RESEARCH ARTICLE


Physiologic Activity of Bisphosphonates – Recent Advances



Ewa Chmielewska*, Paweł Kafarski
Department of Bioorganic Chemistry, Faculty of Chemistry, Wrocław University of Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland


Article Metrics

CrossRef Citations:
27
Total Statistics:

Full-Text HTML Views: 2057
Abstract HTML Views: 453
PDF Downloads: 3
ePub Downloads: 2
Total Views/Downloads: 2515
Unique Statistics:

Full-Text HTML Views: 995
Abstract HTML Views: 301
PDF Downloads: 3
ePub Downloads: 2
Total Views/Downloads: 1301



© Chmielewska and Kafarski; Bentham Open.

open-access license: This is an open access article licensed under the terms of the Creative Commons Attribution-Non-Commercial 4.0 International Public License (CC BY-NC 4.0) (https://creativecommons.org/licenses/by-nc/4.0/legalcode), which permits unrestricted, non-commercial use, distribution and reproduction in any medium, provided the work is properly cited.

* Address correspondence to this author at the Department of Bioorganic Chemistry, Faculty of Chemistry, Wrocław University of Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland; E-mail: ewa.chmielewska@pwr.edu.pl


Abstract

Background:

Bisphosphonates are drugs commonly used for the medication and prevention of diseases caused by decreased mineral density. Despite such important medicinal use, they display a variety of physiologic activities, which make them promising anti-cancer, anti-protozoal, antibacterial and antiviral agents.

Objective:

To review physiological activity of bisphosphonates with special emphasis on their ongoing and potential applications in medicine and agriculture.

Method:

Critical review of recent literature data.

Results:

Comprehensive review of activities revealed by bisphosphonates.

Conclusion:

although bisphosphonates are mostly recognized by their profound effects on bone physiology their medicinal potential has not been fully evaluated yet. Literature data considering enzyme inhibition suggest possibilities of far more wide application of these compounds. These applications are, however, limited by their low bioavailability and therefore intensive search for new chemical entities overcoming this shortage are carried out.

Keywords: Antibacterial agents, Anti-cancer agents, Anti-protozoal agents, Anti-resorptive agents, Antiviral agents, Osteopenia, Osteoporosis, Plant growth regulators.



INTRODUCTION

Bisphosphonates were first synthesized in the 19th century, and the history of their discovery was not without some drama [1]. Soon, these compounds had been found to act as strong metal ion complexones [2] with useful industrial and household applications, including detergents, water treatment agents, dispersants preventing re-disposition of insoluble inorganic matter and compounds avoiding the metal-catalyzed decomposition of hydrogen peroxide in bleaching formulations [3].

In the early 1960s, inorganic pyrophosphate was found to act as a natural inhibitor of calcification by its interaction with hydroxyapatite. This finding made it interesting for pharmacologic applications in the treatment of medical states related to bone resorption. Unfortunately, pyrophosphate is metabolically unstable because it is rapidly hydrolyzed in the gastrointestinal tract. Thus, seeking more stable compounds, attention turned to its analogs, bisphosphonates. Similarly to pyrophosphate, bisphosphonates exhibit high affinity for bone hydroxyapatite and effectively prevent calcification. Today, these compounds have become a powerful family of pharmaceuticals for the treatment of skeletal complications of malignancy, Paget’s disease, osteoporosis, multiple myeloma, hypercalcemia and fibrous dysplasia. Their applications, their clinical implications and their mechanism of action have been reviewed in detail [4-14].

Therefore, in this review, recent studies on the biological activity of bisphosphonates will be presented with some general background provided if necessary.

ANTI-RESORPTIVE BISPHOSPHONATES

It is estimated that osteoporosis affects 200 million women worldwide and causes a huge personal and economic problems. In Europe, disability caused by osteoporosis surpasses that caused by cancer (with the exception of lung cancer) and is proportional to or even exceeds that lost to a variety of chronic non-communicable diseases, such as rheumatoid arthritis, asthma and high blood pressure-related heart disease [15].

The treatment of osteoporosis consists of lifestyle measures and pharmacologic therapy. Approval by Food and Drug Administration (FDA) of alendronate in 1995 resulted in widespread use of bisphosphonates in clinical practice. Today, numerous members of this class of compounds are available as drugs. It is evident that over 50 years, the bisphosphonates revolutionized the treatment of osteoporosis being first-line therapy for the postmenopausal-type of this disease.

Early-generation bisphosphonates differ from later-generation bisphosphonates (representative examples are shown in Scheme (1) by the absence of a nitrogen atom in their structures. It was found further that nitrogen-containing bisphosphonates are able to inhibit bone resorption 100 to 10,000 times more effectively than non-nitrogen ones, causing nearly complete displacement of the latter compounds from therapy [16].

Scheme (1).

Clinically used bisphosphonates.


Drug efficacy is governed by its potency to inhibit farnesyl pyrophosphate synthase (FPPS), a primary target for the bisphosphonates) and its affinity to bone mineral. The latter influences uptake and retention of the drug by the skeleton, differential distribution with the bone and diffusion through the osteocyte lacunar-canalicular system [17].

Over the years, a huge variety of bisphosphonates have been obtained, and their influence on the process of bone resorption was evaluated. Because of the efficiency of currently used drugs, not much research is now being done on the design, synthesis and evaluation of new drug candidates. The large amount of data considers rather modes of their applications, longevity of treatment, recommendations and side-effects [18-20]. Most important is the evaluation of the risk of developing bisphosphonate-related osteonecrosis of the jaw – a fatal side effect of chronic use of medication [21].

Two types of efforts have been undertaken in the last decade to design new bisphosphonate anti-osteoporotic agents. The first effort is based on knowledge of the three-dimensional structure of human farnesyl pyrophosphate synthase (hFPPS) [22-24]. FPPS controls intracellular levels of farnesyl pyrophosphate, and thus, the process of protein farnesylation, which is critical for the proper subcellular localization and function of many proteins, including small GTPases that regulate a wide variety of cellular processes. In the case of bone growth modulation, inhibitors of this enzyme block excessive bone resorption in osteoclasts by causing apoptosis. The availability of the crystal structures of structurally variable bisphosphonates bound to the human enzyme allowed for the determination of its structural requirements [5, 25-27]. The knowledge of three-dimensional requirements of the enzyme active and binding sites, in turn resulted in the rationale (mostly computer-aided) design of novel effective inhibitors (representative structures are shown in Scheme 2) [5, 27-35]. However, none of these inhibitors were competitive with drugs already used to treat osteoporosis because most research effort was concentrated on finding possible anti-cancer rather than bone anti-resorptive agents.

Scheme (2).

Representative examples of bisphosphonates designed as human FPPS inhibitors.


Because some of the side-effects caused by bisphosphonates (especially jaw osteonecrosis) may be a result of their permanent binding to bone tissue, there has been interest in the development of more non-polar compounds and even non-bisphosphonate FPPS inhibitors [36, 37]. Such inhibitors should have also higher bioavailability and thus, are also considered better candidates for anti-cancer agents. The most successful implementation of this approach was the discovery of phosphonocarboxylate cognates of risendronate and minodronate (Scheme 3). These compounds inhibit bone resorption in vivo, however, to a lesser extent than the parent drugs [5, 32, 38-40]. Later, it was found that they act as inhibitors of Rab geranylgeranyl transferase and prevent geranylgeranylation of small Rab GTPases [5, 40, 41]. Thus, these compounds might be considered as a novel class of anti-cancer agents.

Since the hydroxyl group present in most commercialized drugs enhances their binding to bone, its removal from the molecule may result in weaker binding and shorter persistence in bones. Incadronate, also called cimadronate (Scheme 3), is a drug in development for treatment of osteoporosis and hypercalcemia and is a successful example of such reasoning [42]. This approach has also been applied to design strong inhibitors of FPPS from various sources,[5, 25, 43-46] with some of them showing unusual patterns of enzyme binding [47]. Interestingly, albeit not fully developed, the approach of replacing the hydroxyl group by fluorine atom has been tested [48]. Representative structures of these two classes of compounds are also shown in Scheme (3).

Scheme (3).

Bisphosphonates with reduced hydrophilicity.


Another approach was to build up libraries of amino-methylenebisphosphonates and screen their ability to inhibit proliferation of macrophage-like J774E cells [49-52]. This choice of a screening system seems to be reasonable because this cell line originates from the same precursors as osteoclasts. Although some of the studied compounds inhibited proliferation of the cells quite potently (representative structures are shown in Scheme 3), preclinical studies performed on sheep indicated that there is no direct relationship between the results of screening and the efficiency of medication in animals with induced osteoporosis [52].

An increase in hydrophobicity of bisphosphonates may also be achieved by modulation of the organic part of the molecule. Examples of a successful implementation of this idea include introduction of long hydrocarbon or aromatic substituents into molecules with previously reported activity (Scheme 4) [5, 25, 30, 31, 53, 54]. Interestingly, hydroxy-bisphosphonic analogs of bile acids appeared to be exceptionally active against L929 cells and cultures of osteoclasts. The lipophilic compounds act similarly as other bisphosphonates and additionally are more bioavailable [55]. Therefore, their influence on prenylation is not limited to bone cells and might be considered as drugs affecting multiple processes, or multi-target drugs.

Scheme (4).

Lipophilic bisphosphonates.


An obvious solution to reduce bone affinity and increase bisphosphonate bioavailability is its esterification, which would mask the negative charge of phosphonic groups [56-60]. The resulting esters should also act as pro-drugs being hydrolyzed within body fluids. Although some of the esters exerted anticancer activity in cell cultures, this effort did not bring promising results so far.

