnosis and Management Issues in Older Adults. Drugs Aging Vol. 2009, 26, 813–831. [Google Scholar] [CrossRef] [PubMed]
  77. Hobbs, B.D.; Morrow, J.D.; Wang, X.W.; Liu, Y.Y.; DeMeo, D.L.; Hersh, C.P.; Celli, B.R.; Bueno, R.; Criner, G.J.; Silverman, E.K.; et al. Identifying chronic obstructive pulmonary disease from integrative omics and clustering in lung tissue. BMC Pulm. Med. 2023, 23, 115. [Google Scholar] [CrossRef]
  78. Kim, V.; Criner, G.J. Chronic bronchitis and chronic obstructive pulmonary disease. Am. J. Respir. Crit. Care Med. 2013, 187, 228–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Kim, V.; Criner, G.J. The Chronic Bronchitis Phenotype in COPD: Features and Implications Victor. Curr. Opin. Pulm. Med. 2016, 21, 133–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Erb-Downward, J.R.; Thompson, D.L.; Han, M.K.; Freeman, C.M.; McCloskey, L.; Schmidt, L.A.; Young, V.B.; Toews, G.B.; Curtis, J.L.; Sundaram, B.; et al. Analysis of the lung microbiome in the “healthy” smoker and in COPD. PLoS ONE 2011, 6, 16384. [Google Scholar] [CrossRef] [Green Version]
  81. Pragman, A.A.; Kim, H.B.; Reilly, C.S.; Wendt, C.; Isaacson, R.E. The Lung Microbiome in Moderate and Severe Chronic Obstructive Pulmonary Disease. PLoS ONE 2012, 7, 47305. [Google Scholar] [CrossRef] [Green Version]
  82. Han, M.L.K.; Huang, Y.J.; LiPuma, J.J.; Boushey, H.A.; Boucher, R.C.; Cookson, W.O.; Curtis, J.L.; Erb-Downward, J.; Lynch, S.V.; Sethi, S.; et al. Significance of the microbiome in obstructive lung disease. Thorax 2012, 67, 456–463. [Google Scholar] [CrossRef] [Green Version]
  83. Larsen, J.M.; Musavian, H.S.; Butt, T.M.; Ingvorsen, C.; Thysen, A.H.; Brix, S. Chronic obstructive pulmonary disease and asthma-associated Proteobacteria, but not commensal Prevotella spp., promote Toll-like receptor 2-independent lung inflammation and pathology. Immunology 2014, 144, 333–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Gaeckle, N.T.; Pragman, A.A.; Pendleton, K.M.; Baldomero, A.K.; Criner, G.J. The oral-lung axis: The impact of oral health on lung health. Respir. Care 2020, 65, 1211–1220. [Google Scholar] [CrossRef]
  85. Garcia-Nuñez, M.; Marti, S.; Puig, C.; Perez-Brocal, V.; Millares, L.; Santos, S.; Ardanuy, C.; Moya, A.; Linãres, J.; Monsó, E. Bronchial microbiome, PA biofilm-forming capacity and exacerbation in severe COPD patients colonized by P. aeruginosa. Future Microbiol. 2017, 12, 379–392. [Google Scholar] [CrossRef]
  86. Wang, Z.; Bafadhel, M.; Haldar, K.; Spivak, A.; Mayhew, D.; Miller, B.E.; Tal-Singer, R.; Johnston, S.L.; Ramsheh, M.Y.; Barer, M.R.; et al. Lung microbiome dynamics in COPD exacerbations. Eur. Respir. J. 2016, 47, 1082–1092. [Google Scholar] [CrossRef] [Green Version]
  87. Dickson, R.P.; Erb-Downward, J.R.; Freeman, C.M.; McCloskey, L.; Beck, J.M.; Huffnagle, G.B.; Curtis, J.L. Spatial variation in the healthy human lung microbiome and the adapted island model of lung biogeography. Ann. Am. Thorac. Soc. 2015, 12, 821–830. [Google Scholar] [CrossRef] [Green Version]
  88. Xue, Q.; Xie, Y.; He, Y.; Yu, Y.; Fang, G.; Yu, W.; Wu, J.; Li, J.; Zhao, L.; Deng, X.; et al. Lung microbiome and cytokine profiles in different disease states of COPD: A cohort study. Sci. Rep. 2023, 13, 5715. [Google Scholar] [CrossRef]
  89. Dicker, A.J.; Huang, J.T.J.; Lonergan, M.; Keir, H.R.; Fong, C.J.; Tan, B.; Cassidy, A.J.; Finch, S.; Mullerova, H.; Miller, B.E.; et al. The sputum microbiome, airway inflammation, and mortality in chronic obstructive pulmonary disease. J. Allergy Clin. Immunol. 2021, 147, 158–167. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, Z.; Singh, R.; Miller, B.E.; Tal-Singer, R.; Van Horn, S.; Tomsho, L.; MacKay, A.; Allinson, J.P.; Webb, A.J.; Brookes, A.J.; et al. Sputum microbiome temporal variability and dysbiosis in chronic obstructive pulmonary disease exacerbations: An analysis of the COPDMAP study. Thorax 2018, 73, 331–338. [Google Scholar] [CrossRef] [Green Version]
  91. Liu, H.; Zheng, D.; Lin, Y.; Liu, Z.; Liang, Z.; Su, J.; Chen, R.; Zhou, H.; Wang, Z. Association of sputum microbiome with clinical outcome of initial antibiotic treatment in hospitalized patients with acute exacerbations of COPD. Pharmacol. Res. 2020, 160, 105095. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, Z.; Yang, Y.; Yan, Z.; Liu, H.; Chen, B.; Liang, Z.; Wang, F.; Miller, B.E.; Tal-Singer, R.; Yi, X.; et al. Multi-omic meta-analysis identifies functional signatures of airway microbiome in chronic obstructive pulmonary disease. ISME J. 2020, 14, 2748–2765. [Google Scholar] [CrossRef]
  93. MacLeod, M.; Papi, A.; Contoli, M.; Beghé, B.; Celli, B.R.; Wedzicha, J.A.; Fabbri, L.M. Chronic obstructive pulmonary disease exacerbation fundamentals: Diagnosis, treatment, prevention and disease impact. Respirology 2021, 26, 532–551. [Google Scholar] [CrossRef] [PubMed]
  94. Fishbein, S.R.S.; Mahmud, B.; Dantas, G. Antibiotic perturbations to the gut microbiome. Nat. Rev. Microbiol. 2023. [Google Scholar] [CrossRef] [PubMed]
  95. Ramirez, J.; Guarner, F.; Bustos Fernandez, L.; Maruy, A.; Sdepanian, V.L.; Cohen, H. Antibiotics as Major Disruptors of Gut Microbiota. Front. Cell. Infect. Microbiol. 2020, 10, 572912. [Google Scholar] [CrossRef] [PubMed]
  96. Yang, L.; Bajinka, O.; Jarju, P.O.; Tan, Y.; Taal, A.M.; Ozdemir, G. The varying effects of antibiotics on gut microbiota. AMB Express 2021, 11, 116. [Google Scholar] [CrossRef] [PubMed]
  97. Anthony, W.E.; Wang, B.; Sukhum, K.V.; D’Souza, A.W.; Hink, T.; Cass, C.; Seiler, S.; Reske, K.A.; Coon, C.; Dubberke, E.R.; et al. Acute and persistent effects of commonly used antibiotics on the gut microbiome and resistome in healthy adults. Cell Rep. 2022, 39, 110649. [Google Scholar] [CrossRef]
  98. Dickson, R.; Martinez, F.; Huffnagle, G. The Role of the Microbiome in Exacerbations of Chronic Lung Diseases. Hosp. Peer Rev. 2014, 384, 691–702. [Google Scholar] [CrossRef] [Green Version]
  99. Assefa, M. Multi-drug resistant gram-negative bacterial pneumonia: Etiology, risk factors, and drug resistance patterns. Pneumonia 2022, 14, 4. [Google Scholar] [CrossRef]
  100. Merker, M.; Tueffers, L.; Vallier, M.; Groth, E.E.; Sonnenkalb, L.; Unterweger, D.; Baines, J.F.; Niemann, S.; Schulenburg, H. Evolutionary Approaches to Combat Antibiotic Resistance: Opportunities and Challenges for Precision Medicine. Front. Immunol. 2020, 11, 1938. [Google Scholar] [CrossRef]
  101. Toraldo, D.M.; Conte, L. Influence of the Lung Microbiota Dysbiosis in Chronic Obstructive Pulmonary Disease Exacerbations: The Controversial Use of Corticosteroid and Antibiotic Treatments and the Role of Eosinophils as a Disease Marker. J. Clin. Med. Res. 2019, 11, 667–675. [Google Scholar] [CrossRef] [Green Version]
  102. Ren, L.; Zhang, R.; Rao, J.; Xiao, Y.; Zhang, Z.; Yang, B.; Cao, D.; Zhong, H.; Ning, P.; Shang, Y.; et al. Transcriptionally Active Lung Microbiome and Its Association with Bacterial Biomass and Host Inflammatory Status. mSystems 2018, 3, 00199-18. [Google Scholar] [CrossRef] [Green Version]
  103. Von, S.T.; Seng, H.L.; Lee, H.B.; Ng, S.W.; Kitamura, Y.; Chikira, M.; Ng, C.H. DNA molecular recognition and cellular selectivity of anticancer metal(II) complexes of ethylenediaminediacetate and phenanthroline: Multiple targets. J. Biol. Inorg. Chem. 2012, 17, 57–69. [Google Scholar] [CrossRef] [PubMed]
  104. Pragman, A.A.; Knutson, K.A.; Gould, T.J.; Isaacson, R.E.; Reilly, C.S.; Wendt, C.H. Chronic obstructive pulmonary disease upper airway microbiota alpha diversity is associated with exacerbation phenotype: A case-control observational study. Respir. Res. 2019, 20, 114. [Google Scholar] [CrossRef] [Green Version]
  105. Sze, M.A.; Dimitriu, P.A.; Hayashi, S.; Elliott, W.M.; McDonough, J.E.; Gosselink, J.V.; Cooper, J.; Sin, D.D.; Mohn, W.W.; Hogge, J.C. The lung tissue microbiome in chronic obstructive pulmonary disease. Am. J. Respir. Crit. Care Med. 2012, 185, 1073–1080. [Google Scholar] [CrossRef] [Green Version]
  106. Zhu, K.; Zhou, S.; Xu, A.; Sun, L.; Li, M.; Jiang, H.; Zhang, B.; Zeng, D.; Fei, G.; Wang, R. Microbiota Imbalance Contributes to COPD Deterioration by Enhancing IL-17a Production via miR-122 and miR-30a. Mol. Ther. Nucleic Acids 2020, 22, 520–529. [Google Scholar] [CrossRef]
  107. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Parkin, D.M.; Piñeros, M.; Znaor, A.; Bray, F. Cancer statistics for the year 2020: An overview. Int. J. Cancer 2021, 149, 778–789. [Google Scholar] [CrossRef] [PubMed]
  108. Anusewicz, D.; Orzechowska, M.; Bednarek, A.K. Lung squamous cell carcinoma and lung adenocarcinoma differential gene expression regulation through pathways of Notch, Hedgehog, Wnt, and ErbB signalling. Sci. Rep. 2020, 10, 21128. [Google Scholar] [CrossRef]
  109. Testa, U.; Castelli, G.; Pelosi, E. Lung cancers: Molecular characterization, clonal heterogeneity and evolution, and cancer stem cells. Cancers 2018, 10, 248. [Google Scholar] [CrossRef]
  110. Chen, Z.; Fillmore, C.M.; Hammerman, P.S.; Kim, C.F.; Wong, K.-K. Non-small-cell lung cancers: A heterogeneous set of diseases. Nat. Rev. Cancer 2014, 14, 535–546. [Google Scholar] [PubMed]
  111. Eapen, M.S.; Hansbro, P.M.; Larsson-Callerfelt, A.K.; Jolly, M.K.; Myers, S.; Sharma, P.; Jones, B.; Rahman, M.A.; Markos, J.; Chia, C.; et al. Chronic Obstructive Pulmonary Disease and Lung Cancer: Underlying Pathophysiology and New Therapeutic Modalities. Drugs 2018, 78, 1717–1740. [Google Scholar] [CrossRef]
  112. Gomes, S.; Cavadas, B.; Ferreira, J.C.; Marques, P.I.; Monteiro, C.; Sucena, M.; Sousa, C.; Vaz Rodrigues, L.; Teixeira, G.; Pinto, P.; et al. Profiling of lung microbiota discloses differences in adenocarcinoma and squamous cell carcinoma. Sci. Rep. 2019, 9, 12838. [Google Scholar] [CrossRef] [Green Version]
  113. Yan, X.; Yang, M.; Liu, J.; Gao, R.; Hu, J.; Li, J.; Zhang, L.; Shi, Y.; Guo, H.; Cheng, J.; et al. Discovery and validation of potential bacterial biomarkers for lung cancer. Am. J. Cancer Res. 2015, 5, 3111–3122. [Google Scholar]
  114. Lee, S.H.; Sung, J.Y.; Yong, D.; Chun, J.; Kim, S.Y.S.K.; Song, J.H.; Chung, K.S.; Kim, E.Y.; Jung, J.Y.; Kang, Y.A.; et al. Characterization of microbiome in bronchoalveolar lavage fluid of patients with lung cancer comparing with benign mass like lesions. Lung Cancer 2016, 102, 89–95. [Google Scholar] [CrossRef] [PubMed]
  115. Apopa, P.L.; Alley, L.; Penney, R.B.; Arnaoutakis, K.; Steliga, M.A.; Jeffus, S.; Bircan, E.; Gopalan, B.; Jin, J.; Patumcharoenpol, P.; et al. PARP1 is up-regulated in non-small cell lung cancer tissues in the presence of the Cyanobacterial toxin microcystin. Front. Microbiol. 2018, 9, 1757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. González, I.; Araya, P.; Roj, A. Helicobacter pylori infection and lung cancer: New insights and future challenges. Chin. J. Lung Cancer 2018, 21, 658–662. [Google Scholar]
  117. Xu, N.; Wang, L.; Li, C.; Ding, C.; Li, C.; Fan, W.; Cheng, C.; Gu, B. Microbiota dysbiosis in lung cancer: Evidence of association and potential mechanisms. Transl. Lung Cancer Res. 2020, 9, 1554–1568. [Google Scholar] [CrossRef] [PubMed]
  118. Liu, N.N.; Ma, Q.; Ge, Y.; Yi, C.X.; Wei, L.Q.; Tan, J.C.; Chu, Q.; Li, J.Q.; Zhang, P.; Wang, H. Microbiome dysbiosis in lung cancer: From composition to therapy. npj Precis. Oncol. 2020, 4, 33. [Google Scholar] [CrossRef]
  119. Karakasidis, E.; Kotsiou, O.S.; Gourgoulianis, K.I. Lung and Gut Microbiome in COPD. J. Pers. Med. 2023, 13, 804. [Google Scholar] [CrossRef]
  120. He, J.Q.; Chen, Q.; Wu, S.J.; Wang, D.Q.; Zhang, S.Y.; Zhang, S.Z.; Chen, R.L.; Wang, J.F.; Wang, Z.; Yu, C.H. Potential Implications of the Lung Microbiota in Patients with Chronic Obstruction Pulmonary Disease and Non-Small Cell Lung Cancer. Front. Cell. Infect. Microbiol. 2022, 12, 937864. [Google Scholar] [CrossRef]
  121. Yu, G.; Gail, M.H.; Consonni, D.; Carugno, M.; Humphrys, M.; Pesatori, A.C.; Caporaso, N.E.; Goedert, J.J.; Ravel, J.; Landi, M.T. Characterizing human lung tissue microbiota and its relationship to epidemiological and clinical features. Genome Biol. 2016, 17, 163. [Google Scholar] [CrossRef] [Green Version]
  122. Huang, D.; Su, X.; Yuan, M.; Zhang, S.; Jing, H.; Deng, Q.; Qiu, W.; Dong, H.; Cai, S. The Characterization of Lung Microbiome in Sputum of Lung Cancer Patients with Different Clinicopathology. Am. J. Cancer Res. 2019, 9, 2047–2063. [Google Scholar]
  123. Najafi, S.; Abedini, F.; Azimzadeh Jamalkandi, S.; Shariati, P.; Ahmadi, A.; Gholami Fesharaki, M. The composition of lung microbiome in lung cancer: A systematic review and meta-analysis. BMC Microbiol. 2021, 21, 315. [Google Scholar] [CrossRef] [PubMed]
  124. Johansson, M.E.V.; Jakobsson, H.E.; Holmén-Larsson, J.; Schütte, A.; Ermund, A.; Rodríguez-Piñeiro, A.M.; Arike, L.; Wising, C.; Svensson, F.; Bäckhed, F.; et al. Normalization of host intestinal mucus layers requires long-term microbial colonization. Cell Host Microbe. 2015, 18, 582–592. [Google Scholar] [CrossRef] [Green Version]
  125. Zhuang, H.; Cheng, L.; Wang, Y.; Zhang, Y.K.; Zhao, M.F.; Liang, G.D.; Zhang, M.C.; Li, Y.G.; Zhao, J.B.; Gao, Y.N.; et al. Dysbiosis of the gut microbiome in lung cancer. Front. Cell. Infect. Microbiol. 2019, 9, 112. [Google Scholar] [CrossRef]
  126. Wang, Z.; Bai, C.; Hu, T.; Luo, C.; Yu, H.; Ma, X.; Liu, T.; Gu, X. Emerging trends and hotspot in gut–lung axis research from 2011 to 2021: A bibliometrics analysis. Biomed. Eng. Online 2022, 21, 27. [Google Scholar] [CrossRef] [PubMed]
  127. Enaud, R.; Prevel, R.; Ciarlo, E.; Beaufils, F.; Wieërs, G.; Guery, B.; Delhaes, L. The Gut-Lung Axis in Health and Respiratory Diseases: A Place for Inter-Organ and Inter-Kingdom Crosstalks. Front. Cell. Infect. Microbiol. 2020, 10, 9. [Google Scholar] [CrossRef] [Green Version]
  128. Jin, C.; Lagoudas, G.K.; Zhao, C.; Bullman, S.; Bhutkar, A.; Hu, B.; Ameh, S.; Sandel, D.; Liang, X.S.; Mazzilli, S.; et al. Commensal Microbiota Promote Lung Cancer Development via γδ T Cells. Cell 2019, 176, 998–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Zhang, H.; García Rodríguez, L.A.; Hernández-Díaz, S. Antibiotic use and the risk of lung cancer. Cancer Epidemiol. Biomarkers Prev. 2008, 17, 1308–1315. [Google Scholar] [CrossRef] [Green Version]
  130. Francescone, R.; Hou, V.; Grivennikov, S.I. Microbiome, inflammation, and cancer. Cancer J. 2014, 20, 181–189. [Google Scholar] [CrossRef] [Green Version]
  131. Sadrekarimi, H.; Gardanova, Z.R.; Bakhshesh, M.; Ebrahimzadeh, F.; Yaseri, A.F.; Thangavelu, L.; Hasanpoor, Z.; Zadeh, F.A.; Kahrizi, M.S. Emerging role of human microbiome in cancer development and response to therapy: Special focus on intestinal microflora. J. Transl. Med. 2022, 20, 301. [Google Scholar] [CrossRef]
  132. Hou, X.; Zheng, Z.; Wei, J.; Zhao, L. Effects of gut microbiota on immune responses and immunotherapy in colorectal cancer. Front. Immunol. 2022, 13, 1030745. [Google Scholar] [CrossRef] [PubMed]
  133. Sheflin, A.M.; Whitney, A.K.; Weir, T.L. Cancer-Promoting Effects of Microbial Dysbiosis. Curr. Oncol. Rep. 2014, 16, 406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Singh, S.; Sharma, P.; Sarma, D.K.; Kumawat, M.; Tiwari, R.; Verma, V.; Nagpal, R.; Kumar, M. Implication of Obesity and Gut Microbiome Dysbiosis in the Etiology of Colorectal Cancer. Cancers 2023, 15, 1913. [Google Scholar] [CrossRef] [PubMed]
  135. Mills, K.H.G. IL-17 and IL-17-producing cells in protection versus pathology. Nat. Rev. Immunol. 2023, 23, 38–54. [Google Scholar] [CrossRef]
  136. Rivas-Domínguez, A.; Pastor, N.; Martínez-López, L.; Colón-Pérez, J.; Bermúdez, B.; Orta, M.L. The role of dna damage response in dysbiosis-induced colorectal cancer. Cells 2021, 10, 1934. [Google Scholar] [CrossRef]
  137. Wang, Z.; Yang, J.; Qi, J.; Jin, Y.; Tong, L. Activation of NADPH/ROS pathway contributes to angiogenesis through JNK signaling in brain endothelial cells. Microvasc. Res. 2020, 131, 104012. [Google Scholar] [CrossRef]
  138. Ochoa Perez, C.E.; Mirabolfathinejad, S.G.; Venado, A.R.; Evans, S.E.; Gagea, M.; Evans, C.M.; Dickey, B.F.; Moghaddam, S.J. Interleukin 6, but not T helper 2 cytokines, promotes lung carcinogenesis. Cancer Prev. Res. 2011, 4, 51–64. [Google Scholar] [CrossRef] [Green Version]
  139. Caetano, M.S.; Zhang, H.; Cumpian, A.M.; Gong, L.; Unver, N.; Ostrin, E.J.; Daliri, S.; Chang, S.H.; Ochoa, C.E.; Hanash, S.; et al. IL-6 blockade reprograms the lung tumor microenvironment to limit the development and progression of K-ras mutant lung cancer. Cancer Res. 2016, 76, 3189–3199. [Google Scholar] [CrossRef] [Green Version]
  140. Chang, S.H.; Mirabolfathinejad, S.G.; Katta, H.; Cumpian, A.M.; Gong, L.; Caetano, M.S.; Moghaddam, S.J.; Dong, C. T helper 17 cells play a critical pathogenic role in lung cancer. Proc. Natl. Acad. Sci. USA 2014, 111, 5664–5669. [Google Scholar] [CrossRef]
  141. Weeks, J.R.; Staples, K.J.; Spalluto, C.M.; Watson, A.; Wilkinson, T.M.A. The Role of Non-Typeable Haemophilus influenzae Biofilms in Chronic Obstructive Pulmonary Disease. Front. Cell. Infect. Microbiol. 2021, 11, 720742. [Google Scholar] [CrossRef]
  142. Sriram, K.B.; Cox, A.J.; Sivakumaran, P.; Singh, M.; Watts, A.M.; West, N.P.; Cripps, A.W. Non-typeable Haemophilus influenzae detection in the lower airways of patients with lung cancer and chronic obstructive pulmonary disease. Multidiscip. Respir. Med. 2018, 13, 11. [Google Scholar] [CrossRef] [Green Version]