An important factor influencing the anti-resorptive action of bisphosphonates is their affinity toward bones. Although intensively studied using various techniques and theoretical approaches, [61-66] this process is not fully understood. Binding is mainly mediated by the presence of two phosphonic groups, but other structural features are also important. This is reflected by a high dependence of bone affinity on the chemical structure of these drugs.

The efficiency of bisphosphonates as anti-resorptive drugs is governed by their molecular mechanism of action, affinity to bones and distribution within this tissue. The latter was studied using either radioactive or fluorescent labeled drugs [2, 67-71]. These studies indicated that distribution is also dependent on the chemical features of the drugs.

In summary, better understanding of the molecular mode of action, an exceptional selectivity of bisphosphonates for bone mineral and the process of their distribution would explain their clinical features and creates new opportunities for further discoveries.

Another, albeit poorly developed, idea is the possible use of bisphosphonates for osseointegration of implants by coating their surface with these drugs and for healing of segmental bone fractures [72-76]. The latter system requires development of a local delivery system that is easy to handle by surgeons (most likely application via syringe) in a form that would remain in the targeted area of the affected tissue. Specific glues composed of hydroxyapatite and calcium phosphate microspheres or obtained by encapsulation of bisphosphonates in hydrogels have been designed for this application [77-83].

It should be noted, however, that bisphosphonates enhance bacterial adhesion and biofilm formation on the surface of bone hydroxyapatite. This may limit their use as osseointegrating agents [84, 85].

BISPHOSPHONATES AS BONE-TARGETING UNITS

The standard routes of administration for bisphosphonates used in clinical practice are either oral or intravenous infusion. Oral administration of bisphosphonates is complicated by poor bioavailability (generally below 1%) and poor gastrointestinal tolerability [86, 87]. Due to their avid affinity to bone, between 30% and 60% of the absorbed substance rapidly binds to bone mineral. This feature was used to construct drug-bisphosphonate conjugates, which might be considered as a promising method for selective drug targeting to the bone [88, 89]. Such an active transport of therapeutic agents to bone is called osteotropic drug delivery system (ODDS). It reduces drug toxicity and improves its bioavailability at the desired site. Bisphosphonates also have an advantage over other molecules because their affinity to bone may be tuned by variation of their chemical structures.

This system was also used directly for construction of bone-regenerative drugs. Parathyroid hormone (PTH) is an 84-amino acid polypeptide that plays an important role in calcium regulation and bone remodeling. Its 34-amino acid analog Teriparatide [90] retains most of the functions of PTH and is an FDA approved drug against osteoporosis. However, being a peptide, it is unstable in body fluids and is readily hydrolyzed by proteinases. Conjugation of this compound to hydrazine bisphosphonates of varying lengths and hydrophobicity was proposed as a way to improve the therapeutic properties of Teriparatide [91]. Another approach was to conjugate a popular anti-osteoporetic drug, pamidronate, with an edible polysaccharide, pullulan, and use this system for targeting fluorescent and magnetic resonance probes. In this manner, a theranostic system was obtained [92].

One of the major reasons of cancer-related women's death is the development of bone metastases. Therefore, the selective targeting of anti-cancer drugs to bone tissue should improve their pharmaceutic performance. For example, polymers conjugated with both bisphosphonates and anti-cancer paclitaxel have been designed [93-97] to synergistically combine the anti-mitotic effect of paclitaxel with the anti-angiogenic and bone-targeting properties of bisphosphonates [98]. Furthermore, direct conjunction of bisphosphonates with popularly used anti-cancer drugs such as camptothecin, [99] bortezomib, [100] doxorubicin [101] or gemcitabine [102] has been used to target these drugs to bone tissue or to multiple myeloma cells. Representative chemical structures of such conjugates are shown in (Scheme 5).

Scheme (5).

Examples of conjugates of bisphosphonates with anti-cancer drugs.


Another solution is to use conventional drug delivery systems, such as liposomes [103-106] or polymeric (preferably biodegradable) nanoparticles [107-110] with surfaces functionalized (incrusted) with bisphosphonates. Construction of such systems is challenging and thus, data on their functional activity are scarce.

Breast microcalcifications are found in about half of all women over the age of 50. They are a natural result of breast aging and are not usually due to cancer but can be a sign of pre-cancerous changes or early breast cancer if a group is found in one area and therefore, are diagnosed by mammography. Diagnosis, however, is limited to the sensitivity and specificity of this technique. Therefore, bisphosphonate-functionalized gold nanoparticles have been designed and used for contrast-enhanced radiographic studies of microcalcifications [111]. Similar approaches using specially designed technetium-99m and rhenium-188 complexes with pedant bisphosphonate groups have been applied for imaging of arterial calcification, [112] a symptom of cardiovascular disease.

Osteomyelitis is a serious bone infection most often caused by bacteria. The most common treatments are antibiotics and surgery to remove portions of bone that are infected or dead. Thus, selective targeting of antibiotics to bone seems to be profitable, and some preliminary attempts to design and obtain osteotropic systems with fluoroquinones, [113, 114] vancomycin [115] and rifamycin [116] as active components have been undertaken.

An interesting system for the treatment of kidney stones has been proposed recently. Kidney stones are endemic, and the use of extracorporeal shock wave lithotripsy where focused shock waves were used to fragment these stones have been studied [117]. The shockwaves induce the formation of cavitation bubbles, whose collapse releases energy at the stone, resulting in fragmentation into pieces small enough to be passed spontaneously. Because there is substantial amount of hydroxyapatite in these stones, bisphosphonates are good candidates to deliver microbubbles into or near urinary stones. Thus, a system has been designed where perfluoropropane is encapsulated in lipid microspheres and delivered to the kidney by a catheter. The lipidic surface is composed of the mixture of commercial lipid (DPCC) and lipid functionalized with bisphosphonate (Scheme 6). This mixture binds to the surface of the stone, and application of external ultrasonic irradiation causes stone fragmentation.

Scheme (6).

Bisphosphonate-based system for the delivery of gas microbubbles into kidney stones.


ANTI-CANCER ACTIVITY

As estimated, farnesylation or geranylgeranylation account for more than 2% of the human proteome. Farnesylation of the small GTP-binding proteins, Ras, is indispensable to regulate proliferation, invasive properties and pro-angiogenic activity in human cancers [118]. Nitrogen-containing bisphosphonates are inhibitors of human FPPS and related enzymes of the mevalonate pathway and also inhibit Ras farnesylation. Thus, they should be considered as possible anti-cancer agents. This reasoning is supported by the ability of these compounds to suppress proliferation of cancer cells of prostate, [119] breast, [120, 121] melanoma, [122] ovarian [123] and colorectal cancers, [124] as well as glioblastoma [125] and multiple myelanoma [126]. Additionally these drugs have also been shown to kill cancers in humans independently on their action on bones.

The most developed strategy is the use of bisphosphonates as adjuvants in breast cancer therapy. These compounds are the major weapon to prevent bone-loss and skeleton-related events (bone pain, pathologic fractures, spinal cord compression and hyperglycemia), which accompany breast cancer therapy with aromatase inhibitors [127-129].

Interestingly, epidemiological studies seem to reveal beneficial and preventive effects of anti-osteoporetic therapy on cancer development [130, 131]. Nonetheless, the mechanisms underscoring these anti-cancer actions are not well understood. This is because the physiologic mechanism of bisphosphonate action is complex, and they have been shown to block tumor growth independently of FPPS inhibition, namely through γ,δ T-cell receptor activation, [132, 133] NF-κB inhibition, [134]. VEGF and hypoxia inducible factor-α suppression [135] as well as inactivation of epidermal growth factor receptors [136]. There is also a report showing that bisphosphonates target the three most epigenetic cell levels, namely DNA methylation, histone deacetylation and microRNAs [137].

Recent study have shown that the use of carrier technology may convert antiresorptive zoledronate into manticancer therapeutic [210]. Reformulation of calcium zolendronate into nanoscale metal-organic framework, functionalized with folate (as a selective carrier to cancer cells) served for this purpose.

Nitrogen-containing bisphosphonates are limited as anti-cancer agents because of high bone affinity and low plasma concentration. Therefore, it will be necessary to use them in combination with other drugs or search for new chemical entities.

The strategy to increase compound plasma level relies on the synthesis of new agents with increased hydrophobicity. This approach has already been described.

Another approach is to search for completely new mechanisms of action. This is offered by the use of bisphosphonate complexes with polyoxometalates [138]. These complexes can act as dual inhibitors, as phosphonate inhibits FPPS, whereas polyoxometalate interferes with redox reactions. However, the mechanism of action of some of these complexes is independent of mevalonic pathway inhibition, which suggests action via an unknown mechanism [139-141].

There are limited reports on the use of bisphosphonates as agents that are synergistic with other drugs. A recent report on the use of lovastatin, zoledronate and digeranylbisphosphonate (Scheme 7) indicated that the simultaneous use of these three compounds as inhibitors of three subsequent steps in geranylgeranyl synthesis led to the inhibition of growth and induction of apoptosis in human chronic myelogenous leukemia cells [142, 143].

Scheme (7).

Three-component drug against myelogenous leukemia.


Being the most common source of cancer-related pain, bone metastases reduce functional capacity of patients and undermine their quality of life. Pain may be a result of disruption of tissue upon invasion and/or the pressure of tumorous tissue on nerve endings. Bone metastases can also activate pain receptors in pathologically changed bones. In the cases when pain is not manageable with analgesics it is commonly medicated by radiation therapy.

The efficiency of bisphosphonates in reducing pain from bone metastases is well-documented [144-146]. Zoledronic acid applied intravenously alone or complexed with radioactive metal ions has demonstrated the broadest clinical activity [147-149].