  143. Su, Y.C.; Jalalvand, F.; Thegerström, J.; Riesbeck, K. The interplay between immune response and bacterial infection in COPD: Focus Upon non-typeable Haemophilus influenzae. Front. Immunol. 2018, 9, 2530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Nobili, S.; Mini, E.; Landini, I.; Gabbiani, C.; Casini, A.; Messori, L. Gold compounds as anticancer agents: Chemistry, cellular pharmacology, and preclinical studies. Med. Res. Rev. 2009, 30, 550–580. [Google Scholar] [CrossRef] [PubMed]
  145. Dilda, P.J.; Hogg, P.J. Arsenical-based cancer drugs. Cancer Treat. Rev. 2007, 33, 542–564. [Google Scholar] [CrossRef]
  146. Klasen, H.J. A historical review of the use of silver in the treatment of burns. II. Renewed interest for silver. Burns 2000, 26, 131–138. [Google Scholar] [CrossRef]
  147. Ehrlich, P. About Salvarsan. Abhandlungen über Salvarsan 1912, 2, 547–563. [Google Scholar]
  148. Ehrlich, P.; Bertheim, A. On the hydrochloric acid 3,3′-diamino-4,4′-dioxyarsenobenzene and its closest relatives. Rep. Ger. Chem. Soc. 1912, 45, 756–766. [Google Scholar]
  149. Rosenberg, B.; Van Camp, L.; Krigas, T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature 1965, 205, 698–699. [Google Scholar] [CrossRef]
  150. Dasari, S.; Tchounwou, P.B. Cisplatin in cancer therapy: Molecular mechanisms of action. Eur. J. Pharmacol. 2014, 740, 364–378. [Google Scholar] [CrossRef] [Green Version]
  151. Aldossary, S.A. Review on pharmacology of cisplatin: Clinical use, toxicity and mechanism of resistance of cisplatin. Biomed. Pharmacol. J. 2019, 12, 7–15. [Google Scholar] [CrossRef]
  152. Makovník, M.; Rejleková, K.; Uhrin, I.; Mego, M.; Chovanec, M. Intricacies of Radiographic Assessment in Testicular Germ Cell Tumors. Front. Oncol. 2021, 10, 587523. [Google Scholar] [CrossRef]
  153. Nieder, C.; Pawinski, A.; Andratschke, N.H. Combined radio- and chemotherapy for non-small cell lung cancer: Systematic review of landmark studies based on acquired citations. Front. Oncol. 2013, 3, 176. [Google Scholar] [CrossRef] [Green Version]
  154. Tchounwou, P.B.; Dasari, S.; Noubissi, F.K.; Ray, P.; Kumar, S. Advances in our understanding of the molecular mechanisms of action of cisplatin in cancer therapy. J. Exp. Pharmacol. 2021, 13, 303–328. [Google Scholar] [CrossRef]
  155. Cocetta, V.; Ragazzi, E.; Montopoli, M. Mitochondrial involvement in cisplatin resistance. Int. J. Mol. Sci. 2019, 20, 3384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Kryczka, J.; Kryczka, J.; Czarnecka-Chrebelska, K.H.; Brzeziańska-Lasota, E. Molecular mechanisms of chemoresistance induced by cisplatin in NSCLC cancer therapy. Int. J. Mol. Sci. 2021, 22, 8885. [Google Scholar] [CrossRef]
  157. Dasari, S.; Njiki, S.; Mbemi, A.; Yedjou, C.G.; Tchounwou, P.B. Pharmacological Effects of Cisplatin Combination with Natural Products in Cancer Chemotherapy. Int. J. Mol. Sci. 2022, 23, 1532. [Google Scholar] [CrossRef] [PubMed]
  158. Iacobucci, I.; La Manna, S.; Cipollone, I.; Can, L.; Cozzolino, F. From the Discovery of Targets to Delivery Systems: How to Decipher and Improve the Metallodrugs’ Actions at a Molecular Level. Pharm. Rev. 2023, 15, 1997. [Google Scholar] [CrossRef]
  159. Gou, Y.; Liu, L.; Liang, H. The developments of metal-based agents against lung cancer Yi. Front. Pharmacol. 2022, 13, 5106–5131. [Google Scholar] [CrossRef]
  160. Riccardi, C.; Piccolo, M. Metal-Based Complexes in Cancer. Int. J. Mol. Sci. 2023, 24, 16–18. [Google Scholar] [CrossRef] [PubMed]
  161. Sharma, A.; Shambhwani, D.; Pandey, S.; Singh, J.; Lalhlenmawia, H.; Kumarasamy, M.; Singh, S.K.; Chellappan, D.K.; Gupta, G.; Prasher, P.; et al. Advances in Lung Cancer Treatment Using Nanomedicines. ACS Omega 2022, 8, 10–41. [Google Scholar] [CrossRef]
  162. Guo, Q.; Liu, L.; Chen, Z.; Fan, Y.; Zhou, Y.; Yuan, Z.; Zhang, W. Current treatments for non-small cell lung cancer. Front. Oncol. 2022, 12, 945102. [Google Scholar] [CrossRef]
  163. Kelland, L. The resurgence of platinum-based cancer chemotherapy. Nat. Rev. Cancer 2007, 7, 573–584. [Google Scholar] [CrossRef]
  164. Zhang, C.; Leighl, N.B.; Wu, Y.L.; Zhong, W.Z. Emerging therapies for non-small cell lung cancer. J. Hematol. Oncol. 2019, 12, 45. [Google Scholar] [CrossRef] [Green Version]
  165. Anand, U.; Dey, A.; Chandel, A.K.S.; Sanyal, R.; Mishra, A.; Pandey, D.K.; De Falco, V.; Upadhyay, A.; Kandimalla, R.; Chaudhary, A.; et al. Cancer chemotherapy and beyond: Current status, drug candidates, associated risks and progress in targeted therapeutics. Genes Dis. 2023, 10, 1367–1401. [Google Scholar] [CrossRef]
  166. Alessio, E.; Messori, L. NAMI-A and KP1019/1339, two iconic ruthenium anticancer drug candidates face-to-face: A case story in medicinal inorganic chemistry. Molecules 2019, 24, 1995. [Google Scholar] [CrossRef] [Green Version]
  167. Leijen, S.; Burgers, S.A.; Baas, P.; Pluim, D.; Tibben, M.; Van Werkhoven, E.; Alessio, E.; Sava, G.; Beijnen, J.H.; Schellens, J.H.M. Phase I/II study with ruthenium compound NAMI-A and gemcitabine in patients with non-small cell lung cancer after first line therapy. Investig. New Drugs 2015, 33, 201–214. [Google Scholar] [CrossRef]
  168. Lee, S.Y.; Kim, C.Y.; Nam, T.G. Ruthenium complexes as anticancer agents: A brief history and perspectives. Drug Des. Devel. Ther. 2020, 14, 5375–5392. [Google Scholar] [CrossRef] [PubMed]
  169. Abdalbari, F.H.; Telleria, C.M. The gold complex auranofin: New perspectives for cancer therapy. Discov. Oncol. 2021, 12, 42. [Google Scholar] [CrossRef]
  170. Gordon, E.M.; Angel, N.L.; Omelchenko, N.; Chua-Alcala, V.S.; Moradkhani, A.; Quon, D.; Wong, S. A Phase I/II Investigation of Safety and Efficacy of Nivolumab and nab-Sirolimus in Patients with a Variety of Tumors with Genetic Mutations in the mTOR Pathway. Anticancer Res. 2023, 43, 1993–2002. [Google Scholar] [CrossRef] [PubMed]
  171. Molinaro, C.; Martoriati, A.; Pelinski, L.; Cailliau, K. Copper complexes as anticancer agents targeting topoisomerases i and ii. Cancers 2020, 12, 2863. [Google Scholar] [CrossRef]
  172. Liu, Y.L.; Bager, C.L.; Willumsen, N.; Ramchandani, D.; Kornhauser, N.; Ling, L.; Cobham, M.; Andreopoulou, E.; Cigler, T.; Moore, A.; et al. Tetrathiomolybdate (TM)-associated copper depletion influences collagen remodeling and immune response in the pre-metastatic niche of breast cancer. npj Breast Cancer 2021, 7, 108. [Google Scholar] [CrossRef] [PubMed]
  173. Guo, Q.; Li, L.; Hou, S.; Yuan, Z.; Li, C.; Zhang, W.; Zheng, L.; Li, X. The Role of Iron in Cancer Progression. Front. Oncol. 2021, 11, 778492. [Google Scholar] [CrossRef] [PubMed]
  174. Vaidya, S.P.; Gadre, S.; Kamisetti, R.T.; Patra, M. Challenges and opportunities in the development of metal-based anticancer theranostic agents. Biosci. Rep. 2022, 42, BSR20212160. [Google Scholar] [CrossRef]
  175. Kroschinsky, F.; Stölzel, F.; von Bonin, S.; Beutel, G.; Kochanek, M.; Kiehl, M.; Schellongowski, P. New drugs, new toxicities: Severe side effects of modern targeted and immunotherapy of cancer and their management. Crit. Care 2017, 21, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Gou, Y.; Huang, G.J.; Li, J.; Yang, F.; Liang, H. Versatile delivery systems for non-platinum metal-based anticancer therapeutic agents. Coord. Chem. Rev. 2021, 441, 213975. [Google Scholar] [CrossRef]
  177. Balali-Mood, M.; Naseri, K.; Tahergorabi, Z.; Khazdair, M.R.; Sadeghi, M. Toxic Mechanisms of Five Heavy Metals: Mercury, Lead, Chromium, Cadmium, and Arsenic. Front. Pharmacol. 2021, 12, 643972. [Google Scholar] [CrossRef]
  178. Wang, X.; Zhang, H.; Chen, X. Drug resistance and combating drug resistance in cancer. Cancer Drug Resist. 2019, 2, 141–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Boros, E.; Dyson, P.J.; Gasser, G. Classification of Metal-Based Drugs according to Their Mechanisms of Action. Chem 2020, 6, 41–60. [Google Scholar] [CrossRef]
  180. Butler, M.S.; Gigante, V.; Sati, H.; Paulin, S.; Al-Sulaiman, L.; Rex, J.H.; Fernandes, P.; Arias, C.A.; Paul, M.; Thwaites, G.E.; et al. Analysis of the Clinical Pipeline of Treatments for Drug-Resistant Bacterial Infections: Despite Progress, More Action Is Needed. Antimicrob. Agents Chemother. 2022, 66, 0199121. [Google Scholar] [CrossRef]
  181. Reig, S.; Le Gouellec, A.; Bleves, S. What Is New in the Anti–Pseudomonas aeruginosa Clinical Development Pipeline Since the 2017 WHO Alert? Front. Cell. Infect. Microbiol. 2022, 12, 909731. [Google Scholar] [CrossRef]
  182. Wang, C.; Yang, D.; Wang, Y.; Ni, W. Cefiderocol for the Treatment of Multidrug-Resistant Gram-Negative Bacteria: A Systematic Review of Currently Available Evidence. Front. Pharmacol. 2022, 13, 896971. [Google Scholar] [CrossRef] [PubMed]