Complex Regional Pain syndrome type is a disease for which no gold-standard treatment exists to date. It may initially affect an arm or leg and spread throughout the body. Bisphosphonates offer some hope here, although the studies on their application are in infancy [150, 151].

ANTI-PROTOZOAL BISPHOSPHONATES

Protozoan infections are so called the “world's neglected diseases,” despite the fact that they affect millions of humans, spreading mostly in the poorest populations. The most important infections are leishmaniases, affecting 12 million inhabitants and Chagas disease with 8 million people affected worldwide [152]. Furthermore, malaria, which is a serious life-threatening infection caused by five different species of protozoa of the genus Plasmodium, is a huge public health concern [153]. According to the WHO report, there were 200 million malaria cases in 2013, with over half a million deaths.

Protozoan farnesyl pyrophosphate synthase (FPPS) is also a valuable target in the search for drugs against protozoa infections, and consequently, bisphosphonates were quite intensively studied here. The initial efforts had been concentrated on screening the influence of various series of these compounds against groups of parasites, such as Leishmania, [154, 155] Plasmodium, [46, 154] Trypanosoma, [154, 156, 157] Toxoplasma, [54, 154, 158] Schistosoma [159] and Cryptosporidium [160]. Thus, intensive screening of large libraries of structurally diverse bisphosphonates resulted in selection of the most effective structures and determination of the molecular mechanisms of their activities. Special emphasis have been put on bisphosphonate interactions with protozoan FPPSs and the related enzyme, namely geranylgeranyl phosphate synthase (GGPPS) [45, 161-167]. These studies have also been concentrated on the synthesis of more effective inhibitors, specifically those of increased lipophilicity appeared to be effective (Scheme 8).

Scheme (8).

Representative anti-protozoal bisphosphonates.


Bisphosphonates also appeared to inhibit other protozoan enzymes, namely hexokinase [168] and squalene synthetase [169]. Additionally, their interesting action on protozoan acidosomes, acidic organelles rich in calcium and phosphorus, have been observed [170]. These findings suggest that bisphosphonates might be considered as multi-target drugs.

The soil-dwelling amoeba Dictyostelium discoideum is known for its remarkable life cycle, which makes it an ideal model organism to study a range of biological problems, such as the mechanisms of action of both first generation [171, 172] and nitrogen-containing bisphosphonates [173]. These studies indicate that bisphosphonates are able to cross peroxisomal membranes before they can exert inhibitory action.

These studies also indicated that this class of compounds might be useful as anti-amoebic agents. This was supported by their inhibitory action towards Entamoeba histolytica, the causative agent of amebiasis and Naegleria fowleria, a vector of primary amebic meningoencephalitis [174, 175]. This was an important finding because the annual number of E. histolytica infections throughout the world is estimated to be approximately 50 million.

ANTIBACTERIAL BISPHOSPHONATES

Antibiotics have always been considered one of the greatest discoveries of the 20th century. This is true, but the real interest is in the rise of antibiotic resistance [176]. In the last two decades, multi-drug resistant microorganisms challenged the scientific community into developing new antimicrobial compounds. Isoprenoid biosynthesis is an important target here because isoprenoids are involved in the early steps of bacterial cell-wall biosynthesis. New compounds affecting cell wall biosynthesis are very promising because they could also cause restoration of sensitivity to existing drugs. Therefore, it is not surprising that a huge library of bisphosphonates were tested as inhibitors of Escherichia coli growth, with some of them exhibiting quite promising activities [177-180]. Additionally, they appeared to be synergistic with other phosphonate antibiotics, such as fosfomycin.

Some of the bisphosphonates designed as FPPS inhibitors appeared to also inhibit undecaprenyl diphosphate synthase, [181, 182] an enzyme that catalyzes the sequential condensations of a farnesyl pyrophosphate with eight isopentenyl pyro-phosphates, which results in the formation of new cis-double bonds and gives undecaprenyl pyrophosphate, a metabolite serving as a lipid carrier for peptidoglycan formation in the bacterial cell wall. Systematic molecular modeling based on crystallographic studies enabled the definition of structural requirements for its inhibitors [183].

Finding that two structurally similar bisphosphonates (Scheme 9) are able to inhibit different Mycobacterium tuberculosis enzymes, decaprenyl diphosphate synthase and tuberculosinol synthase, suggest that bisphosphonates could act as multi-targeting agents [184]. This possibility was confirmed by the ability of one of the inhibitors of undecaprenyl pyrophosphate synthase, shown in Scheme (9), to additionally intercalate with DNA [185].

Other enzymes have also been considered as targets for specific antibacterial agents, namely thymidyltransferase [186] and δ1-pyrroline-5-carboxylate reductase as an inhibitor of Streptococcus pyogenes , [187] which is the cause of many important human diseases ranging from mild superficial skin infections to life-threatening systemic infections. Additionally, glutamine synthetase could also be a target of a selective anti-tuberculosis agent for use against infections of the musculoskeletal system [188].

Preparation of hydroxyphosphonate derivatives of ciprofloxacin and moxifloxacin (representative structure is shown in Scheme 9) represents a similar concept of combining bone-targeting and antibacterial properties in one molecule [189].

Scheme (9).

Chosen antibacterials.


In summary, the application of bisphosphonates as potential antibacterial agents has been considered very recently, but the number of reports is limited. The design and synthesis of specific chemicals inhibiting target enzymes of chosen microorganisms is a solution.

ANTIVIRAL ACTIVITY

The finding that keto- and diketo-acids bind to the Mg(II)/Asp domain of HIV integrase in a manner similar to that observed upon their binding with prenyl transferases stimulated interest in possible use of bisphosphonates as its inhibitors [178]. This resulted in the synthesis and evaluation of several inhibitors of this enzyme (Scheme 10) [190-192]. However, integrase is a difficult target for the development of efficient anti-HIV drugs. This results from the fact that the compounds of interest have to be transported to the nuclei of infected cells, and polar bisphosphonates cannot be successfully transported.

There are also selected reports on the action of these compounds on HIV reverse transcriptase (Scheme 10)[191, 193].

Scheme (10).

Chosen antibacterials.


In considering the above information, it is not surprising that pamidronate was shown to inhibit influenza virus infections in mice [194].

PLANT GROWTH REGULATORS

The herbicidal effects of bisphosphonates were first discovered in 1979 according to the patent literature, but this did not attract much attention until 1995 [195]. Their herbicidal action was then intensively studied, and they were the first reported inhibitors of farnesyl pyrophosphate synthetase (FPPS) [196, 197]. This inhibitory activity was observed when studying bleaching herbicides influencing biosynthesis of caretonoids. Further studies have shown that herbicidal aminomethylene-bisphosphonates may be considered multi-target substances, which was documented by their inhibitory activities towards glutamine synthetase, [198, 199] 3-deoxy-D-arabinoheptulosonate-7-phosphate (DAHP) synthase, [200] δ1-pyrroline-5-carboxylate (P5CR) reductase [201, 202] and pyrophosphatase [203]. Thus, these compounds may be considered a heterogeneous group of compounds with variable modes of action (Scheme 11). Quite interestingly, most active compounds contain a halogen atom or atoms in the N-aromatic substituent (Scheme 11).

Unfortunately, despite their excellent action under laboratory conditions, they have not been introduced to agriculture. Thus, recent and scarce studies on the influence of these compounds on plant growth used bisphosphonates as tools to study the role of isoprenoid biosynthesis [204-207]. These compounds enabled us to determine the high elasticity of the chloroplastic isoprenoid pathway.

Scheme (11).

Plant growth regulating bisphosphonates.


Alpine pennycress (Noccaea caerulescens)is native to the mountains of central and southern Europe. It arrived in Finland at the end of the 19th century and is now used as a hyperaccumulator of zinc, cadmium and nickel ions for phytoremediation purposes [208]. The simultaneous use of the water insoluble, extremely effective metal chelator, hydroxyundecylidene-1,1,-bisphosphonic acid (Scheme 11), and this plant caused a significant increase of the metal sequestering properties of Alpine pennycress [209].

CONCLUSION

Bisphosphonates are mostly recognized by their profound effects on bone physiology. They inhibit bone resorption by inducing apoptosis of osteoclasts and thus, preventing age-related bone loss and deterioration of bone microarchitecture. Because some advanced cancers, such as breast or prostate cancer, can spread to the bone, bisphosphonates could modify the process of metastasis. Furthermore, bisphosphonates possess other useful physiologic properties, which make them promising anti-cancer, anti-protozoal, antibacterial and antiviral agents. However, their high polarity and thus low absorption in humans limit these applications. Therefore, the careful design of their chemical structures devoted to specific applications is required.

CONFLICT OF INTEREST

The authors confirm that this article content has no conflict of interest.

ACKNOWLEDGEMENTS

The work was financed by a statutory activity subsidy from the Polish Ministry of Science and Education for the Faculty of Chemistry of Wrocław University of Technology. Linguistic correction was financed by Wrocław Center of Biotechnology, program The Leading National Research Center (KNOW) for years 2014-2018.