  183. Zhanel, G.G.; Golden, A.R.; Zelenitsky, S.; Wiebe, K.; Lawrence, C.K.; Adam, H.J.; Idowu, T.; Domalaon, R.; Schweizer, F.; Zhanel, M.A.; et al. Cefiderocol: A Siderophore Cephalosporin with Activity Against Carbapenem-Resistant and Multidrug-Resistant Gram-Negative Bacilli. Drugs 2019, 79, 271–289. [Google Scholar] [CrossRef] [PubMed]
  184. Theuretzbacher, U.; Piddock, L.J.V. Non-traditional Antibacterial Therapeutic Options and Challenges. Cell Host Microbe 2019, 26, 61–72. [Google Scholar] [CrossRef] [PubMed]
  185. Theuretzbacher, U. Dual-mechanism antibiotics. Nat. Microbiol. 2020, 5, 984–985. [Google Scholar] [CrossRef]
  186. Rex, J.H.; Fernandez Lynch, H.; Cohen, I.G.; Darrow, J.J.; Outterson, K. Designing development programs for non-traditional antibacterial agents. Nat. Commun. 2019, 10, 3416. [Google Scholar] [CrossRef] [Green Version]
  187. Langendonk, R.F.; Neill, D.R.; Fothergill, J.L. The Building Blocks of Antimicrobial Resistance in Pseudomonas aeruginosa: Implications for Current Resistance-Breaking Therapies. Front. Cell. Infect. Microbiol. 2021, 11, 665759. [Google Scholar] [CrossRef]
  188. Nasiri Sovari, S.; Zobi, F. Recent Studies on the Antimicrobial Activity of Transition Metal Complexes of Groups 6–12. Chemistry 2020, 2, 418–452. [Google Scholar] [CrossRef]
  189. Frei, A.; Zuegg, J.; Elliott, A.G.; Baker, M.; Braese, S.; Brown, C.; Chen, F.; Dowson, C.G.; Dujardin, G.; Jung, N.; et al. Metal complexes as a promising source for new antibiotics. Chem. Sci. 2020, 11, 2627–2639. [Google Scholar] [CrossRef] [Green Version]
  190. Viganor, L.; Howe, O.; McCarron, P.; McCann, M.; Devereux, M. The Antibacterial Activity of Metal Complexes Containing 1,10- phenanthroline: Potential as Alternative Therapeutics in the Era of Antibiotic Resistance. Curr. Top. Med. Chem. 2016, 17, 1280–1302. [Google Scholar] [CrossRef]
  191. Tempera, P.J.; Michael, M.; Tageldin, O.; Hasak, S. Gastric Cancer Due to Chronic H. pylori Infection: What We Know and Where We Are Going. Diseases 2022, 10, 57. [Google Scholar] [CrossRef]
  192. Peek, R.M.; Blaser, M.J. Helicobacter pylori and gastrointestinal tract adenocarcinomas. Nat. Rev. Cancer 2002, 2, 28–37. [Google Scholar] [CrossRef]
  193. Suerbaum, S.; Michetti, P. INCE the first culture of. N. Engl. J. Med. 2002, 347, 1175–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Robin Warren, J.; Marshall, B. Unidentified Curved Bacilli on Gastric Epithelium in Active Chronic Gastritis. Lancet 1983, 321, 1273–1275. [Google Scholar] [CrossRef]
  195. Hooi, J.K.Y.; Lai, W.Y.; Ng, W.K.; Suen, M.M.Y.; Underwood, F.E.; Tanyingoh, D.; Malfertheiner, P.; Graham, D.Y.; Wong, V.W.S.; Wu, J.C.Y.; et al. Global Prevalence of Helicobacter pylori Infection: Systematic Review and Meta-Analysis. Gastroenterology 2017, 153, 420–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Liou, J.M.; Malfertheiner, P.; Lee, Y.C.; Sheu, B.S.; Sugano, K.; Cheng, H.C.; Yeoh, K.G.; Hsu, P.I.; Goh, K.L.; Mahachai, V.; et al. Screening and eradication of Helicobacter pylori for gastric cancer prevention: The Taipei global consensus. Gut 2020, 69, 2093–2112. [Google Scholar] [CrossRef] [PubMed]
  197. Piscione, M.; Mazzone, M.; Di Marcantonio, M.C.; Muraro, R.; Mincione, G. Eradication of Helicobacter pylori and Gastric Cancer: A Controversial Relationship. Front. Microbiol. 2021, 12, 630852. [Google Scholar] [CrossRef]
  198. Boyanova, L.; Hadzhiyski, P.; Gergova, R.; Markovska, R. Evolution of Helicobacter pylori Resistance to Antibiotics: A Topic of Increasing Concern. Antibiotics 2023, 12, 332. [Google Scholar] [CrossRef]
  199. Wang, H.; Yan, A.; Liu, Z.; Yang, X.; Xu, Z.; Wang, Y.; Wang, R.; Koohi-Moghadam, M.; Hu, L.; Xia, W.; et al. Deciphering molecular mechanism of silver by integrated omic approaches enables enhancing its antimicrobial efficacy in E. coli. PLoS Biol. 2019, 17, 3000292. [Google Scholar] [CrossRef] [Green Version]
  200. Frei, A.; Verderosa, A.D.; Elliott, A.G.; Zuegg, J.; Blaskovich, M.A.T. Metals to combat antimicrobial resistance. Nat. Rev. Chem. 2023, 7, 202–224. [Google Scholar] [CrossRef]
  201. Maier, R.J.; Benoit, S.L. Role of nickel in microbial pathogenesis. Inorganics 2019, 7, 80. [Google Scholar] [CrossRef] [Green Version]
  202. Keogan, D.M.; Griffith, D.M. Current and potential applications of bismuth-based drugs. Molecules 2014, 19, 15258–15297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Shetu, S.A.; Sanchez-Palestino, L.M.; Rivera, G.; Bandyopadhyay, D. Medicinal bismuth: Bismuth-organic frameworks as pharmaceutically privileged compounds. Tetrahedron 2022, 129, 133117. [Google Scholar] [CrossRef]
  204. Abdelkhalek, A.; Abutaleb, N.S.; Mohammad, H.; Mohamed, N.; Lafayette, W.; Disease, I.; Lafayette, W. Antibacterial and antivirulence activities of auranofin against Clostridium difficile. Int. J. Antimicrob. Agents 2020, 53, 54–62. [Google Scholar] [CrossRef] [PubMed]
  205. Thangamani, S.; Mohammad, H.; Abushahba, M.F.N.; Sobreira, T.J.P.; Hedrick, V.E.; Paul, L.N.; Seleem, M.N. Antibacterial activity and mechanism of action of auranofin against multi-drug resistant bacterial pathogens. Sci. Rep. 2016, 6, 22571. [Google Scholar] [CrossRef] [Green Version]
  206. AbdelKhalek, A.; Abutaleb, N.S.; Elmagarmid, K.A.; Seleem, M.N. Repurposing auranofin as an intestinal decolonizing agent for vancomycin-resistant enterococci. Sci. Rep. 2018, 8, 8353. [Google Scholar] [CrossRef] [Green Version]
  207. Abutaleb, N.S.; Seleem, M.N. Antivirulence activity of auranofin against vancomycin-resistant enterococci: In vitro and in vivo studies. Int. J. Antimicrob. Agents 2020, 55, 105828. [Google Scholar] [CrossRef]
  208. Kim, N.-H.; Lee, M.-Y.; Park, S.-J.; Choi, J.-S.; Oh, M.-K.; Kim, I.-S. Auranofin blocks interleukin-6 signalling by inhibiting phosphorylation of JAK1 and STAT3 Nam-Hoon. Immunology 2007, 122, 607–614. [Google Scholar] [CrossRef]
  209. Han, Y.; Chen, P.; Zhang, Y.; Lu, W.; Ding, W.; Luo, Y.; Wen, S.; Xu, R.; Liu, P.; Huang, P. Synergy between auranofin and celecoxib against colon cancer in vitro and in vivo through a novel redox-mediated mechanism. Cancers 2019, 11, 931. [Google Scholar] [CrossRef] [Green Version]
  210. Fiskus, W.; Saba, N.; Shen, M.; Ghias, M.; Liu, J.; Gupta, S.D.; Chauhan, L.; Rao, R.; Gunewardena, S.; Schorno, K.; et al. Auranofin induces lethal oxidative and endoplasmic reticulum stress and exerts potent preclinical activity against chronic lymphocytic leukemia. Cancer Res. 2014, 74, 2520–2532. [Google Scholar] [CrossRef]
  211. Sharlow, E.R.; Leimgruber, S.; Murray, S.; Lira, A.; Sciotti, R.J.; Hickman, M.; Hudson, T.; Leed, S.; Caridha, D.; Barrios, A.M.; et al. Auranofin is an apoptosis-simulating agent with in vitro and in vivo anti-leishmanial activity. ACS Chem. Biol. 2014, 9, 663–672. [Google Scholar] [CrossRef]
  212. Yang, L.; Wang, H.; Yang, X.; Wu, Q.; An, P.; Jin, X.; Liu, W.; Huang, X.; Li, Y.; Yan, S.; et al. Auranofin mitigates systemic iron overload and induces ferroptosis via distinct mechanisms. Signal Transduct. Target. Ther. 2020, 5, 138. [Google Scholar] [CrossRef] [PubMed]
  213. Sen, S.; Perrin, M.W.; Sedgwick, A.C.; Lynch, V.M.; Sessler, J.L.; Arambula, J.F. Covalent and non-covalent albumin binding of Au(i) bis-NHCsviapost-synthetic amide modification. Chem. Sci. 2021, 12, 7547–7553. [Google Scholar] [CrossRef] [PubMed]
  214. Martín-Encinas, E.; Conejo-Rodríguez, V.; Miguel, J.A.; Martínez-Ilarduya, J.M.; Rubiales, G.; Knudsen, B.R.; Palacios, F.; Alonso, C. Novel phosphine sulphide gold(i) complexes: Topoisomerase I inhibitors and antiproliferative agents. Dalton Trans. 2020, 49, 7852–7861. [Google Scholar] [CrossRef]
  215. Hu, J.; Zhang, H.; Cao, M.; Wang, L.; Wu, S.; Fang, B. Auranofin enhances ibrutinib’s anticancer activity in EGFR-mutant lung adenocarcinoma. Mol. Cancer Ther. 2018, 17, 2156–2163. [Google Scholar] [CrossRef] [Green Version]
  216. Park, S.-J.; Kim, I.-S. The role of p38 MAPK activation in auranofin-induced apoptosis of human promyelocytic leukaemia HL-60 cells. Br. J. Pharmacol. 2005, 146, 506–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Liu, N.; Li, X.; Huang, H.; Zhao, C.; Liao, S.; Yang, C.; Liu, S.; Song, W.; Lu, X.; Lan, X.; et al. Clinically used antirheumatic agent auranofin is a proteasomal deubiquitinase inhibitor and inhibits tumor growth. Oncotarget 2014, 5, 5453–5471. [Google Scholar] [CrossRef] [Green Version]
  218. Husain, S.; Nandi, A.; Simnani, F.Z.; Saha, U.; Ghosh, A.; Sinha, A.; Sahay, A.; Samal, S.K.; Panda, P.K.; Verma, S.K. Emerging Trends in Advanced Translational Applications of Silver Nanoparticles: A Progressing Dawn of Nanotechnology. J. Funct. Biomater. 2023, 14, 47. [Google Scholar] [CrossRef]