REFERENCES

[1] Petroianu GA. Pharmacist Theodor Salzer (1833-1900) and the discovery of bisphosphonates. Pharmazie 2011; 66(10): 804-7.
[2] Gałęzowska J, Gumienna-Kontecka E. Phosphonates, their complexes and bio-applications: a spectrum of surprising diversity. Coord Chem Rev 2012; 256(1-2): 105-24.
[3] Studnik H, Liebsch S, Forlani G, Wieczorek D, Kafarski P, Lipok J. Amino polyphosphonates - chemical features and practical uses, environmental durability and biodegradation. N Biotechnol 2015; 32(1): 1-6.
[4] Russell RG. Bisphosphonates: the first 40 years. Bone 2011; 49(1): 2-19.
[5] Ebetino FH, Hogan AM, Sun S, et al. The relationship between the chemistry and biological activity of the bisphosphonates. Bone 2011; 49(1): 20-33.
[6] Cakarer S, Selvi F, Keskin C. Bisphosphonates and Bone. In: Al-Aubaidi Z, Ed. Orthopedic Surgery. Croatia: InTech Rijeka 2012.
[7] Widler L, Jahnke W, Green JR. The chemistry of bisphosphonates: from antiscaling agents to clinical therapeutics. Anticancer Agents Med Chem 2012; 12(2): 95-101.
[8] Lozano-Calderon SA, Colman MW, Raskin KA, Hornicek FJ, Gebhardt M. Use of bisphosphonates in orthopedic surgery: pearls and pitfalls. Orthop Clin North Am 2014; 45(3): 403-16.
[9] Pazianas M, van der Geest S, Miller P. Bisphosphonates and bone quality. Bonekey Rep 2014; 3(529): 529.
[10] Xue D, Li F, Chen G, Yan S. Pan. Do bisphosphonates affect bone healing? A meta-analysis of randomized controlled trials. J Orthop Surg 2014; 9(45): 1-7.
[11] Xu XL, Gou WL, Wang AY, et al. Basic research and clinical applications of bisphosphonates in bone disease: what have we learned over the last 40 years? J Transl Med 2013; 11(11)(303): 1-8.
[12] Heng C, Badner VM, Vakkas TG, Johnson R, Yeo Y. Bisphosphonate-related osteonecrosis of the jaw in patients with osteoporosis. Am Fam Physician 2012; 85(12): 1134-41.
[13] Anagha PP, Sen S. The efficacy of bisphosphonates in preventing aromatase inhibitor induced bone loss for postmenopausal women with early breast cancer: a systematic review and meta-analysis. J Oncol 2014; 2014: 625060.
[14] Moore SN, Tanner SB, Schoenecker JG. Bisphosphonates: from softening water to treating PXE. Cell Cycle 2015; 14(9): 1354-5.
[15] Johnell O, Kanis JA. An estimate of the worldwide prevalence and disability associated with osteoporotic fractures. Osteoporos Int 2006; 17(12): 1726-33.
[16] Kennel KA, Drake MT. Adverse effects of bisphosphonates: implications for osteoporosis management. Mayo Clin Proc 2009; 84(7): 632-7.
[17] Russell RG, Xia Z, Dunford JE, et al. Bisphosphonates: an update on mechanisms of action and how these relate to clinical efficacy. Ann NY Acad Sci 2007; 1117: 209-57.
[18] Black DM, Bauer DC, Schwartz AV, Cummings SR, Rosen CJ. Continuing bisphosphonate treatment for osteoporosis-for whom and for how long? N Engl J Med 2012; 366(22): 2051-3.
[19] McClung M, Harris ST, Miller PD, et al. Bisphosphonate therapy for osteoporosis: benefits, risks, and drug holiday. Am J Med 2013; 126(1): 13-20.
[20] Molvik H, Khan W. Bisphosphonates and their influence on fracture healing: a systematic review. Osteoporos Int 2015; 26(4): 1251-60.
[21] Rasmusson L, Abtahi J. Associated Osteonecrosis of the Jaw: An Update on Pathophysiology, Risk Factors, and Treatment. Int J Dent 2014; 1-9. ID 471035
[22] Kavanagh KL, Dunford JE, Bunkoczi G, Russell RG, Oppermann U. The crystal structure of human geranylgeranyl pyrophosphate synthase reveals a novel hexameric arrangement and inhibitory product binding. J Biol Chem 2006; 281(31): 22004-12.
[23] Evdokimov A, Pokross M, Barnett BL, et al. Human farnesyl diphosphate synthase crystal structures with active and inactive bisphosphonates. Bone 2006; 38(31): 49-1.
[24] Rondeau JM, Bitsch F, Bourgier E, et al. Structural basis for the exceptional in vivo efficacy of bisphosphonate drugs. Chem Med Chem 2006; 1(2): 267-73.
[25] Guo RT, Cao R, Liang PH, et al. Bisphosphonates target multiple sites in both cis and trans-prenyltransferases. Proc Natl Acad Sci USA 2007; 104(24): 10022-7.
[26] Park J, Lin YS, Tsantrizos YS, Berghuis AM. Structure of human farnesyl pyrophosphate synthase in complex with an aminopyridine bisphosphonate and two molecules of inorganic phosphate. Acta Crystallogr F Struct Biol Commun 2014; 70(Pt 3): 299-304.
[27] Ohno K, Mori K, Orita M, Takeuchi M. Computational insights into binding of bisphosphates to farnesyl pyrophosphate synthase. Curr Med Chem 2011; 18(2): 220-33.
[28] Kawasaki Y, Sekiguchi M, Kawasaki M, Hirakura Y. Thermodynamic evaluation of the binding of bisphosphonates to human farnesyl pyrophosphate synthase. Chem Pharm Bull (Tokyo) 2014; 62(1): 77-83.
[29] Sun S, McKenna CE. Farnesyl pyrophosphate synthase modulators: a patent review (2006 - 2010). Expert Opin Ther Pat 2011; 21(9): 1433-51.
[30] Park J, Lin YS, De Schutter JW, Tsantrizos YS, Berghuis AM. Ternary complex structures of human farnesyl pyrophosphate synthase bound with a novel inhibitor and secondary ligands provide insights into the molecular details of the enzyme’s active site closure. BMC Struct Biol 2012; 12(32): 32.
[31] De Schutter JW, Shaw J, Lin YS, Tsantrizos YS. Design of potent bisphosphonate inhibitors of the human farnesyl pyrophosphate synthase via targeted interactions with the active site ‘capping’ phenyls. Bioorg Med Chem 2012; 20(18): 5583-91.
[32] Lindert S, Zhu W, Liu YL, Pang R, Oldfield E, McCammon JA. Farnesyl diphosphate synthase inhibitors from in silico screening. Chem Biol Drug Des 2013; 81(6): 742-8.
[33] Leung CY, Park J, De Schutter JW, Sebag M, Berghuis AM, Tsantrizos YS. Thienopyrimidine bisphosphonate (ThPBP) inhibitors of the human farnesyl pyrophosphate synthase: optimization and characterization of the mode of inhibition. J Med Chem 2013; 56(20): 7939-50.
[34] Jahnke W, Rondeau JM, Cotesta S, et al. Allosteric non-bisphosphonate FPPS inhibitors identified by fragment-based discovery. Nat Chem Biol 2010; 6(9): 660-6.
[35] Durrant JD, Cao R, Gorfe AA, et al. Non-bisphosphonate inhibitors of isoprenoid biosynthesis identified via computer-aided drug design. Chem Biol Drug Des 2011; 78(3): 323-32.
[36] Fernández D, Ortega-Castro J, Frau J. Human farnesyl pyrophosphate synthase inhibition by nitrogen bisphosphonates: a 3D-QSAR study. J Comput Aided Mol Des 2013; 27(8): 739-54.
[37] Ferrer-Casal M, Li C, Galizzi M, et al. New insights into molecular recognition of 1,1-bisphosphonic acids by farnesyl diphosphate synthase. Bioorg Med Chem 2014; 22(1): 398-405.
[38] McKenna CE, Kashemirov BA, Błazewska KM, et al. Synthesis, chiral high performance liquid chromatographic resolution and enantiospecific activity of a potent new geranylgeranyl transferase inhibitor, 2-hydroxy-3-imidazo[1,2-a]pyridin-3-yl-2-phosphonopropionic acid. J Med Chem 2010; 53(9): 3454-64.
[39] Błażewska KM, Ni F, Haiges R, et al. Synthesis, stereochemistry and SAR of a series of minodronate analogues as RGGT inhibitors. Eur J Med Chem 2011; 46(10): 4820-6.
[40] Stewart A, Baron R, Marma MS, et al. Structure activity relationships of phosphonocarboxylate inhibitors of rab geranylgeranyl transferase. Bone 2008; 42(Suppl. 1): 86-7.
[41] Coxon FP, Joachimiak L, Najumudeen AK, et al. Synthesis and characterization of novel phosphonocarboxylate inhibitors of RGGT. Eur J Med Chem 2014; 84: 77-89.
[42] Shipman CM, Croucher PI, Russell RG, Helfrich MH, Rogers MJ. The bisphosphonate incadronate (YM175) causes apoptosis of human myeloma cells in vitro by inhibiting the mevalonate pathway. Cancer Res 1998; 58(23): 5294-7.
[43] Mukkamala D, No JH, Cass LM, Chang TK, Oldfield E. Bisphosphonate inhibition of a Plasmodium farnesyl diphosphate synthase and a general method for predicting cell-based activity from enzyme data. J Med Chem 2008; 51(24): 7827-33.
[44] K-M Chen C, Hudock MP, Zhang Y, et al. Inhibition of geranylgeranyl diphosphate synthase by bisphosphonates: a crystallographic and computational investigation. J Med Chem 2008; 51(18): 5594-607.