  219. O’Shaughnessy, M.; Hurley, J.; Dillon, S.C.; Herra, C.; McCarron, P.; McCann, M.; Devereux, M.; Howe, O. Antibacterial activity of metal–phenanthroline complexes against multidrug-resistant Irish clinical isolates: A whole genome sequencing approach. J. Biol. Inorg. Chem. 2023, 28, 153–171. [Google Scholar] [CrossRef]
  220. Wang, H.; Wang, M.; Xu, X.; Gao, P.; Xu, Z.; Zhang, Q.; Li, H.; Yan, A.; Kao, R.Y.T.; Sun, H. Multi-target mode of action of silver against Staphylococcus aureus endows it with capability to combat antibiotic resistance. Nat. Commun. 2021, 12, 3331. [Google Scholar] [CrossRef]
  221. Jung, W.K.; Koo, H.C.; Kim, K.W.; Shin, S.; Kim, S.H.; Park, Y.H. Antibacterial Activity and Mechanism of Action of the Silver Ion in Staphylococcus aureus and Escherichia coli. Appl. Environ. Microbiol. 2008, 74, 2171–2178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. McQuillan, J.S.; Groenaga Infante, H.; Stokes, E.; Shaw, A.M. Silver nanoparticle enhanced silver ion stress response in Escherichia coli K12. Nanotoxicology 2012, 6, 857–866. [Google Scholar] [CrossRef]
  223. Arakawa, H.; Neault, J.F.; Tajmir-Riahi, H.A. Silver(I) complexes with DNA and RNA studied by fourier transform infrared spectroscopy and capillary electrophoresis. Biophys. J. 2001, 81, 1580–1587. [Google Scholar] [CrossRef] [Green Version]
  224. Gordon, O.; Slenters, T.V.; Brunetto, P.S.; Villaruz, A.E.; Sturdevant, D.E.; Otto, M.; Landmann, R.; Fromm, K.M. Silver coordination polymers for prevention of implant infection: Thiol interaction, impact on respiratory chain enzymes, and hydroxyl radical induction. Antimicrob. Agents Chemother. 2010, 54, 4208–4218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Gu, M.; Imlay, J. The SoxRS response of Escherichia coli is directly activated by redox-cycling drugs rather than by superoxide. Mol. Microbiol. 2011, 4, 1136–1150. [Google Scholar] [CrossRef] [Green Version]
  226. Barras, F.; Aussel, L.; Ezraty, B. Silver and antibiotic, new facts to an old story. Antibiotics 2018, 7, 79. [Google Scholar] [CrossRef] [Green Version]
  227. Saulou-Bérion, C.; Gonzalez, I.; Enjalbert, B.; Audinot, J.N.; Fourquaux, I.; Jamme, F.; Cocaign-Bousquet, M.; Mercier-Bonin, M.; Girbal, L. Escherichia coli under ionic silver stress: An integrative approach to explore transcriptional, physiological and biochemical responses. PLoS ONE 2015, 10, 0145748. [Google Scholar] [CrossRef] [Green Version]
  228. Surwade, P.; Ghildyal, C.; Weikel, C.; Luxton, T.; Peloquin, D.; Fan, X.; Shah, V. Augmented antibacterial activity of ampicillin with silver nanoparticles against methicillin-resistant Staphylococcus aureus (MRSA). J. Antibiot. 2019, 72, 50–53. [Google Scholar] [CrossRef]
  229. Morones-Ramirez, J.; Winkler, J.A.; Spina, C.S.; Collins, J.J.; Morones-Ramirez, J.R.; Winkler, J.A.; Spina, C.S.; Collins, J.J.; RubenMorones-Ramirez, J.; Winkler, J.A.; et al. Silver Enhances Antibiotic Activity Against Gram-negative Bacteria. Sci. Transl. Med. 2013, 5, 190ra81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Panáček, A.; Smékalová, M.; Večeřová, R.; Bogdanová, K.; Röderová, M.; Kolář, M.; Kilianová, M.; Hradilová, Š.; Froning, J.P.; Havrdová, M.; et al. Silver nanoparticles strongly enhance and restore bactericidal activity of inactive antibiotics against multiresistant Enterobacteriaceae. Colloids Surf. B Biointerfaces 2016, 142, 392–399. [Google Scholar] [CrossRef]
  231. Habash, M.B.; Goodyear, M.C.; Park, A.J.; Surette, M.D.; Vis, E.C.; Harris, R.J.; Khursigara, C.M.; Khursigara, M. Potentiation of Tobramycin by Silver Nanoparticles against Pseudomonas aeruginosa Biofilm. Antimicrob. Agents Chemother. 2017, 61, e00415-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. O’Shaughnessy, M.; McCarron, P.; Viganor, L.; McCann, M.; Devereux, M.; Howe, O. The antibacterial and anti-biofilm activity of metal complexes incorporating 3,6,9-trioxaundecanedioate and 1,10-phenanthroline ligands in clinical isolates of Pseudomonas aeruginosa from irish cystic fibrosis patients. Antibiotics 2020, 9, 674. [Google Scholar] [CrossRef]
  233. Williams, K.; Milner, J.; Boudreau, M.D.; Gokulan, K.; Cerniglia, C.E.; Khare, S. Effects of subchronic exposure of silver nanoparticles on intestinal microbiota and gut-associated immune responses in the ileum of Sprague-Dawley rats. Nanotoxicology 2015, 9, 279–289. [Google Scholar] [CrossRef]
  234. Hadrup, N.; Gao, X.; Lam, H.R.; Loeschner, K.; Vogel, U.; Bergström, A.; Frandsen, H.L.; Mortensen, A.; Wilcks, A.; Larsen, E.H. Subacute oral toxicity investigation of nanoparticulate and ionic silver in rats. Arch. Toxicol. 2012, 86, 543–551. [Google Scholar] [CrossRef]
  235. Javurek, A.B.; Suresh, D.; Spollen, W.G.; Hart, M.L.; Hansen, A.; Ellersieck, M.R.; Bivens, N.J.; Givan, S.A. Gut Dysbiosis and Neurobehavioral Alterations in Rats Exposed to Silver Nanoparticles. Sci. Rep. 2017, 7, 2822. [Google Scholar] [CrossRef] [PubMed]
  236. Lyu, Z.; Ghoshdastidar, S.; Rekha, K.R.; Suresh, D. Developmental exposure to silver nanoparticles leads to long term gut dysbiosis and neurobehavioral alterations. Sci. Rep. 2021, 11, 6558. [Google Scholar] [CrossRef] [PubMed]
  237. Kankala, S.; Thota, N.; Björkling, F.; Taylor, M.K.; Vadde, R.; Balusu, R. Silver carbene complexes: An emerging class of anticancer agents. Drug Dev. Res. 2019, 80, 188–199. [Google Scholar] [CrossRef] [PubMed]
  238. Akkoç, M.; Khan, S.; Yüce, H.; Türkmen, N.B.; Yaşar, Ş.; Yaşar, S.; Özdemir, İ. Molecular docking and in vitro anticancer studies of silver(I)-N-heterocyclic carbene complexes. Heliyon 2022, 8, 10133. [Google Scholar] [CrossRef]
  239. Sammes, P.G.; Yahioglu, G. 1,10-Phenanthroline: A versatile ligand. Chem. Soc. Rev. 1994, 23, 327–334. [Google Scholar] [CrossRef]
  240. Bencini, A.; Lippolis, V. 1,10-Phenanthroline: A versatile building block for the construction of ligands for various purposes. Coord. Chem. Rev. 2010, 254, 2096–2180. [Google Scholar] [CrossRef]
  241. Turian, G. Tuberculostatic action of o-phenanthroline. Schweiz. Z. Pathol. Bakteriol. 1951, 14, 338–344. [Google Scholar]
  242. Kilah, N.L.; Meggers, E. Sixty years young: The diverse biological activities of metal polypyridyl complexes pioneered by Francis P. Dwyer. Aust. J. Chem. 2012, 65, 1325–1332. [Google Scholar] [CrossRef]
  243. McCann, M.; Santos, A.L.S.; Da Silva, B.A.; Romanos, M.T.V.; Pyrrho, A.S.; Devereux, M.; Kavanagh, K.; Fichtner, I.; Kellett, A. In vitro and in vivo studies into the biological activities of 1,10-phenanthroline, 1,10-phenanthroline-5,6-dione and its copper(ii) and silver(i) complexes. Toxicol. Res. 2012, 1, 47–54. [Google Scholar] [CrossRef] [Green Version]
  244. Santos, A.L.; Sodre, C.L.; Valle, R.S.; Silva, B.A.; Abi-Chacra, E.A.; Silva, L.V.; Souza-Goncalves, A.L.; Sangenito, L.S.; Goncalves, D.S.; Souza, L.O.; et al. Antimicrobial Action of Chelating Agents: Repercussions on the Microorganism Development, Virulence and Pathogenesis. Curr. Med. Chem. 2012, 19, 2715–2737. [Google Scholar] [CrossRef] [PubMed]
  245. Dos Santos, M.H.B.; Da Costa, A.F.E.; Santos, G.D.S.; Dos Santos, A.L.S.; Nagao, P.E. Effect of chelating agents on the growth, surface polypeptide synthesis and interaction of Streptococcus agalactiae with human epithelial cells. Mol. Med. Rep. 2009, 2, 81–84. [Google Scholar] [PubMed] [Green Version]
  246. Husseini, R.; Stretton, R.J. Studies on the antibacterial activity of phanquone: Chelating properties in relation to mode of action against Escherichia coli and Staphylococcus aureus. Microbios 1980, 29, 109–125. [Google Scholar]
  247. Zoroddu, M.A.; Zanetti, S.; Pogni, R.; Basosi, R. An electron spin resonance study and antimicrobial activity of copper(II)-phenanthroline complexes. J. Inorg. Biochem. 1996, 63, 291–300. [Google Scholar] [CrossRef]
  248. Kellett, A.; Howe, O.; O’Connor, M.; McCann, M.; Creaven, B.S.; McClean, S.; Foltyn-Arfa Kia, A.; Casey, A.; Devereux, M. Radical-induced DNA damage by cytotoxic square-planar copper(II) complexes incorporating o-phthalate and 1,10-phenanthroline or 2,2′-dipyridyl. Free Radic. Biol. Med. 2012, 53, 564–576. [Google Scholar] [CrossRef]
  249. Rochford, G.; Molphy, Z.; Browne, N.; Surlis, C.; Devereux, M.; McCann, M.; Kellett, A.; Howe, O.; Kavanagh, K. In-vivo evaluation of the response of Galleria mellonella larvae to novel copper(II) phenanthroline-phenazine complexes. J. Inorg. Biochem. 2018, 186, 135–146. [Google Scholar] [CrossRef] [Green Version]
  250. Rochford, G.; Molphy, Z.; Kavanagh, K.; McCann, M.; Devereux, M.; Kellett, A.; Howe, O. Cu(ii) phenanthroline-phenazine complexes dysregulate mitochondrial function and stimulate apoptosis. Metallomics 2020, 12, 65–78. [Google Scholar] [CrossRef]
  251. Thornton, L.; Dixit, V.; Assad, L.O.N.; Ribeiro, T.P.; Queiroz, D.D.; Kellett, A.; Casey, A.; Colleran, J.; Pereira, M.D.; Rochford, G.; et al. Water-soluble and photo-stable silver(I) dicarboxylate complexes containing 1,10-phenanthroline ligands: Antimicrobial and anticancer chemotherapeutic potential, DNA interactions and antioxidant activity. J. Inorg. Biochem. 2016, 159, 120–132. [Google Scholar] [CrossRef]
  252. Zhang, J.; Li, Y.; Fang, R.; Wei, W.; Wang, Y.; Jin, J.; Yang, F.; Chen, J. Organometallic gold(I) and gold(III) complexes for lung cancer treatment. Front. Pharmacol. 2022, 13, 979951. [Google Scholar] [CrossRef]
  253. Iglesias, S.; Alvarez, N.; Torre, M.H.; Kremer, E.; Ellena, J.; Ribeiro, R.R.; Barroso, R.P.; Costa-Filho, A.J.; Kramer, G.M.; Facchin, G. Synthesis, structural characterization and cytotoxic activity of ternary copper(II)-dipeptide-phenanthroline complexes. A step towards the development of new copper compounds for the treatment of cancer. J. Inorg. Biochem. 2014, 139, 117–123. [Google Scholar] [CrossRef] [PubMed]
  254. Pivetta, T.; Isaia, F.; Verani, G.; Cannas, C.; Serra, L.; Castellano, C.; Demartin, F.; Pilla, F.; Manca, M.; Pani, A. Mixed-1,10-phenanthroline-Cu(II) complexes: Synthesis, cytotoxic activity versus hematological and solid tumor cells and complex formation equilibria with glutathione. J. Inorg. Biochem. 2012, 114, 28–37. [Google Scholar] [CrossRef] [PubMed]
  255. Masuri, S.; Cadoni, E.; Cabiddu, M.G.; Isaia, F.; Demuru, M.G.; Morán, L.; Morán, L.; Bucek, D.; Vanhara, P.; Vanhara, P.; et al. The first copper(ii) complex with 1,10-phenanthroline and salubrinal with interesting biochemical properties. Metallomics 2020, 12, 891–901. [Google Scholar] [CrossRef] [PubMed]
  256. Olsen, P.M.; Ruiz, C.; Lussier, D.; Le, B.K.; Angel, N.; Smith, M.; Hwang, C.; Khatib, R.; Jenkins, J.; Adams, K.; et al. Synthesis, characterization, and antitumor activity of unusual pseudo five coordinate gold(III) complexes: Distinct cytotoxic mechanism or expensive ligand delivery systems? J. Inorg. Biochem. 2014, 141, 121–131. [Google Scholar] [CrossRef]
  257. Wenzel, M.N.; Mósca, A.F.; Graziani, V.; Aikman, B.; Thomas, S.R.; De Almeida, A.; Platts, J.A.; Re, N.; Coletti, C.; Marrone, A.; et al. Insights into the Mechanisms of Aquaporin-3 Inhibition by Gold(III) Complexes: The Importance of Non-Coordinative Adduct Formation. Inorg. Chem. 2019, 58, 2140–2148. [Google Scholar] [CrossRef] [PubMed]
  258. Yu, H.-j.; Liu, J.-p.; Hao, Z.-f.; He, J.; Sun, M.; Hu, S.; Yu, L.; Chao, H. Synthesis, characterization and biological evaluation of ruthenium(II) complexes [Ru(dtzp)(dppz)Cl]+ and [Ru(dtzp)(dppz)CH3CN]2+ for photodynamic therapy. Dye. Pigment. 2017, 136, 416–426. [Google Scholar] [CrossRef]
  259. Deo, K.M.; Pages, B.J.; Ang, D.L.; Gordon, C.P.; Aldrich-Wright, J.R. Transition Metal Intercalators as Anticancer Agents-Recent Advances. Int. J. Mol. Sci. 2016, 17, 1818. [Google Scholar] [CrossRef] [Green Version]
  260. Sidambaram, P.; Colleran, J. Evaluating the anticancer properties and real-time electrochemical extracellular bio-speciation of bis (1,10-phenanthroline) silver (I) acetate monohydrate in the presence of A549 lung cancer cells. Biosens. Bioelectron. 2021, 175, 112876. [Google Scholar] [CrossRef]
  261. Gandra, R.M.; Carron, P.M.; Fernandes, M.F.; Ramos, L.S.; Mello, T.P.; Aor, A.C.; Branquinha, M.H.; McCann, M.; Devereux, M.; Santos, A.L.S. Antifungal potential of copper(II), Manganese(II) and silver(I) 1,10-phenanthroline chelates against multidrug-resistant fungal species forming the Candida haemulonii Complex: Impact on the planktonic and biofilm lifestyles. Front. Microbiol. 2017, 8, 1257. [Google Scholar] [CrossRef]
  262. Gandra, R.M.; McCarron, P.; Viganor, L.; Fernandes, M.F.; Kavanagh, K.; McCann, M.; Branquinha, M.H.; Santos, A.L.S.; Howe, O.; Devereux, M. In vivo Activity of Copper(II), Manganese(II), and Silver(I) 1,10-Phenanthroline Chelates Against Candida haemulonii Using the Galleria mellonella Model. Front. Microbiol. 2020, 11, 470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Granato, M.Q.; Gonçalves, D.d.S.; Seabra, S.H.; McCann, M.; Devereux, M.; dos Santos, A.L.S.; Kneipp, L.F. 1,10-phenanthroline-5,6-dione-based compounds are effective in disturbing crucial physiological events of Phialophora verrucosa. Front. Microbiol. 2017, 8, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Granato, M.Q.; Mello, T.P.; Nascimento, R.S.; Pereira, M.D.; Rosa, T.L.S.A.; Pessolani, M.C.V.; McCann, M.; Devereux, M.; Branquinha, M.H.; Santos, A.L.S.; et al. Silver(I) and Copper(II) Complexes of 1,10-Phenanthroline-5,6-Dione Against Phialophora verrucosa: A Focus on the Interaction With Human Macrophages and Galleria mellonella Larvae. Front. Microbiol. 2021, 12, 641258. [Google Scholar] [CrossRef] [PubMed]
  265. Lima, A.K.C.; Elias, C.G.R.; Oliveira, S.S.C.; Santos-Mallet, J.R.; McCann, M.; Devereux, M.; Branquinha, M.H.; Dutra, P.M.L.; Santos, A.L.S. Anti-Leishmania braziliensis activity of 1,10-phenanthroline-5,6-dione and its Cu(II) and Ag(I) complexes. Parasitol. Res. 2021, 120, 3273–3285. [Google Scholar] [CrossRef] [PubMed]
  266. Vargas Rigo, G.; Petro-Silveira, B.; Devereux, M.; McCann, M.; Souza Dos Santos, A.L.; Tasca, T. Anti-Trichomonas vaginalis activity of 1,10-phenanthroline-5,6-dione-based metallodrugs and synergistic effect with metronidazole. Parasitology 2019, 146, 1179–1183. [Google Scholar] [CrossRef]
  267. Rigo, G.V.; Cardoso, F.G.; Pereira, M.M.; Devereux, M. Peptidases Are Potential Targets of and Potent New Drug against Trichomonas vaginalis. Pathogens 2023, 12, 745. [Google Scholar] [CrossRef]
  268. Silva-oliveira, R.; Sangenito, L.S.; Reddy, A.; Velasco-torrijos, T. Tropical Medicine and Infectious Disease In Vitro Effects of Aminopyridyl Ligands Complexed to Copper (II) on the Physiology and Interaction Process of In Vitro Effects of Aminopyridyl Ligands Complexed to Copper (II) on the Physiology and Interaction. Trop. Med. Infect. Dis. 2023, 8, 288. [Google Scholar] [CrossRef]
  269. Papadia, P.; Margiotta, N.; Bergamo, A.; Sava, G.; Natile, G. Platinum(II) complexes with antitumoral/antiviral aromatic heterocycles: Effect of glutathione upon in vitro cell growth inhibition. J. Med. Chem. 2005, 48, 3364–3371. [Google Scholar] [CrossRef]
  270. Shulman, A.; White, D.O. Virostatic activity of 1,10-phenanthroline transition metal chelates: A structure-activity analysis. Chem. Biol. Interact. 1973, 6, 407–413. [Google Scholar] [CrossRef]
  271. Mazumder, A.; Gupta, M.; Perrin, D.M.; Sigman, D.S.; Rabinovitz, M.; Pommier, Y. Inhibition of Human Immunodeficiency Virus Type 1 Integrase by a Hydrophobic Cation: The Phenanthroline-Cuprous Complex. AIDS Res. Hum. Retroviruses 1995, 11, 115–125. [Google Scholar] [CrossRef]
  272. Chang, E.L.; Simmers, C.; Knight, D.A. Cobalt complexes as antiviral and antibacterial agents. Pharmaceuticals 2010, 3, 1711–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Viganor, L.; Galdino, A.C.M.; Nunes, A.P.F.; Santos, K.R.N.; Branquinha, M.H.; Devereux, M.; Kellett, A.; McCann, M.; Santos, A.L.S. Anti-Pseudomonas aeruginosa activity of 1,10-phenanthroline-based drugs against both planktonic- and biofilm-growing cells. J. Antimicrob. Chemother. 2016, 71, 128–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. McCarron, P.; McCann, M.; Devereux, M.; Kavanagh, K.; Skerry, C.; Karakousis, P.C.; Aor, A.C.; Mello, T.P.; Santos, A.L.S.; Campos, D.L.; et al. Unprecedented in vitro antitubercular activitiy of manganese(II) complexes containing 1,10-phenanthroline and dicarboxylate ligands: Increased activity, superior selectivity, and lower toxicity in comparison to their copper(II) analogs. Front. Microbiol. 2018, 9, 1432. [Google Scholar] [CrossRef]
  275. Ahmed, M.; Rooney, D.; McCann, M.; Devereux, M.; Twamley, B.; Galdino, A.C.M.; Sangenito, L.S.; Souza, L.O.P.; Lourenço, M.C.; Gomes, K.; et al. Synthesis and antimicrobial activity of a phenanthroline-isoniazid hybrid ligand and its Ag+ and Mn2+ complexes. BioMetals 2019, 32, 671–682. [Google Scholar] [CrossRef]
  276. Ventura, R.F.; Galdino, A.C.M.; Viganor, L.; Schuenck, R.P.; Devereux, M.; McCann, M.; Santos, A.L.S.; Nunes, A.P.F. Antimicrobial action of 1,10-phenanthroline-based compounds on carbapenemase-producing Acinetobacter baumannii clinical strains: Efficacy against planktonic- and biofilm-growing cells. Braz. J. Microbiol. 2020, 51, 1703–1710. [Google Scholar] [CrossRef] [PubMed]
  277. Vianez Peregrino, I.; Ferreira Ventura, R.; Borghi, M.; Pinto Schuenck, R.; Devereux, M.; McCann, M.; Souza dos Santos, A.L.; FerreiraNunes, A.P. Antibacterial activity and carbapenem re-sensitizing ability of 1,10-phenanthroline-5,6-dione and its metal complexes against KPC-producing Klebsiella pneumoniae clinical strains. Lett. Appl. Microbiol. 2021, 73, 139–148. [Google Scholar] [CrossRef]