[45] Gao J, Chu X, Qiu Y, et al. Discovery of potent inhibitor for farnesyl pyrophosphate synthase in the mevalonate pathway. Chem Commun (Camb) 2010; 46(29): 5340-2.
[46] No JH, de Macedo Dossin F, Zhang Y, et al. Lipophilic analogs of zoledronate and risedronate inhibit Plasmodium geranylgeranyl diphosphate synthase (GGPPS) and exhibit potent antimalarial activity. Proc Natl Acad Sci USA 2012; 109(11): 4058-63.
[47] Liu YL, Cao R, Wang Y, Oldfield E. Farnesyl diphosphate synthase inhibitors with unique ligand-binding geometries. ACS Med Chem Lett 2015; 6(3): 349-54.
[48] Marma MS, Xia Z, Stewart C, et al. Synthesis and biological evaluation of α-halogenated bisphosphonate and phosphonocarboxylate analogues of risedronate. J Med Chem 2007; 50(24): 5967-75.
[49] Goldeman W, Nasulewicz-Goldeman A. Synthesis and antiproliferative activity of aromatic and aliphatic bis[aminomethylidene(bisphosphonic)] acids. Bioorg Med Chem Lett 2014; 24(15): 3475-9.
[50] Goldeman W, Nasulewicz-Goldeman A. Synthesis and biological evaluation of aminomethylidenebisphosphonic derivatives of β-arylethylamines. Tetrahedron 2015; 71(21): 3282-9.
[51] Chmielewska E, Mazur Z, Kempińska K, et al. N-Arylaminomethylenebisphosphonates bearing fluorine atoms: synthesis and antiosteoporotic activity. Phosphorus Silicon Sulfur Relat Elemen 2015; 190(12): 2164-72.
[52] Chmielewska E, Kempińska K, Wietrzyk J, et al. Novel bisphosphonates and their use. International Patent Application WO2015159153, 2015.
[53] Zhang Y, Cao R, Yin F, et al. Lipophilic pyridinium bisphosphonates: potent gammadelta T cell stimulators. Angew Chem Int Ed Engl 2010; 49(6): 1136-8.
[54] Zhang Y, Cao R, Yin F, et al. Lipophilic bisphosphonates as dual farnesyl/geranylgeranyl diphosphate synthase inhibitors: an X-ray and NMR investigation. J Am Chem Soc 2009; 131(14): 5153-62.
[55] Bortolini O, Fantin G, Fogagnolo M, et al. Synthesis, characterization and biological activity of hydroxyl-bisphosphonic analogs of bile acids. Eur J Med Chem 2012; 52: 221-9.
[56] Vepsäläinen JJ. Bisphosphonate prodrugs. Curr Med Chem 2002; 9(12): 1201-8.
[57] Niemi R, Vepsäläinen J, Järvinen T, Järvinen T. Bisphosphonate prodrugs: synthesis and in vitro evaluation of novel acyloxyalkyl esters of clodronic acid. J Med Chem 1999; 42(24): 5053-8.
[58] Ledoux D, Hamma-Kourbali Y, Di Benedetto M, et al. A new dimethyl ester bisphosphonate inhibits angiogenesis and growth of human epidermoid carcinoma xenograft in nude mice. Anticancer Drugs 2006; 17(4): 479-85.
[59] Monteil M, Migianu-Griffoni E, Sainte-Catherine O, Di Benedetto M, Lecouvey M. Bisphosphonate prodrugs: synthesis and biological evaluation in HuH7 hepatocarcinoma cells. Eur J Med Chem 2014; 77: 56-64.
[60] Joachimiak Ł, Janczewski Ł, Ciekot J, Boratyński J, Błażewska K. Applying the prodrug strategy to α-phosphonocarboxylate inhibitors of Rab GGTase--synthesis and stability studies. Org Biomol Chem 2015; 13(24): 6844-56.
[61] Grossmann G, Grossmann A, Ohms G, et al. Solid-state NMR of bisphosphonates adsorbed on hydroxyapatite. Magn Reson Chem 2000; 38(1): 11-7.
[62] Henneman ZJ, Nancollas GH, Ebetino FH, Russell RG, Phipps RJ. Bisphosphonate binding affinity as assessed by inhibition of carbonated apatite dissolution in vitro. J Biomed Mater Res A 2008; 85(4): 993-1000.
[63] Lawson MA, Xia Z, Barnett BL, et al. Differences between bisphosphonates in binding affinities for hydroxyapatite. J Biomed Mater Res B Appl Biomater 2010; 92(1): 149-55.
[64] Mukherjee S, Song Y, Oldfield E. NMR investigations of the static and dynamic structures of bisphosphonates on human bone: a molecular model. J Am Chem Soc 2008; 130(4): 1264-73.
[65] Juillard A, Falgayrac G, Cortet B, et al. Molecular interactions between zoledronic acid and bone: An in vitro Raman microspectroscopic study. Bone 2010; 47(5): 895-904.
[66] Puljula E, Turhanen P, Vepsäläinen J, Monteil M, Lecouvey M, Weisell J. Structural requirements for bisphosphonate binding on hydroxyapatite: NMR study of bisphosphonate partial esters. ACS Med Chem Lett 2015; 6(4): 397-401.
[67] Fernandes C, Monteiro S, Mendes P, et al. Biological assessment of novel bisphosphonate-containing 99mTc/Re-organometallic complexes. J Organomet Chem 2014; 760: 197-204.
[68] Turek J, Ebetino FH, Lundy MW, et al. Bisphosphonate binding affinity affects drug distribution in both intracortical and trabecular bone of rabbits. Calcif Tissue Int 2012; 90(3): 202-10.
[69] Bae S, Sun S, Aghaloo T, et al. Development of oral osteomucosal tissue constructs in vitro and localization of fluorescently-labeled bisphosphonates to hard and soft tissue. Int J Mol Med 2014; 34(2): 559-63.
[70] Kozloff KM, Volakis LI, Marini JC, Caird MS. Near-infrared fluorescent probe traces bisphosphonate delivery and retention in vivo. J Bone Miner Res 2010; 25(8): 1748-58.
[71] Roelofs AJ, Stewart CA, Sun S, et al. Influence of bone affinity on the skeletal distribution of fluorescently labeled bisphosphonates in vivo. J Bone Miner Res 2012; 27(4): 835-47.
[72] Kurth AH, Eberhardt C, Müller S, Steinacker M, Schwarz M, Bauss F. The bisphosphonate ibandronate improves implant integration in osteopenic ovariectomized rats. Bone 2005; 37(2): 204-10.
[73] Vohra F, Al-Rifaiy MQ, Almas K, Javed F. Efficacy of systemic bisphosphonate delivery on osseointegration of implants under osteoporotic conditions: lessons from animal studies. Arch Oral Biol 2014; 59(9): 912-20.
[74] Javed F, Almas K. Osseointegration of dental implants in patients undergoing bisphosphonate treatment: a literature review. J Periodontol 2010; 81(4): 479-84.
[75] Moaddabi A, Shariati M, Hossein-Moaddabii A, Soltani P. Osseointegration of dental implants in patients with oral bisphosphonate intake: a review. J Craniomax Res 2014; 1(3): 74-7.
[76] Ballantyne E. Bisphosphonates: possible modes of action and implications for dental implant treatment. A review of the literature. J Gent Pract 2014; 3(1): 28.
[77] Samdancioglu S, Calis S, Sumnu M, Atilla Hincal A. Formulation and in vitro evaluation of bisphosphonate loaded microspheres for implantation in osteolysis. Drug Dev Ind Pharm 2006; 32(4): 473-81.
[78] Bobyn JD, McKenzie K, Karabasz D, Krygier JJ, Tanzer M. Locally delivered bisphosphonate for enhancement of bone formation and implant fixation. J Bone Joint Surg Am 2009; 91(Suppl. 6): 23-31.
[79] Yun YP, Kim SE, Kang EY, Kim HJ, Park K, Song HR. The effect of bone morphogenic protein-2 (BMP-2)-immobilizing heparinized-chitosan scaffolds for enhanced osteoblast activity. Tissue Eng Regen Med 2013; 10(3): 122-30.
[80] Liu J, Zhang H, Dong Y, et al. Bi-directionally selective bone targeting delivery for anabolic and antiresorptive drugs: a novel combined therapy for osteoporosis? Med Hypotheses 2014; 83(6): 694-6.
[81] Posadowska U, Parizek M, Filova E, et al. Injectable nanoparticle-loaded hydrogel system for local delivery of sodium alendronate. Int J Pharm 2015; 485(1-2): 31-40.
[82] Aderibigbe BA, Varaprasad K, Sadiku ER, et al. Kinetic release studies of nitrogen-containing bisphosphonate from gum acacia crosslinked hydrogels. Int J Biol Macromol 2015; 73: 115-23.
[83] Ossipov DA. Bisphosphonate-modified biomaterials for drug delivery and bone tissue engineering. Expert Opin Drug Deliv 2015; 12(9): 1443-58.
[84] Kos M, Junka A, Smutnicka D, Bartoszewicz M, Kurzynowski T, Gluza K. Pamidronate enhances bacterial adhesion to bone hydroxyapatite. Another puzzle in the pathology of bisphosphonate-related osteonecrosis of the jaw? J Oral Maxillofac Surg 2013; 71(6): 1010-6.
[85] Kos M, Junka A, Smutnicka D, Szymczyk P, Gluza K, Bartoszewicz M. Bisphosphonates enhance bacterial adhesion and biofilm formation on bone hydroxyapatite. J Craniomaxillofac Surg 2015; 43(6): 863-9.
[86] Conte P, Guarneri V. Safety of intravenous and oral bisphosphonates and compliance with dosing regimens. Oncologist 2004; 9(Suppl. 4): 28-37.
[87] Miller PD. The kidney and bisphosphonates. Bone 2011; 49(1): 77-81.
[88] Hirabayashi H, Fujisaki J. Bone-specific drug delivery systems: approaches via chemical modification of bone-seeking agents. Clin Pharmacokinet 2003; 42(15): 1319-30.
[89] De Rossa G, Misso G, Salzano G, Caraglia M. Bisphosphonates and Cancer: What Opportunities from Nanotechnology? J Drug Deliv 2013; 12: 17.
[90] Eriksen EF, Keaveny TM, Gallagher ER, Krege JH. Literature review: The effects of teriparatide therapy at the hip in patients with osteoporosis. Bone 2014; 67: 246-56.
[91] Yewle JN, Puleo DA, Bachas LG. Bifunctional bisphosphonates for delivering PTH (1-34) to bone mineral with enhanced bioactivity. Biomaterials 2013; 34(12): 3141-9.
[92] Liu J, Jo J, Kawai Y, et al. Preparation of polymer-based multimodal imaging agent to visualize the process of bone regeneration. J Control Release 2012; 157(3): 398-405.
[93] Miller K, Erez R, Segal E, Shabat D, Satchi-Fainaro R. Targeting bone metastases with a bispecific anticancer and antiangiogenic polymer-alendronate-taxane conjugate. Angew Chem Int Ed Engl 2009; 48(16): 2949-54.
[94] Miller K, Eldar-Boock A, Polyak D, et al. Antiangiogenic antitumor activity of HPMA copolymer-paclitaxel-alendronate conjugate on breast cancer bone metastasis mouse model. Mol Pharm 2011; 8(4): 1052-62.
[95] Clementi C, Miller K, Mero A, Satchi-Fainaro R, Pasut G. Dendritic poly(ethylene glycol) bearing paclitaxel and alendronate for targeting bone neoplasms. Mol Pharm 2011; 8(4): 1063-72.
[96] Miller K, Clementi C, Polyak D, et al. Poly(ethylene glycol)-paclitaxel-alendronate self-assembled micelles for the targeted treatment of breast cancer bone metastases. Biomaterials 2013; 34(15): 3795-806.
[97] Bonzi G, Salmaso S, Scomparin A, Eldar-Boock A, Satchi-Fainaro R, Caliceti P. Novel pullulan bioconjugate for selective breast cancer bone metastases treatment. Bioconjug Chem 2015; 26(3): 489-501.
[98] Neville-Webbe HL, Evans CA, Coleman RE, Holen I. Mechanisms of the synergistic interaction between the bisphosphonate zoledronic acid and the chemotherapy agent paclitaxel in breast cancer cells in vitro. Tumour Biol 2006; 27(2): 92-103.
[99] Erez R, Ebner S, Attali B, Shabat D. Chemotherapeutic bone-targeted bisphosphonate prodrugs with hydrolytic mode of activation. Bioorg Med Chem Lett 2008; 18(2): 816-20.
[100] Agyin JK, Santhamma B, Roy SS. Design, synthesis, and biological evaluation of bone-targeted proteasome inhibitors for multiple myeloma. Bioorg Med Chem Lett 2013; 23(23): 6455-8.
[101] Pignatello R. Biomaterials Application for Nanomedicine. Croatia: InTech Rijeka 2011.
[102] El-Mabhouh AA, Angelov CA, Cavell R, Mercer JR Jr. A 99mTc-labeled gemcitabine bisphosphonate drug conjugate as a probe to assess the potential for targeted chemotherapy of metastatic bone cancer. Nucl Med Biol 2006; 33(6): 715-22.
[103] Anada T, Takeda Y, Honda Y, Sakurai K, Suzuki O. Synthesis of calcium phosphate-binding liposome for drug delivery. Bioorg Med Chem Lett 2009; 19(15): 4148-50.
[104] Wang G, Babadağli ME, Uludağ H. Bisphosphonate-derivatized liposomes to control drug release from collagen/hydroxyapatite scaffolds. Mol Pharm 2011; 8(4): 1025-34.
[105] Wang G, Mostafa NZ, Incani V, Kucharski C, Uludağ H. Bisphosphonate-decorated lipid nanoparticles designed as drug carriers for bone diseases. J Biomed Mater Res A 2012; 100(3): 684-93.
[106] Song H, Zhang J, Liu X, et al. Development of a bone targeted thermosensitive liposomal doxorubicin formulation based on a bisphosphonate modified non-ionic surfactant. Pharm Dev Technol 2015; 1-8.
[107] Paolino D, Licciardi M, Ceila C, Giammona G, Fresta M, Cavallaro G. Bisphosphonate–polyaspartamide conjugates as bone targeted drug delivery systems. J Mater Chem B Mater Biol Med 2015; 3: 250-9.
[108] Pignatello R, Sarpietro MG, Castelli F. Synthesis and biological evaluation of a new polymeric conjugate and nanocarrier with osteotropic properties. J Funct Biomater 2012; 3(1): 79-99.
[109] Rudnick-Glick S, Corem-Salkmon E, Grinberg I, Gluz E, Margel S. Doxorubicin-conjugated bisphosphonate nanoparticles for the therapy of osteosarcoma. JSM Nanotechnol Nanomed 2014; 2(2): 1022-30.
[110] Swami A, Reagan MR, Basto P, et al. Engineered nanomedicine for myeloma and bone microenvironment targeting. Proc Natl Acad Sci USA 2014; 111(28): 10287-92.
[111] Cole LE, Vargo-Gogola T, Roeder RK. Bisphosphonate-functionalized gold nanoparticles for contrast-enhanced X-ray detection of breast microcalcifications. Biomaterials 2014; 35(7): 2312-21.
[112] Bordoloi JK, Berry D, Khan IU, et al. Technetium-99m and rhenium-188 complexes with one and two pendant bisphosphonate groups for imaging arterial calcification. Dalton Trans 2015; 44(11): 4963-75.
[113] McPherson JC III, Runner R, Buxton TB, et al. Synthesis of osteotropic hydroxybisphosphonate derivatives of fluoroquinolone antibacterials. Eur J Med Chem 2012; 47(1): 615-8.
[114] Houghton TJ, Tanaka KS, Kang T, et al. Linking bisphosphonates to the free amino groups in fluoroquinolones: preparation of osteotropic prodrugs for the prevention of osteomyelitis. J Med Chem 2008; 51(21): 6955-69.
[115] Cong Y, Quan C, Liu M, et al. Alendronate-decorated biodegradable polymeric micelles for potential bone-targeted delivery of vancomycin. J Biomat Sci Ed 2015; 26(11): 629-43.
[116] Reddy R, Dietrich E, Lafontaine Y, et al. Bisphosphonated benzoxazinorifamycin prodrugs for the prevention and treatment of osteomyelitis. Chem Med Chem 2008; 3(12): 1863-8.
[117] Ramaswamy K, Marx V, Laser D, et al. Targeted microbubbles: a novel application for the treatment of kidney stones. BJU Int 2015; 116(1): 9-16.
[118] Sebti SM. Protein farnesylation: implications for normal physiology, malignant transformation, and cancer therapy. Cancer Cell 2005; 7(4): 297-300.
[119] Iguchi K, Tatsuda Y, Usui S, Hirano K. Pamidronate inhibits antiapoptotic bcl-2 expression through inhibition of the mevalonate pathway in prostate cancer PC-3 cells. Eur J Pharmacol 2010; 641(1): 35-40.
[120] Dedes PG, Gialeli Ch, Tsonis AI, et al. Expression of matrix macromolecules and functional properties of breast cancer cells are modulated by the bisphosphonate zoledronic acid. Biochim Biophys Acta 2012; 1820(12): 1926-39.
[121] Mathew A, Brufsky A. Bisphosphonates in breast cancer. Int J Cancer 2015; 137(4): 753-64.
[122] Garay T, Kenessey I, Molnár E, et al. Prenylation inhibition-induced cell death in melanoma: reduced sensitivity in BRAF mutant/PTEN wild-type melanoma cells. PLoS One 2015; 10(2): 753-64. e0117021
[123] Rennert G, Rennert HS, Pinchev M, Lavie O. The effect of bisphosphonates on the risk of endometrial and ovarian malignancies. Gynecol Oncol 2014; 133(2): 309-13.
[124] Notarnicola M, Messa C, Cavallini A, et al. Higher farnesyl diphosphate synthase activity in human colorectal cancer inhibition of cellular apoptosis. Oncology 2004; 67(5-6): 351-8.
[125] Cimini E, Piacentini P, Sacchi A, et al. Zoledronic acid enhances Vδ2 T-lymphocyte antitumor response to human glioma cell lines. Int J Immunopathol Pharmacol 2011; 24(1): 139-48.
[126] Terpos E, Roodman GD, Dimopoulos MA. Optimal use of bisphosphonates in patients with multiple myeloma. Blood 2013; 121(17): 3325-8.
[127] Ben-Aharon I, Vidal L, Rizel S, et al. Bisphosphonates in the adjuvant setting of breast cancer therapy-effect on survival: a systematic review and meta-analysis. PloS One 2013; 8(8): e70044-9.
[128] Li BT, Wong MH, Pavlakis N. Treatment and prevention of bone metastases from breast cancer: a comprehensive review of evidence for clinical practice. J Clin Med 2014; 3(1): 1-24.
[129] Jacobs C, Amir E, Paterson A, Zhu X, Clemons M. Are adjuvant bisphosphonates now standard of care of women with early stage breast cancer? A debate from the Canadian Bone and the Oncologist New Updates meeting. J Bone Oncol 2015; 4(2): 54-8.
[130] Liu Y, Du C, Zhang Y, et al. Bisphosphonate and risk of cancer recurrence: protocol for a systematic review, meta-analysis and trial sequential analysis of randomised controlled trials. BMJ Open 2015; 5(4): e007215-6.
[131] Stachnik A, Yuen T, Iqbal J, et al. Repurposing of bisphosphonates for the prevention and therapy of nonsmall cell lung and breast cancer. Proc Natl Acad Sci USA 2014; 111(50): 17995-8000.
[132] Sanders JM, Ghosh S, Chan JM, et al. Quantitative structure-activity relationships for gammadelta T cell activation by bisphosphonates. J Med Chem 2004; 47(2): 375-84.
[133] Thompson K, Roelofs AJ, Jauhiainen M, Mönkkönen H, Mönkkönen J, Rogers MJ. Activation of γδ T cells by bisphosphonates. Adv Exp Med Biol 2010; 658: 11-20. [In: Osteoimmunology. Interactions of the Immune and Skeletal Systems II. Choi J Ed; Springer].
[134] Vassiliou V. Management of metastatic bone disease in the elderly with bisphosphonates and receptor activator of NF-kB ligand inhibitors: effectiveness and safety. Clin Oncol (R Coll Radiol) 2013; 25(5): 290-7.
[135] Tang X, Zhang Q, Shi S, et al. Bisphosphonates suppress insulin-like growth factor 1-induced angiogenesis via the HIF-1alpha/VEGF signaling pathways in human breast cancer cells. Int J Cancer 2010; 126(1): 90-103.
[136] Yuen T, Stachnik A, Iqbal J, et al. Bisphosphonates inactivate human EGFRs to exert antitumor actions. Proc Natl Acad Sci USA 2014; 111(50): 17989-94.
[137] Karlic H, Thaler R, Gerner C, et al. Inhibition of the mevalonate pathway affects epigenetic regulation in cancer cells. Cancer Genet 2015; 208(5): 241-52.
[138] Saad A, Zhu W, Rousseau G, et al. Polyoxomolybdate bisphosphonate heterometallic complexes: synthesis, structure, and activity on a breast cancer cell line. Chemistry 2015; 21(29): 10537-47.
[139] Nishiguchi T, Akiyoshi T, Anami S, Nakabayashi T, Matsuyama K, Matzno S. Synergistic action of statins and nitrogen-containing bisphosphonates in the development of rhabdomyolysis in L6 rat skeletal myoblasts. J Pharm Pharmacol 2009; 61(6): 781-8.
[140] Santini D, Caraglia M, Vincenzi B, et al. Mechanisms of disease: Preclinical reports of antineoplastic synergistic action of bisphosphonates. Nat Clin Pract Oncol 2006; 3(6): 325-38.
[141] Mathavan N, Bosemark P, Isaksson H, Tägil M. Investigating the synergistic efficacy of BMP-7 and zoledronate on bone allografts using an open rat osteotomy model. Bone 2013; 56(2): 440-8.
[142] Dudakovic A, Wiemer AJ, Lamb KM, Vonnahme LA, Dietz SE, Hohl RJ. Inhibition of geranylgeranyl diphosphate synthase induces apoptosis through multiple mechanisms and displays synergy with inhibition of other isoprenoid biosynthetic enzymes. J Pharmacol Exp Ther 2008; 324(3): 1028-36.
[143] Reilly JE, Zhou X, Tong H, Kuder CH, Wiemer DF, Hohl RJ. In vitro studies in a myelogenous leukemia cell line suggest an organized binding of geranylgeranyl diphosphate synthase inhibitors. Biochem Pharmacol 2015; 96(2): 83-92.
[144] Yoshida K, Hiratsuka J. Palliative radiotherapy for metastatic bone tumor. Clin Calcium 2006; 16(4): 641-5.
[145] Smith HS. Painful boney metastases. Ann Palliat Med 2012; 1(1): 14-31.
[146] Santangelo A, Testai M, Barbagallo P, et al. The use of bisphosphonates in palliative treatment of bone metastases in a terminally ill, oncological elderly population. Arch Gerontol Geriatr 2006; 43(2): 187-92.
[147] Vogel CL, Yanagihara RH, Wood AJ, et al. Safety and pain palliation of zoledronic acid in patients with breast cancer, prostate cancer, or multiple myeloma who previously received bisphosphonate therapy. Oncologist 2004; 9(6): 687-95.
[148] Nikzad M, Jalilian AR, Shirvani-Arani S, Bahrami-Samani A, Golchoubian H. Production, quality control and pharmacokinetic studies of 177Lu-zoledronate for bone pain palliation therapy. J Radioanal Nucl Chem 2013; 298(2): 1273-81.
[149] Di Franco R, Calvanese M, Cuomo M, et al. Management of painful bone metastases: the interaction between radiation therapy and zoledronate. J Cancer Ther 2011; 2(5): 697.
[150] Littlejohn G. Therapy: Bisphosphonates for early complex regional pain syndrome. Nat Rev Rheumatol 2013; 9(4): 199-200.
[151] Varenna M, Adami S, Sinigaglia L. Bisphosphonates in Complex Regional Pain syndrome type I: how do they work? Clin Exp Rheumatol 2014; 32(4): 451-4.
[152] Feasey N, Wansbrough-Jones M, Mabey DC, Solomon AW. Neglected tropical diseases. Br Med Bull 2010; 93(1): 179-200.
[153] Cohen JM, Smith DL, Cotter C, et al. Malaria resurgence: a systematic review and assessment of its causes. Malar J 2012; 11(11): 122.
[154] Martin MB, Grimley JS, Lewis JC, et al. Bisphosphonates inhibit the growth of Trypanosoma brucei, Trypanosoma cruzi, Leishmania donovani, Toxoplasma gondii, and Plasmodium falciparum: a potential route to chemotherapy. J Med Chem 2001; 44(6): 909-16.
[155] Sanders JM, Gómez AO, Mao J, et al. 3-D QSAR investigations of the inhibition of Leishmania major farnesyl pyrophosphate synthase by bisphosphonates. J Med Chem 2003; 46(24): 5171-83.
[156] Martin MB, Sanders JM, Kendrick H, et al. Activity of bisphosphonates against Trypanosoma brucei rhodesiense. J Med Chem 2002; 45(14): 2904-14.
[157] Liu YL, Lindert S, Zhu W, Wang K, McCammon JA, Oldfield E. Taxodione and arenarone inhibit farnesyl diphosphate synthase by binding to the isopentenyl diphosphate site. Proc Natl Acad Sci USA 2014; 111(25): E2530-9.
[158] Shubar HM, Mayer JP, Hopfenmüller W, Liesenfeld O. A new combined flow-cytometry-based assay reveals excellent activity against Toxoplasma gondii and low toxicity of new bisphosphonates in vitro and in vivo. J Antimicrob Chemother 2008; 61(5): 1110-9.
[159] Ziniel PD, Desai J, Cass CL, Gatto C, Oldfield E, Williams DL. Characterization of potential drug targets farnesyl diphosphate synthase and geranylgeranyl diphosphate synthase in Schistosoma mansoni. Antimicrob Agents Chemother 2013; 57(12): 5969-76.
[160] Artz JD, Dunford JE, Arrowood MJ, et al. Targeting a uniquely nonspecific prenyl synthase with bisphosphonates to combat cryptosporidiosis. Chem Biol 2008; 15(12): 1296-306.
[161] Cao R, Chen CK, Guo RT, Wang AH, Oldfield E. Structures of a potent phenylalkyl bisphosphonate inhibitor bound to farnesyl and geranylgeranyl diphosphate synthases. Proteins 2008; 73(2): 431-9.
[162] Mukkamala D, No JH, Cass LM, Chang TK, Oldfield E. Bisphosphonate inhibition of a Plasmodium farnesyl diphosphate synthase and a general method for predicting cell-based activity from enzyme data. J Med Chem 2008; 51(24): 7827-33.
[163] Szajnman SH, García Liñares GE, Li ZH, et al. Synthesis and biological evaluation of 2-alkylaminoethyl-1,1-bisphosphonic acids against Trypanosoma cruzi and Toxoplasma gondii targeting farnesyl diphosphate synthase. Bioorg Med Chem 2008; 16(6): 3283-90.
[164] Huang CH, Gabelli SB, Oldfield E, Amzel LM. Binding of nitrogen-containing bisphosphonates (N-BPs) to the Trypanosoma cruzi farnesyl diphosphate synthase homodimer. Proteins 2010; 78(4): 888-99.
[165] Singh AP, Zhang Y, No JH, Docampo R, Nussenzweig V, Oldfield E. Lipophilic bisphosphonates are potent inhibitors of Plasmodium liver-stage growth. Antimicrob Agents Chemother 2010; 54(7): 2987-93.
[166] Rosso VS, Szajnman SH, Malayil L, et al. Synthesis and biological evaluation of new 2-alkylaminoethyl-1,1-bisphosphonic acids against Trypanosoma cruzi and Toxoplasma gondii targeting farnesyl diphosphate synthase. Bioorg Med Chem 2011; 19(7): 2211-7.
[167] Srivastava A, Mukherjee P, Desai PV, Avery MA, Tekwani BL. Structural analysis of farnesyl pyrophosphate synthase from parasitic protozoa, a potential chemotherapeutic target. Infect Disord Drug Targets 2008; 8(1): 16-30.
[168] Sanz-Rodríguez CE, Concepción JL, Pekerar S, Oldfield E, Urbina JA. Bisphosphonates as inhibitors of Trypanosoma cruzi hexokinase: kinetic and metabolic studies. J Biol Chem 2007; 282(17): 12377-87.
[169] Shang N, Li Q, Ko TP, et al. Squalene synthase as a target for Chagas disease therapeutics. PLOS Pathog 2014; 10(5): e1004114-.
[170] Docampo R, Moreno SN. The acidocalcisome as a target for chemotherapeutic agents in protozoan parasites. Curr Pharm Des 2008; 14(9): 882-8.
[171] Rogers MJ, Ji X, Russell RG, et al. Incorporation of bisphosphonates into adenine nucleotides by amoebae of the cellular slime mould Dictyostelium discoideum. Biochem J 1994; 303(Pt 1): 303-11.
[172] Rogers MJ, Xiong X, Ji X, et al. Inhibition of growth of Dictyostelium discoideum amoebae by bisphosphonate drugs is dependent on cellular uptake. Pharm Res 1997; 14(5): 625-30.
[173] Nuttall JM, Hettema EH, Watts DJ. Farnesyl diphosphate synthase, the target for nitrogen-containing bisphosphonate drugs, is a peroxisomal enzyme in the model system Dictyostelium discoideum. Biochem J 2012; 447(3): 353-61.
[174] Ghosh S, Chan JM, Lea CR, et al. Effects of bisphosphonates on the growth of Entamoeba histolytica and Plasmodium species in vitro and in vivo. J Med Chem 2004; 47(1): 175-87.
[175] Ondraza RN. Drug effects on drug targets: inhibition of enzymes by neuroleptics, antimycotics, antibiotics and other drugs on human pathogenic amoebas and their antiproliferative effects. Recent Patents Anti-Infect Drug Disc 2007; 2(3): 206-16.
[176] Davies J, Davies D. Origins and evolution of antibiotic resistance. Microbiol Mol Biol Rev 2010; 74(3): 417-33.
[177] Leon A, Liu L, Yang Y, et al. Isoprenoid biosynthesis as a drug target: bisphosphonate inhibition of Escherichia coli K12 growth and synergistic effects of fosmidomycin. J Med Chem 2006; 49(25): 7331-41.
[178] Zhang Y, Fu-Yang Lin , Li K, et al. HIV-1 integrase inhibitor-inspired antibacterials targeting isoprenoid biosynthesis. ACS Med Chem Lett 2012; 3(5): 402-6.
[179] Teng KH, Liang PH. Structures, mechanisms and inhibitors of undecaprenyl diphosphate synthase: a cis-prenyltransferase for bacterial peptidoglycan biosynthesis. Bioorg Chem 2012; 43: 51-7.
[180] Kopernyk IM, Kalashnikova LE, Golod LP, Metelitza LO. Antibacterial properties of nitrogen-containing bisphosphonates as farnesyl diphosphate synthases inhibitors. Ukrain Bioorg Acta 2013; 11(2): 8-11.
[181] Schmidberger JW, Schnell R, Schneider G. Structural characterization of substrate and inhibitor binding to farnesyl pyrophosphate synthase from Pseudomonas aeruginosa. Acta Crystallogr D Biol Crystallogr 2015; 71(Pt 3): 721-31.
[182] Zhu W, Zhang Y, Sinko W, et al. Antibacterial drug leads targeting isoprenoid biosynthesis. Proc Natl Acad Sci USA 2013; 110(1): 123-8.
[183] Sinko W, de Oliveira C, Williams S, et al. Applying molecular dynamics simulations to identify rarely sampled ligand-bound conformational states of undecaprenyl pyrophosphate synthase, an antibacterial target. Chem Biol Drug Des 2011; 77(6): 412-20.
[184] Chan HC, Feng X, Ko TP, et al. Structure and inhibition of tuberculosinol synthase and decaprenyl diphosphate synthase from Mycobacterium tuberculosis. J Am Chem Soc 2014; 136(7): 2892-6.
[185] Zhu W, Wang Y, Li K, et al. Antibacterial drug leads: DNA and enzyme multitargeting. J Med Chem 2015; 58(3): 1215-27.
[186] Beaton SA, Jiang PM, Melong JC, et al. The effect of bisphosphonate acidity on the activity of a thymidylyltransferase. Org Biomol Chem 2013; 11(33): 5473-80.
[187] Forlani G, Petrollino D, Fusetti M, et al. Δ1-pyrroline-5-carboxylate reductase as a new target for therapeutics: inhibition of the enzyme from Streptococcus pyogenes and effects in vivo. Amino Acids 2012; 42(6): 2283-91.
[188] Kosikowska P, Bochno M, Macegoniuk K, Forlani G, Kafarski P, Berlicki Ł. Bisphosphonic acids as effective inhibitors of Mycobacterium tuberculosis glutamine synthetase. J Enz Inh Med Chem 2015; 1-8. [Epub ahead of print].
[189] McPherson JC III, Runner R, Buxton TB, et al. Synthesis of osteotropic hydroxybisphosphonate derivatives of fluoroquinolone antibacterials. Eur J Med Chem 2012; 47(1): 615-8.
[190] Anisenko A, Agapkina J, Zatsepin T, Yanvarev D, Gottikh M. A new fluorometric assay for the study of DNA-binding and 3′-processing activities of retroviral integrases and its use for screening of HIV-1 integrase inhibitors. Biochimie 2012; 94(11): 2382-90.
[191] Lacbay CM, Mancuso J, Lin YS, Bennett N, Götte M, Tsantrizos YS. Modular assembly of purine-like bisphosphonates as inhibitors of HIV-1 reverse transcriptase. J Med Chem 2014; 57(17): 7435-49.
[192] Agapkina J, Yanvarev D, Anisenko A, et al. Specific features of HIV-1 integrase inhibition by bisphosphonate derivatives. Eur J Med Chem 2014; 73: 73-82.
[193] Song Y, Chan JM, Tovian Z, et al. Bisphosphonate inhibitors of ATP-mediated HIV-1 reverse transcriptase catalyzed excision of chain-terminating 3′-azido, 3′-deoxythymidine: a QSAR investigation. Bioorg Med Chem 2008; 16(19): 8959-67.
[194] Tan KS, Ng WC, Seet JE, Olfat F, Engelward BP, Chow VT. Investigating the efficacy of pamidronate, a chemical inhibitor of farnesyl pyrophosphate synthase, in the inhibition of influenza virus infection in vitro and in vivo. Mol Med Rep 2014; 9(1): 51-6.
[195] Kafarski P, Forlani G, Lejczak B. Herbicidally active aminomethylenebisphosphonic acids. Heteroatom Chem 2000; 11(7): 449-53.
[196] Oberhauser V, Gaudin J, Fonné-Pfister R, Schär HP. New target enzyme(s) for bisphosphonates: inhibition of geranylgeranyl diphosphate synthase. Pestic Biochem Physiol 1998; 60(2): 111-7.
[197] Cromartie TH, Fisher KJ, Grossmann JN. The discovery of a novel site of action for herbicidal bisphosphonates. Pestic Biochem Physiol 1999; 63(2): 114-26.
[198] Forlani G, Lejczak B, Kafarski P. The herbicidally active compound N-2-(5-chloro-pyridyl) aminomethylene bisphosphonic acid acts by inhibiting both glutamine and aromatic amino acid biosynthesis. Aust J Plant Physiol 2000; 27(7): 677-83.
[199] Occhipinti A, Berlicki Ł, Giberti S, Dziedzioła G, Kafarski P, Forlani G. Effectiveness and mode of action of phosphonate inhibitors of plant glutamine synthetase. Pest Manag Sci 2010; 66(1): 51-8.
[200] Forlani G, Lejczak B, Kafarski P. The herbicidally active compound n-2-(6-methyl-pyridyl)-aminomethylene bisphosphonic acid inhibits in vivo aromatic biosynthesis. J Plant Growth Regul 1999; 18(2): 73-9.
[201] Forlani G, Giberti S, Berlicki L, Petrollino D, Kafarski P. Plant P5C reductase as a new target for aminomethylenebisphosphonates. J Agric Food Chem 2007; 55(11): 4340-7.
[202] Forlani G, Occhipinti A, Berlicki L, Dziedzioła G, Wieczorek A, Kafarski P. Tailoring the structure of aminobisphosphonates to target plant P5C reductase. J Agric Food Chem 2008; 56(9): 3193-9.
[203] Szabo CM, Oldfield E. An investigation of bisphosphonate inhibition of a vacuolar proton-pumping pyrophosphatase. Biochem Biophys Res Commun 2001; 287(2): 468-73.
[204] Chang TH, Hsieh FL, Ko TP, Teng KH, Liang PH, Wang AH. Structure of a heterotetrameric geranyl pyrophosphate synthase from mint (Mentha piperita) reveals intersubunit regulation. Plant Cell 2010; 22(2): 454-67.
[205] Vishwakarma RK, Patel KA, Sonawane P, et al. Molecular characterization of farnesyl pyrophosphate synthase from Bacopa monniera by comparative modeling and docking studies. Bioinformation 2012; 8(22): 1075-81.
[206] Manzano AI, Medina FJ, Pérez-Zuniga FJ. Effect of bisphosphonates on root growth and on chlorophyll formation in Arabidopsis thaliana seedlings. In: Osteoporosis. Croatia: InTech, Rijeka 2012; p. 13.
[207] Rasulov B, Talts E, Kännaste A, Niinemets Ü. Bisphosphonate inhibitors reveal a large elasticity of plastidic isoprenoid synthesis pathway in isoprene-emitting hybrid aspen. Plant Physiol 2015; 168(2): 532-48.
[208] Halimaa P, Lin YF, Ahonen VH, et al. Gene expression differences between Noccaea caerulescens ecotypes help to identify candidate genes for metal phytoremediation. Environ Sci Technol 2014; 48(6): 3344-53.
[209] Alanne AL, Peräniemi S, Turhanen P, Tuomainen M, Vepsäläinen J, Tervahauta A. A bisphosphonate increasing the shoot biomass of the metal hyperaccumulator Noccaea caerulescens. Chemosphere 2014; 95: 566-71.
[210] Au KM, Satterlee A, Min Y, et al. Folate-targeted pH-responsive calcium zoledronate nanoscale metal-organic frameworks: Turning a bone antiresorptive agent into an anticancer therapeutic. Biomaterials 2016; 82: 178-93.