  278. Dwyer, F.P.; Reid, I.K.; Shulman, A.; Laycock, G.M.; Dixson, S. The biological actions of 1,10-phenanthroline and 2,2′-bipyridine hydrochlorides, quaternary salts and metal chelates and related compounds. Aust. J. Exp. Biol. Med. 1969, 47, 203–218. [Google Scholar] [CrossRef]
  279. McCann, M.; Kellett, A.; Kavanagh, K.; Devereux, M.; Santos, A.L.S. Deciphering the Antimicrobial Activity of Phenanthroline Chelators. Curr. Med. Chem. 2012, 19, 2703–2714. [Google Scholar] [CrossRef]
  280. Raman, N.; Dhaveethu Raja, J.; Sakthivel, A. Synthesis, spectral characterization of Schiff base transition metal complexes: DNA cleavage and antimicrobial activity studies. J. Chem. Sci. 2007, 119, 303–310. [Google Scholar] [CrossRef]
  281. Butler, H.M.; Hurse, A.; Thursky, E.; Shulman, A. Bactericidal action of selected phenanthroline chelates and related compounds. Aust. J. Exp. Biol. Med. Sci. 1969, 47, 541–552. [Google Scholar] [CrossRef]
  282. Wang, S.; König, G.; Roth, H.J.; Fouché, M.; Rodde, S.; Riniker, S. Effect of Flexibility, Lipophilicity, and the Location of Polar Residues on the Passive Membrane Permeability of a Series of Cyclic Decapeptides. J. Med. Chem. 2021, 64, 12761–12773. [Google Scholar] [CrossRef] [PubMed]
  283. Dwyer, F.P.; Gyarfas, E.C.; Rogers, W.P.; Koch, J.H. Biological activity of complex ions. Nature 1952, 170, 190–191. [Google Scholar] [CrossRef] [PubMed]
  284. Kumar, S.V.; Scottwell, S.O.; Waugh, E.; McAdam, C.J.; Hanton, L.R.; Brooks, H.J.L.; Crowley, J.D. Antimicrobial Properties of Tris(homoleptic) Ruthenium(II) 2-Pyridyl-1,2,3-triazole “click” Complexes against Pathogenic Bacteria, Including Methicillin-Resistant Staphylococcus aureus (MRSA). Inorg. Chem. 2016, 55, 9767–9777. [Google Scholar] [CrossRef]
  285. Yang, X.; Sun, B.; Zhang, L.; Li, N.; Han, J.; Zhang, J.; Sun, X.; He, Q. Chemical Interference with Iron Transport Systems to Suppress Bacterial Growth of Streptococcus pneumoniae. PLoS ONE 2014, 9, 105953. [Google Scholar] [CrossRef] [Green Version]
  286. Yang, X.; Zhang, L.; Liu, J.; Li, N.; Yu, G.; Cao, K.; Han, H.; Zeng, G.; Pan, Y.; Sun, X.; et al. Proteomic analysis on the antibacterial activity of a Ru(II) complex against Streptococcus pneumoniae. J. Proteom. 2015, 115, 107–116. [Google Scholar] [CrossRef] [PubMed]
  287. Weber, D.K.; Sani, M.A.; Downton, M.T.; Separovic, F.; Keene, F.R.; Collins, J.G. Membrane Insertion of a Dinuclear Polypyridylruthenium(II) Complex Revealed by Solid-State NMR and Molecular Dynamics Simulation: Implications for Selective Antibacterial Activity. J. Am. Chem. Soc. 2016, 138, 15267–15277. [Google Scholar] [CrossRef]
  288. Li, F.; Harry, E.J.; Bottomley, A.L.; Edstein, M.D.; Birrell, G.W.; Woodward, C.E.; Keene, F.R.; Collins, J.G. Dinuclear ruthenium(ii) antimicrobial agents that selectively target polysomes in vivo. Chem. Sci. 2014, 5, 685–693. [Google Scholar] [CrossRef] [Green Version]
  289. Gorle, A.K.; Feterl, M.; Warner, J.M.; Wallace, L.; Keene, F.R.; Collins, J.G. Tri- and tetra-nuclear polypyridyl ruthenium(ii) complexes as antimicrobial agents. Dalton Trans. 2014, 43, 16713–16725. [Google Scholar] [CrossRef] [Green Version]
  290. Li, F.; Collins, J.G.; Keene, F.R. Ruthenium complexes as antimicrobial agents. Chem. Soc. Rev. 2015, 44, 2529–2542. [Google Scholar] [CrossRef] [Green Version]
  291. Gorle, A.K.; Li, X.; Primrose, S.; Li, F.; Feterl, M.; Kinobe, R.T.; Heimann, K.; Warner, J.M.; Richard Keene, F.; Grant Collins, J. Oligonuclear polypyridylruthenium(II) complexes: Selectivity between bacteria and eukaryotic cells. J. Antimicrob. Chemother. 2016, 71, 1547–1555. [Google Scholar] [CrossRef]
  292. Sigman, D.S.; Graham, D.R.; D’Aurora, V.; Stern, A.M. Inhibition polymerase of. J. Biol. Chem. 1979, 254, 12269–12272. [Google Scholar] [CrossRef]
  293. Bolhuis, A.; Aldrich-Wright, J.R. DNA as a target for antimicrobials. Bioorg. Chem. 2014, 55, 51–59. [Google Scholar] [CrossRef] [Green Version]
  294. Psomas, G.; Kessissoglou, D.P. Quinolones and non-steroidal anti-inflammatory drugs interacting with copper(ii), nickel(ii), cobalt(ii) and zinc(ii): Structural features, biological evaluation and perspectives. Dalton Trans. 2013, 42, 6252–6276. [Google Scholar] [CrossRef]
  295. Sousa, I.; Claro, V.; Pereira, J.L.; Amaral, A.L.; Cunha-Silva, L.; De Castro, B.; Feio, M.J.; Pereira, E.; Gameiro, P. Synthesis, characterization and antibacterial studies of a copper(II) levofloxacin ternary complex. J. Inorg. Biochem. 2012, 110, 64–71. [Google Scholar] [CrossRef]
  296. Fernandes, P.; Sousa, I.; Cunha-Silva, L.; Ferreira, M.; De Castro, B.; Pereira, E.F.; Feio, M.J.; Gameiro, P. Synthesis, characterization and antibacterial studies of a copper(II) lomefloxacin ternary complex. J. Inorg. Biochem. 2014, 131, 21–29. [Google Scholar] [CrossRef] [PubMed]
  297. Gameiro, P.; Rodrigues, C.; Baptista, T.; Sousa, I.; de Castro, B. Solution studies on binary and ternary complexes of copper(II) with some fluoroquinolones and 1,10-phenanthroline: Antimicrobial activity of ternary metalloantibiotics. Int. J. Pharm. 2007, 334, 129–136. [Google Scholar] [CrossRef]
  298. Ude, Z.; Kavanagh, K.; Twamley, B.; Pour, M.; Gathergood, N.; Kellett, A.; Marmion, C.J. A new class of prophylactic metallo-antibiotic possessing potent anti-cancer and anti-microbial properties. Dalton Trans. 2019, 48, 8578–8593. [Google Scholar] [CrossRef] [PubMed]
  299. Ude, Z.; Flothkötter, N.; Sheehan, G.; Brennan, M.; Kavanagh, K.; Marmion, C.J. Multi-targeted metallo-ciprofloxacin derivatives rationally designed and developed to overcome antimicrobial resistance. Int. J. Antimicrob. Agents 2021, 58, 106449. [Google Scholar] [CrossRef] [PubMed]
  300. Smoleński, P.; Jaros, S.W.; Pettinari, C.; Lupidi, G.; Quassinti, L.; Bramucci, M.; Vitali, L.A.; Petrelli, D.; Kochel, A.; Kirillov, A.M. New water-soluble polypyridine silver(i) derivatives of 1,3,5-triaza-7-phosphaadamantane (PTA) with significant antimicrobial and antiproliferative activities. Dalton Trans. 2013, 42, 6572–6581. [Google Scholar] [CrossRef]
  301. Chetana, P.R.; Srinatha, B.S.; Somashekar, M.N.; Policegoudra, R.S. Synthesis, spectroscopic characterisation, thermal analysis, DNA interaction and antibacterial activity of copper(I) complexes with N, N′- disubstituted thiourea. J. Mol. Struct. 2016, 45, 352–365. [Google Scholar] [CrossRef] [Green Version]
  302. Raman, N.; Raja, S.J. DNA cleavage, structural elucidation and anti-microbial studies of three novel mixed ligand Schiff base complexes of copper(II). J. Serbian Chem. Soc. 2007, 72, 983–992. [Google Scholar] [CrossRef]
  303. Tabassum, S.; Asim, A.; Arjmand, F.; Afzal, M.; Bagchi, V. Synthesis and characterization of copper(II) and zinc(II)-based potential chemotherapeutic compounds: Their biological evaluation viz. DNA binding profile, cleavage and antimicrobial activity. Eur. J. Med. Chem. 2012, 58, 308–316. [Google Scholar] [CrossRef] [PubMed]
  304. Ng, N.S.; Leverett, P.; Hibbs, D.E.; Yang, Q.; Bulanadi, J.C.; Jie Wu, M.; Aldrich-Wright, J.R. The antimicrobial properties of some copper(ii) and platinum(ii) 1,10-phenanthroline complexes. Dalton Trans. 2013, 42, 3196–3209. [Google Scholar] [CrossRef] [PubMed]
  305. Shivakumar, L.; Shivaprasad, K.; Revanasiddappa, H.D. SODs, DNA binding and cleavage studies of new Mn(III) complexes with 2-((3-(benzyloxy)pyridin-2-ylimino)methyl)phenol. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2013, 107, 203–212. [Google Scholar] [CrossRef] [PubMed]
  306. Dimitrakopoulou, A.; Dendrinou-samara, C.; Pantazaki, A.A.; Raptopoulou, C.; Terzis, A.; Samaras, E.; Kessissoglou, D.P. Interaction of Fe (III) with herbicide-carboxylato ligands. Di-, tri- and tetra-nuclear compounds: Structure, antimicrobial study and DNA interaction. Inorganica Chim. Acta 2007, 360, 546–556. [Google Scholar] [CrossRef]
  307. Sharma, D.; Misba, L.; Khan, A.U. Antibiotics versus biofilm: An emerging battleground in microbial communities. Antimicrob. Resist. Infect. Control 2019, 8, 76. [Google Scholar] [CrossRef] [Green Version]
  308. Flemming, H.C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S.A.; Kjelleberg, S. Biofilms: An emergent form of bacterial life. Nat. Rev. Microbiol. 2016, 14, 563–575. [Google Scholar] [CrossRef]
  309. Flemming, H.C.; Wingender, J. The biofilm matrix. Nat. Rev. Microbiol. 2010, 8, 623–633. [Google Scholar] [CrossRef]
  310. Fenker, D.E.; Mcdaniel, C.T.; Panmanee, W.; Panos, R.J.; Sorscher, E.J.; Sabusap, C.; Clancy, J.P.; Hassett, D.J. A Comparison between Two Pathophysiologically Different yet Microbiologically Similar Lung Diseases: Cystic Fibrosis and Chronic Obstructive Pulmonary Disease. Int. J. Respir. Pulm. Med. 2018, 5, 098. [Google Scholar]
  311. Vestby, L.K.; Grønseth, T.; Simm, R.; Nesse, L.L. Bacterial biofilm and its role in the pathogenesis of disease. Antibiotics 2020, 9, 59. [Google Scholar] [CrossRef] [Green Version]
  312. Kolpen, M.; Kragh, K.N.; Enciso, J.B.; Faurholt-Jepsen, D.; Lindegaard, B.; Egelund, G.B.; Jensen, A.V.; Ravn, P.; Mathiesen, I.H.M.; Gheorge, A.G.; et al. Bacterial biofilms predominate in both acute and chronic human lung infections. Eur. Respir. J. 2022, 60, 4. [Google Scholar]
  313. Domingue, J.C.; Drewes, J.L.; Merlo, C.A.; Housseau, F.; Sears, C.L. Host Responses to Mucosal Biofilms in the Lung and Gut. Mucosal Immunol. 2020, 13, 413–422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Sharma, A.; Kumar, D.; Dahiya, K.; Hawthorne, S.; Jha, S.K.; Jha, N.K.; Nand, P.; Girgis, S.; Raj, S.; Srivastava, R.; et al. Advances in pulmonary drug delivery targeting microbial biofilms in respiratory diseases. Nanomedicine 2021, 16, 1905–1923. [Google Scholar] [CrossRef] [PubMed]
  315. Beeton, M.L.; Aldrich-Wright, J.R.; Bolhuis, A. The antimicrobial and antibiofilm activities of copper(II) complexes. J. Inorg. Biochem. 2014, 140, 167–172. [Google Scholar] [CrossRef]
  316. O’Shaughnessy, M.; Piatek, M.; McCarron, P.; McCann, M.; Devereux, M.; Kavanagh, K.; Howe, O. In Vivo Activity of Metal Complexes Containing 1,10-Phenanthroline and 3,6,9-Trioxaundecanedioate Ligands against Pseudomonas aeruginosa Infection in Galleria mellonella Larvae. Biomedicines 2022, 10, 222. [Google Scholar] [CrossRef]
  317. Zhang, H.; Niu, Y.; Tan, J.; Liu, W.; Sun, M.; Yang, E.; Wang, Q.; Li, R.; Wang, Y.; Liu, W. Global Screening of Genomic and Transcriptomic Factors Associated with Phenotype Differences between Multidrug-Resistant and -Susceptible Candida haemulonii Strains. mSystems 2019, 4, 00459-19. [Google Scholar] [CrossRef] [Green Version]
  318. Zuo, Y.H.; Wang, W.Q.; Chen, Q.J.; Liu, B.; Zhang, F.Y.; Jin, X.Y.; Hang, J.Q.; Li, H.Y.; Bao, Z.Y.; Jie, Z.J.; et al. Candida in Lower Respiratory Tract Increases the Frequency of Acute Exacerbation of Chronic Obstructive Pulmonary Disease: A Retrospective Case-Control Study. Front. Cell. Infect. Microbiol. 2020, 10, 538005. [Google Scholar] [CrossRef]
  319. Falkievich, D.B.; Martínez Medina, J.J.; Alegre, W.S.; López Tévez, L.L.; Franca, C.A.; Ferrer, E.G.; Williams, P.A.M. Computational studies, antimicrobial activity, inhibition of biofilm production and safety profile of the cadmium complex of 1,10-phenanthroline and cyanoguanidine. Appl. Organomet. Chem. 2022, 36, 6695. [Google Scholar] [CrossRef]
  320. Tay, C.X.; Quah, S.Y.; Lui, J.N.; Yu, V.S.H.; Tan, K.S. Matrix metalloproteinase inhibitor as an antimicrobial agent to eradicate Enterococcus faecalis biofilm. J. Endod. 2015, 41, 858–863. [Google Scholar] [CrossRef]
  321. McCann, M.; Coyle, B.; McKay, S.; McCormack, P.; Kavanagh, K.; Devereux, M.; McKee, V.; Kinsella, P.; O’Connor, R.; Clynes, M. Synthesis and X-ray crystal structure of [Ag(phendio)2]ClO4 (phendio = 1,10-phenanthroline-5,6-dione) and its effects on fungal and mammalian cells. BioMetals 2004, 17, 635–645. [Google Scholar] [CrossRef] [Green Version]
  322. Galdino, A.C.M.; Viganor, L.; De Castro, A.A.; Da Cunha, E.F.F.; Mello, T.P.; Mattos, L.M.; Pereira, M.D.; Hunt, M.C.; O’Shaughnessy, M.; Howe, O.; et al. Disarming Pseudomonas aeruginosa virulence by the inhibitory action of 1,10-phenanthroline-5,6-dione-based compounds: Elastase B (lasB) as a chemotherapeutic target. Front. Microbiol. 2019, 10, 1701. [Google Scholar] [CrossRef] [Green Version]
  323. Galdino, A.C.M.; Viganor, L.; Pereira, M.M.; Devereux, M.; McCann, M.; Branquinha, M.H.; Molphy, Z.; O’Carroll, S.; Bain, C.; Menounou, G.; et al. Copper(II) and silver(I)-1,10-phenanthroline-5,6-dione complexes interact with double-stranded DNA: Further evidence of their apparent multi-modal activity towards Pseudomonas aeruginosa. J. Biol. Inorg. Chem. 2022, 27, 201–213. [Google Scholar] [CrossRef] [PubMed]
  324. Sulaiman, I.; Schuster, S.; Segal, L. Perspectives in lung microbiome research. Curr Opin Microbiol. 2020, 56, 24–29. [Google Scholar] [CrossRef] [PubMed]
  325. Chen, J.; Li, T.; Ye, C.; Zhong, J.; Huang, J.D.; Ke, Y.; Sun, H. The Lung Microbiome: A New Frontier for Lung and Brain Disease. Int. J. Mol. Sci. 2023, 24, 2170. [Google Scholar] [CrossRef] [PubMed]
  326. Broderick, D.; March, R.; Waite, D.; Pillarisetti, N.; Chang, A.B.; Taylor, M.W. Realising respiratory microbiomic meta-analyses: Time for a standardised framework. Microbiome 2023, 11, 57. [Google Scholar] [CrossRef]
  327. Pandey, A.; Boros, E. Coordination Complexes to Combat Bacterial Infections: Recent Developments, Current Directions and Future Opportunities. Chem. Eur. J. 2021, 27, 7340–7350. [Google Scholar] [CrossRef]
  328. Butler, M.S.; Henderson, I.R.; Capon, R.J.; Blaskovich, M.A.T. Antibiotics in the clinical pipeline as of December 2022. J Antibiot 2023, 76, 431–473. [Google Scholar] [CrossRef]
  329. Huynh, M.; Crane, M.J.; Jamieson, A.M. The lung, the niche, and the microbe: Exploring the lung microbiome in cancer and immunity. Front. Immunol. 2023, 13, 1094110. [Google Scholar] [CrossRef]
  330. Martins, D.; Mendes, F.; Schmitt, F. Microbiome: A Supportive or a Leading Actor in Lung Cancer? Pathobiology 2021, 88, 198–207. [Google Scholar] [CrossRef]
Figure 1. Structure of (A) 1,10-phenanthroline (phen), and examples of its derivatives, (B) 3,4,7,8-tetramethyl-1,10-phenanthroline, (C) 5-nitro-1,10-phenanthroline, (D) 1,10-phenanthroline-5,6-dione (phendione) and (E) 2,9-dimethyl-1,10-phenathroline (neocuproine).
Figure 1. Structure of (A) 1,10-phenanthroline (phen), and examples of its derivatives, (B) 3,4,7,8-tetramethyl-1,10-phenanthroline, (C) 5-nitro-1,10-phenanthroline, (D) 1,10-phenanthroline-5,6-dione (phendione) and (E) 2,9-dimethyl-1,10-phenathroline (neocuproine).
Ijms 24 12296 g001
Figure 2. Metal-tdda-phen complexes: {[Cu(3,6,9-tdda)(phen)2]·3H2O·EtOH}n (Cu-tdda-phen), {[Mn(3,6,9-tdda)(phen)2]·3H2O·EtOH}n (Mn-tdda-phen) and [Ag2(3,6,9-tdda)(phen)4]·EtOH (Ag-tdda-phen). Adapted from [261].
Figure 2. Metal-tdda-phen complexes: {[Cu(3,6,9-tdda)(phen)2]·3H2O·EtOH}n (Cu-tdda-phen), {[Mn(3,6,9-tdda)(phen)2]·3H2O·EtOH}n (Mn-tdda-phen) and [Ag2(3,6,9-tdda)(phen)4]·EtOH (Ag-tdda-phen). Adapted from [261].
Ijms 24 12296 g002
Figure 3. Structure of phendione containing metal complexes [Cu(phendione)3]2+ and [Ag(phendione)2]+. Adapted from [321].
Figure 3. Structure of phendione containing metal complexes [Cu(phendione)3]2+ and [Ag(phendione)2]+. Adapted from [321].
Ijms 24 12296 g003
Table 1. Food and Drug Administration (FDA) and European Medicines Agency (EMA) approved metal-containing drugs, their therapeutic applications, and year of approval.
Table 1. Food and Drug Administration (FDA) and European Medicines Agency (EMA) approved metal-containing drugs, their therapeutic applications, and year of approval.
MetalDrugTherapeutic ApplicationYear Approved
Platinum (Pt)CisplatinCancer: lung, testicular, ovarian, bladder1978
CarboplatinCancer: lung, ovarian, head and neck1989
OxaliplatinCancer: colorectal1999
Gold (Au)AuranofinRheumatoid arthritis1985
Sodium aurothiomalateRheumatoid arthritis1985
Iron (Fe)DeferoxamineIron chelation (iron overload)1968
DeferiproneIron chelation (iron overload)1999
DeferasiroxIron chelation (iron overload)2005
Ferric carboxymaltoseIron deficiency anaemia2007
Iron isomaltoside 1000Iron deficiency anaemia2009
FerumoxytolIron deficiency anaemia in chronic kidney disease2009
Ferric citrateHyperphosphatemia in chronic kidney disease2014
Sucroferric oxyhydroxideHyperphosphatemia in chronic kidney disease2013
Bismuth (Bi)Bismuth subsalicylateGastrointestinal issues (ulcers, diarrhoea, acid reflux)1939
Bismuth subcitrate potassiumHelicobacter pylori infection, peptic ulcers1998
Lithium (Li)Lithium carbonateBipolar disorder1970
Gallium (Ga)Gallium nitrateHypercalcemia of malignancy2001
Lanthanum (La)Lanthanum carbonateRenal failure2004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

O’Shaughnessy, M.; Sheils, O.; Baird, A.-M. The Lung Microbiome in COPD and Lung Cancer: Exploring the Potential of Metal-Based Drugs. Int. J. Mol. Sci. 2023, 24, 12296. https://doi.org/10.3390/ijms241512296

AMA Style

O’Shaughnessy M, Sheils O, Baird A-M. The Lung Microbiome in COPD and Lung Cancer: Exploring the Potential of Metal-Based Drugs. International Journal of Molecular Sciences. 2023; 24(15):12296. https://doi.org/10.3390/ijms241512296

Chicago/Turabian Style

O’Shaughnessy, Megan, Orla Sheils, and Anne-Marie Baird. 2023. "The Lung Microbiome in COPD and Lung Cancer: Exploring the Potential of Metal-Based Drugs" International Journal of Molecular Sciences 24, no. 15: 12296. https://doi.org/10.3390/ijms241512296

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop