BOTELLA Rivero-Buceta - Safe Approaches For Camptothecin Delivery - Structural Analogues and Nanom...
BOTELLA Rivero-Buceta - Safe Approaches For Camptothecin Delivery - Structural Analogues and Nanom...
http://hdl.handle.net/10251/148360
Copyright Elsevier
Additional Information
Safe approaches for camptothecin delivery: structural analogues and
nanomedicines
activity over a wide spectrum of human cancers. However, the direct formulation is
limited by its poor water solubility, low plasmatic stability and severe toxicity, which
currently limits its clinical use. As a consequence, two strategies have been developed
in order to achieve safe and efficient delivery of camptothecin to target cells: structural
toxicity. Nanocarriers which respond to a variety of stimuli endogenously (e.g., pH, redox
ultrasound) seem the best positioned therapeutic materials for optimal spatial and
temporal control over drug release. The main goal of this review is to be used as a source
therapeutics.To this end, final remarks on the most important formulations currently
Contents
1. Introduction
2. Camptothecin analogues
2
2.2. Camptothecin structural derivatives at the clinical stage
3.1.1. Micelles
3.1.2. Liposomes
3.1.3. Dendrimers
4. Stimuli-responsive systems
5. Clinical testing
3
6. Conclusions and future directions
Acknowledgements
References
1. Introduction
Camptothecin (CPT) is a water insoluble, natural pentacyclic alkaloid isolated from the
oriental tree Camptotheca accuminata by Wall et al. in 1966 [1]. CPT is well-known for
its antitumor activity against a wide spectrum of human cancers [2-8]. In addition, a
[13], antifungal [14], antimicrobial [15] and antiviral [16-17] has been described for this
compound. Certainly, several researchers have reported that CPT inhibits a cellular
enzyme DNA topoisomerase I and induces apoptosis in various cancer cells. [18-19]
CPT is a planar pentacyclic quinoline that includes 3 rings of pyrrolo- (3,4-β)- quinoline
part (A, B, and C rings), an unsaturated pyridone moiety in ring D and a α-hydroxy lactone
ring (E ring), containing one chiral center with (S)-configuration (Fig. 1) [20-21]. It is worth
noting that the isomer 20(R) hydroxyl has little activity while the isomer 20(S) is between
10 and 100 fold more active [22]. Also, it is known that E-ring exists in equilibrium
between the lactone form (closed ring, not water insoluble) and the carboxylate form
(open ring, water soluble). In acidic pH the lactone predominates which, according to
different clinical trials and structure-activities studies, is the only therapeutic molecule.
O O
5 OH
7
9 N 22 N
C
B D OH O
10
A E O
N 20
N
14 H
11 S´ O
12 O
OH OH
Fig. 1. CPT equilibrium between lactone (active form) and carboxylate (inactive form)
4
However, at physiological pH, the lactone ring is converted into the carboxylate, much
application, like poor water solubility and the rapid lactone ring hydrolysis at physiological
pH, which gives rise to the inactive carboxylate form [25]. In addition, CPT is extremely
insoluble in organic compounds except for dimethyl sulfoxide, in which it shows moderate
solubility [26]. Due to CPT insolubility in biocompatible solvents, it is very difficult to apply
injection, to distribute this compound throughout the body [27]. Furthermore, there are
other negatives aspects that limit the use of CPT in clinical trials: pronounced loss of
toxicity, including hemorrhagic cystitis and myelotoxicity [28]. All these drawbacks have
topotecan and irinotecan have been developed [29-30] (see Table 1). Unfortunately, as
it will be reported later on, CPT quinoline structure modification usually brings out a
dramatic drop of cytotoxic activity, which results one order or even lower than pristine
To overcome these issues two strategies have been pointed out for safety and efficient
CPT delivery to target cells: i) structural analogues, in which the CPT molecule is
chemically modified for increased solubility and stability in biological fluids; and ii)
nanoparticles which act as stable carriers for drug delivery. In this review, we summarize
the most advanced and recent achievements in the design, synthesis and development
of CPT analogues and nanomedicines, including the most relevant formulations currently
5
2. Campthotecin analogues
Since the discovery of CPT and the associated therapeutic limitations, huge scientific
efforts are being made to improve the pharmacokinetics, drug resistance, clinical
efficacy, and toxicity profiles of the original molecule, by introducing organic ligands [31-
36].
(Table 1 and Fig. 2). It was reported that the complete pentacyclic ring structure was
essential for its activity and the planar structure of this system was suggested to be one
of the most influential structural element. In addition, the hydroxylactone ring is the most
Most derivatives of CPT have been obtained through modifications of the quinolone A-B
ring. However, to date, only two of these CPT analogues have been approved for clinical
general, substitution at C12 is unfavorable to biological activity. This loss in the activity
can be due to steric and electronic disturbances at the quinoline nitrogen, which might
interfere interaction with DNA [44]. Monosubstitution at C9, C10 or C11 by amine or
hydroxyl group make it possible to obtain compounds with more antitumoral activity,
substitution of position 11 with fluorine or cyano groups increase the DNA topo I inhibition
ability [44-45]. Moreover, modifications at C10 and C11 positions are adverse for
6
positions 7 and 9 do not positively affect DNA topo I activity, whereas substitutions at C7
Irinotecanc O
H N N H CH2CH3 1.14 290-320 Approved [30,60]
(CPT-11) O
Exatecan NH
F CH3 __ 0.008 0.3-0.5d Phase III [34,70]
(DX-8951f)
Lurtotecan O O
__ H N NCH3 0.006 1.2 Phase II [34,71]
(GG-211)
N
N
O
O N
N -
O +O
Irinotecan O N
OH O O
N
HO N
O
N O
N N
Rubitecan
NH O OH O
O
SN-38 N
O OH O
N
N C9 O
N C7 + C10
O OH O
C7 10-(4-pyridyl)-CPT
7 5 C10
OH O 9 O
Belotecan
10 C N
A B 16 22
D N
C5+C9 N
11 E O C9+C10
OH 12 1 14 HO
20 O
O N
N C10 + C11 R1 O N
N C11 O
O Topotecan
O O OH O
OH O N
5-ethyl-9-hydroxy-CPT O N O
O N
R1-4 N
10,11-methylenedioxy-CPT
OH O O
R1=OH OH O
O
O
N R2= F
O N R3=CN
O R4=NH2
10,11-ethylenedioxy-CPT
OH O
Fig. 2. Scheme of different CPT structural modifications. Where possible, the structure
of the resulting analogue is depicted.
7
Also, a series of 7-substituted CPTs were developed to overcome the instability of
lactone ring. All of them showed significantly improved cytotoxic/antitumor potency and
stability [50]. On the other hand, few studies have focused on modifications of C and D
as CPT loses its planarity. Substitution of alkoxy or several other groups at position 5 are
reasonably well tolerated when the substituents at position 9 are nitro or hydroxyl groups
At this point, it must be emphasized that E ring and stereochemistry at 20 position are
crucial for CPT activity. One of the major drawbacks observed in ring E modification is
that, due to the presence of α-OH group, under physiological conditions the lactone ring
is cleavaged to inactive carboxylate form. Here, with the aim of increasing lactone ring
stability, several derivatives have been reported but with lower cytotoxicity. Also, it is
More recent studies have been focused on different conjugates (ester, amide, carbonate,
etc.) at C20 position aimed at new CPT structures. As it will be detailed later on (section
3.1), CPT can be covalently linked by 20-hydroxyl group to numerous organic moieties,
Actually, these systems improve CPT delivery and bioavailability, and may also introduce
Although a variety of CPT derivatives have been developed, only a short list has been
investigated in clinical trials [59]. Herein, we try to mention the most representative CPT-
8
based molecules at the clinical stage. Such illustrative examples are presented in Table
1.
In particular, CPT-11 (also known as irinotecan) was approved by the Food and Drug
Administration (FDA) and is used to treat colorectal and other cancers. In vivo, CPT-11
is converted into the active metabolite SN38 by the hepatic carboxylesterase enzyme.
toxicity and neutropenia [60]. Furthermore, SN-38 is one of the most active analogues
and shows higher cytotoxic activity than irinotecan [61-62]. However, the principle issue
Topotecan was the first topoisomerase I inhibitor approved by the FDA and is used to
treat ovarian and lung cancers. This compound has shown activity against a broad
spectrum of cancers cell lines. A report about Phase II testing [64], mentions that the
92 patients. Although this compound was tolerated but different side effects such as
analogue currently in Phase III trial. This compound can be metabolically converted in
immunodeficient (nude) mice of the NIH-1 high fertility strain inoculated with different
several-fold lower than CPT [65]. Also, rubitecan, showed similar activity than CPT
against a broad spectrum of tumor types in human tumor xenograft models [66], and in
a Phase II trial over 58 advanced pancreatic carcinoma patients, it achieved >50% tumor
shrinkage in 7 cases (16%), being well tolerated by most of patients (only 3 patients
the problem of water solubility. This compound showed more activity than topotecan
9
against a broad spectrum of cancers cell lines and acceptable toxicity [68]. Recently,
different clinical studies of belotecan in combination with other chemotherapy drugs such
demonstrated a promising efficacy in patients with advanced stage small cell lung
cancer, although the hematologic toxicity of this regimen requires substantial attention.
Exatecan (DX-8951f) is an intrinsically active compound also in Phase II. This CPT
analogue showed higher efficacy against human tumor xenografts (colon, lung, breast,
renal) than topotecan and irinotecan. [70]. In an experiment that included over 31
21-day continuous i.v. infusion showed limited systemic toxicity (with a significant
anticancer activity.
Finally, lurtotecan (GG-211) is also a water soluble CPT analogue currently in Phase II.
This compound showed significant cytotoxicity against HT-29 and SW-48 human colon
cancer cells, PC-3 human prostate cancer cell line, MX-1 human breast cancer cells,
H460 human lung cancer cell line, SKOV3 human ovarian cancer cells and KB human
Many types of nanovehicles have been developed for CPT delivery, and these drug
delivery systems (DDSs) may be classified in several ways, according to their chemical
and hybrid materials, presenting different properties and behaviors [72-75]. These novel
nanoplatforms have been designed to solve most of the antitumor drug limitations,
improving CPT pharmacokinetics, which is affected by poor water solubility, rapid plasma
10
clearance, high systemic toxicity and poor selectivity towards cancer cells [36]. A variety
of DDSs have been developed and allow CPT incorporation. In addition, these DDSs
bonding [76], working as inert supports to transport the drug to tumor [77]. In this sense,
some major advantages of these DDSs over free CPT are: improved solubility, lactone
ring stability, half-life extension, biocompatibility and control drug release rates, which
The use of organic pharmaceutical nanocarriers to enhance the in vivo efficacy of many
drugs has been well established in the last two decades [80]. To prepare such smart
the surface [81]. Organic carriers are carbon-based (with the exception of carbon
nanotubes that are considered inorganic) and are generally characterized by their high
However, their usual limitations include restricted functionalization (with hinders not only
drug loading, but also the incorporation of targeting molecules and cromophores), low
stability (especially under enzymatic activity), and the lack of specific physicochemical
properties, as optical activity or magnetism, which are useful in diagnostics [82]. In these
Although different classifications may be used, we have organized the various organic
nanocontainers according to their chemical nature, resulting in five key groups: micelles,
11
3.1.1. Micelles
Polymeric micelles are expected to increase the accumulation of drugs in tumor tissues
by taking advantage of the enhanced permeability retention (EPR) effect [83]. Actually,
micelle systems can also incorporate various kinds of drugs into their inner core with
relatively high stability by chemical conjugation or physical entrapment [84-85]. Also, the
size of micelles can be controlled within the diameter range of 20-100 nm. Therefore, it
is expected that the incidence of drug-induced side effects may be decreased due to
poly(aspartate ester) block copolymers have been physically loaded with CPT, to
improve drug aqueous solubility and stability in biological fluids [87-89]. Here, it was
observed that the stability of CPT-loaded micelles depended on the amount of benzyl
esters and poly(ethylene glycol) (PEG) length in the polymers. After intravenous
administration of CPT-loaded micelles to male ddY mice via lateral veins, it was
concluded that sample using PEG 5000, 27 aspartic acid (Asp) units and 57-75% benzyl
esterification of Asp residue was the most stable formulation, showing an extended
circulation period in blood stream [82]. However, it is not clear if CPT is safely packaged
in these micelles, which is crucial to avoid premature release before reaching tumor cells.
between PBAsp carboxyl groups and CPT 20-hydroxyl [90-91] (Fig. 3). Subsequently,
spherical nanomicelles were produced in aqueous medium. In this case, CPT was
sequestered toward the core whereas PBAsp backbone formed the hydrophilic shell.
Anyway, in vitro cytotoxicity assays of PBAsp-CPT (2,1% CPT) against L929 mouse
muscular cell line showed less toxicity than free CPT (5-10% activity drop depending of
12
O
O
NH
NH
O O
β
HN α HN
O
OH
O
O
O
O
N
N
O
Then, it was studied CPT release in response to cytosol reductive conditions was studied
urothelial cancer cells confirmed that after endosomal permeabilization CPT loaded
micelles escaped (from endocytic vesicles) into the cytosol, promoting drug release of
more than 70% from micelles in the irradiated tissue (fluence rate to induce the
photochemical internalization effect, 100 mW/cm2; time, 1000 s) and better antitumor
3.1.2. Liposomes
based drug delivery [93]. Liposomes protect the encapsulated drugs form structural
technical issues, as the drug shows poor solubility in nearly all pharmaceutically
13
acceptable solvents, as well as it associates with simple phospholipid bilayers. Actually,
and liposomal CPT, which affects to biological activity and biodistribution. In that sense,
a study over several CPT liposome formulations of different nature (cationic, anionic and
In one of the first attempts to deliver CPT within liposomes, the lipid-complexed
formulation (particle size range 20-200 nm) showed significant advantages with regards
to the naked drug for intravenous administration in clinically relevant lipid-drug ratios
(12.5:1 w/w) [95]. CPT-loaded liposomes (LC-CPT) had an in vitro antitumor activity
similar to that of free CPT and displayed similar cytotoxicity against multidrug resistance
(MDR-1) negative human cervix carcinoma KB-V1 cells and MDR-1 positive human
cervix carcinoma KB-3-1 cell line. However, the biodistribution of CPT in ICR Swiss mice
was severely affected by lipid complexation, as CPT achieved the greatest concentration
probably due to the rapid processing of LC-CPT in liver and elimination via the gut.
Furthermore, LC-CPT showed higher antitumor activity than free CPT in in vivo testing
over B6D2F1 mice bearing intraperitoneal tumors of L1210 mouse lymphocytic leukemia
CPT-loaded liposomes with average particle size of about 150 nm were synthesized by
a lipid-film hydration method [96]. In a first step, liposomes were formulated by addition
glycol-modified liposomes and, subsequently, liposome surface was coated with human
serum albumin. Here, CPT was entrapped with about 80% efficiency due to the
protected liposomes showed much higher therapeutic activity than CPT in CDF1 female
14
specimens increased up to 5-fold that of liposome-treated mice), with high accumulation
3.1.3 Dendrimers
Dendrimers have received much attention in drug delivery, as they are highly branched,
multivalent and tunable polymers for different architecture, size, shape and surface
properties [97]. Dendrimers are usually classified by generation (denominated G), which
refers to the number of repeated branching cycles that are performed during its
(PAMAM) and triazine derivatives conjugate easily with the phenolic 20-hydroxyl group
of CPT, resulting in stable nanomedicines. At this point, it has been proved that CPT
solubility increases with PAMAM generation [98]. Here, PAMAM dendrimers of four (G4),
five (G5) and six (G6) generation were prepared and their solubility in aqueous medium
and the interaction with CPT (carboxylate form or lactone form) were evaluated. The
study showed that G6 PAMAN enhanced CPT water solubility in comparison with G4
PAMAN. This can be partly explained by the fact that the number of primary and tertiary
amines in G6 dendrimer increases and could entrap more hydrophobic molecules inside.
promotes a strong reaction with serum proteins, hemolysis and hepatic toxicity.
A relevant PAMAM second generation (G2) model was synthesized with four molecules
of CPT per unit and a trigger that allows activation with penicillin-G-amidase (PGA). In
this system, the cleavage of the phenylacetamide group by PGA triggered the
as cell-growth inhibition assays indicated that half maximal inhibitory concentration (IC50)
value for the dendrimer was between 100- and 1000-fold lower than free CPT against
15
MOLT-3 lymphoblastic leukemia, JURKAT human leukemia T and HEK-293 human
N O
n
N
O
O
O O
O NH
O
N O
O NH
N
O N N
HN
O N O N
O
O N O O
H H N
N O O N
O N O
N O
O N O O O O
O
O
O N
N O
N N O N
O O N
O H O
OO
O
O O
NH N N
O NH O N O
N H O OH
n
N O
O
O N
O N
O
In one important contribution, triazine dendrimer CPT derivatives were synthesized and
their cytotoxic activity evaluated against MCF-7 human breast carcinoma cells and HT-
29 colon carcinoma cell lines [101]. For this sake, CPT was conjugated with isonipecotic
acid through an ester bond, and the obtained secondary amine was highly reactive and
combined with the triazine, allowing to obtain the desired dendrimers, which showed
Click chemistry has been widely used to improve coupling reaction efficacy in the
the other hand, propargylamine (PPA) and methoxypoly(ethylene glycol) amine were
conjugated to PAMAM dendrimer G4.5. Then, CPT-APO was coupled to the modified
16
dendrimer (G4.5-PPA) via click chemistry. Nearly 100% of the CPT molecules were
covalently conjugated to the dendrimer and the drug was released by ester bond
hydrolysis. This new dendrimer-CPT was more cytotoxic than free CPT (IC50 value
In the last years, a myriad of polymer therapeutics, including polymeric drugs, polymer-
105], have received special attention as drug delivery systems and have been
particle size, allowing them to circulate in the bloodstream for long periods of time [106-
107]. Conversely, the main issues of these materials are their fast biological degradation,
which may cause leaching of CPT important quantities of CPT into the blood stream and
interstitial fluids, and their variable loading capability due to limited functionalization.
glutamic acid) and β-cyclodextrins, and natural polymers such as hyaluronic acid,
chitosan and polypeptides, are the most commonly reported for CPT delivery [108-109].
carrier and the bioactive molecule confers enhanced solubility and stability, allowing to
delivery. At this point, one interesting approach that allows exceptionally high drug
loading is to polymerize the drug itself. In one representative example, a dimeric CPT
derivative, CPT-SS-CPT, bearing a trigger responsive domain was used as the core-
constructing unit of the nanoparticle [113]. Here, CPT molecules are stably conjugated
via carbonate linkages that are subject to triggered bond cleavage and subsequent drug
17
release by a reducing reagent (Fig. 5). Upon triggering, cleavage of the disulfide bond in
CPT-SS-CPT would result in decomposition of the drug dimer, releasing CPT in its
authentic form. Unfortunately, CPT-SS-CPT has much lower water solubility (less than
loaded nanoparticles stable in phosphate buffer saline (PBS) and human serum for at
least 3 days were formed, with particle diameter of about 180 nm. The anticancer effect
assay over HeLa human cervix carcinoma cell line. The IC50 of CPT-SS-CPT/mPEG-
PLA was 114 nM, compared to 17 nM for free CPT, but the main advantages of this
nanoconjugate are the lack of systemic toxicity, as CPT release only takes place through
disulfide bond reduction by intracellular thiols, and the stabilization of the lactone ring.
O O N3
S
S O -
O S
O O O
O O O O O
O O O NH O NH O
O O O
O O O O
O O O O O
O N O N
N N O
N N
N O
N S
CPT-SS-CPT
O O O
O O H 2N O
O
O O O O
O N
N
N
N
O
N
O N O + CO2
O
OH CO2 O O O
O O NH2 O OH
O NH2 O O
O O O
N HO O O N
N
N H2O
N
N
18
One of the best performances to improve CPT stability and solubility in biological
weight (over 50 kDa) β-cyclodextrin-CPT based polymers [114-117]. For this purpose, a
CPT-glycine ester derivative was prepared, which was subsequently linked to the β-CD
derivative by amide bond (Fig. 6). This way, the solubility of CPT is increased by more
than three orders of magnitude [114]. Moreover, this nanomedicine showed higher
concentration in tumor and superior efficacy than irinotecan in the treatment of CD-1
female athymic nude mice bearing LS174T human colon carcinoma, HT-29 human
colorectal carcinoma, H1299 human non-small-cell lung cancer model or Panc-1 human
product (CRLX-101, formerly IT-101) has been released, currently under clinical trials.
β-cyclodextrine has also been the aim of further development to improve CPT solubility
the inclusion of CPT and lactone ring stability. In vitro cytotoxicity assays in L929 mouse
19
muscular cell line suggested a decrease of cytotoxicity at higher dose of CPT@β-CD-
graft-PAsp [120].
refractory (DU-145 and PC3 human prostate cancer cell lines) and sensitive (LnCaP
human prostate cancer cells) cell lines. CPT was incorporated in the CDN structure and
its interaction with the nanosponge was confirmed by thermal analysis. Cell proliferation
and clonogenic assays showed a greater responsiveness of CDN-CPT against PC3 and
DU145 cell lines, blocking colony formation and cell growth starting at 1 nM dose.
Besides, in vitro assays against LNCaP cell line, demonstrated that CDN-CPT inhibited
Poly(L-glutamic acid) (PGA) polymers have been combined with CPT by direct
esterification [122], or by using a short linker, like glycine [123], obtaining loadings of
nearly 40%. The administration of four doses of 7% CPT-loaded PGA by iv injection (one
single dose every four days, with an equivalent CPT dose of 40 mg/kg) over Sprague-
Dawley female athymic nude mice bearing NCI-H322 human lung carcinoma tumors
showed a 32-day tumor growth delay with regards to untreated mice, prolonging the
molecular weight up to 50 kDa and CPT loading up to 37%, it was reported a tumor
growth was delayed to less than half in comparison to untreated mice after a 30 mg/kg
CPT equivalent dose over athymic NCr (nu/nu) mice bearing NCI-H460 human lung large
In the last decade, among different biopolymers, the use of chitosan for drug delivery
20
chemical conjugation of hydrophobic 5b-cholanic acid moieties to the hydrophilic glycol
chitosan backbone [125]. CPT was easily encapsulated into HGC nanoparticles (CPT-
HGC, 280-330 nm in diameter), which showed sustained release for one week. CPT-
HGC exhibited significant antitumor effects and high tumor targeting ability towards
More recently, CPT has been conjugated with hyaluronic acid (HA-CPT) with variable
molecular weight, in order to overcome the low solubility and poor targeting ability of the
naked drug [32]. Here, CPT was incorporated using two linkage molecules such as
succinate and adipic acid dihydrazide. The authors studied the influence of the polymer
molecular weight over solubility, drug loading and stability. Solubility assays confirmed
that these conjugates increase CPT solubility, mostly those samples with lower
molecular weight. Furthermore, in vitro efficacy assays demonstrated that the antitumor
activity of HA-CPT polymers were higher than the naked drug against HepG2 human
A singular and very efficient approach for CPT delivery has been achieved by
through a solvent evaporation technique [126]. In vitro efficacy assays against BT-474
human breast cancer cells, MCF-7 human breast adenocarcinoma cell line and SKBR-3
human breast cancer cell line, demonstrated that the intracellular release of the cargo
induced cell death with half maximal inhibitory concentration (IC50) values similar to those
observed with the naked drug. Furthermore, in vivo assays of the PA-encapsulated CPT
21
3.2. Inorganic nanocarriers
They usually show high stability in the biological environment, ease of functionalization
and electronic) that can be tailored by controlling particle structure, composition, size
plasmonic nanoparticles, among others, have been proposed for imaging and therapy
most of these systems suffer of biocompatibility issues, mostly related to the toxicity of
some structural components and the strong interaction with serum proteins, which results
in rapid blood clearance and hepatic removal [129]. Moreover, occasionally, the
elimination from organic tissues may be an issue, involving the accumulation in the body
nanoparticles that have found application for delivery of CPT, metal oxides and carbon
nanotubes.
The most significant use of transition metal oxides for biological purposes are magnetic
10-20 nm, which is dependent on the material, have a large magnetic moment and are
Iron oxide (Fe3O4) hollow monodispersed spheres of about 200 nm mean diameter, and
synthesized using a one-pot solvothermal method, and further loaded with CPT by
leakage before cell internalization, and exhibited slightly higher cytotoxicity than free CPT
against 786-O, HepG2 and HeLa cell lines, due to the enhanced solubility of the
22
transported drug. Furthermore, non-loaded Fe3O4 hollow spheres showed nearly no
cytotoxicity towards the cells, although there is no report on the stability of these
Very high CPT loading has been achieved into iron oxide (Fe3O4) superparamagnetic
CPT were produced by incorporation of the drug to a SPION aqueous solution; this
increased up to 26 wt% CPT by functionalizing SPION surface with a PEG. The high
loading capacity was attributed to the amphiphilic nature of the PEG molecule.
Nevertheless, this did not bring about any improvements in cytotoxic activity, as CPT-
loaded nanocarriers demonstrated about 10-15% lower apoptotic levels over H460 cell
culture than the free drug. This was justified on the basis of CPT slow release profile
Carbon nanotubes (CNTs) have recently emerged as a new alternative vehicle for
transporting and releasing therapeutic molecules, due to their high surface areas
toxicity (mostly by liver and kidney accumulation), must be solved before they can go
into clinical application. In this sense, surface chemical modification becomes a crucial
step for preparation of soluble CNTs and efficient drug incorporation [127].
(PVA) allowed CPT loading via π-π interactions (Fig. 7) [133]. This construct was found
to be very stable in simulated body fluid and, in in vitro testing, it was approximately 15-
fold more cytotoxic than free CPT against MDA-MB231 human breast cancer and
Fig. 7. (A) Schematic depiction of CPT loaded MWCNT-PVA. The figure shows CPT
molecule in red, MWCNT structure in black and polymer PVA in blue. (B) UV spectra of
multiwall CNT-PVA and CNT-PVA-CPT samples with different concentrations.
(Reprinted from ref. 133 with permission of The Royal Society of Chemistry)
Another singular CNT-based material suitable for CPT delivery are glyconanosomes
determine their shape and topology (Fig. 8) [134]. Experiments against MCF-7 human
antitumor activity than the naked drug (killed more than 65% of MCF-7 human breast
cancer cells while CPT alone affected less than 55% MCF-7 cells), which could be
explained because the interaction of GNS/CPT could interfere with cellular CPT
detoxification mechanism in such a way that intracellular CPT retention time increases,
24
OH O
OH O
HO O S NH O
O
O HO HN
OH 4N 7
OH H
OH O
10 HN
HN HN HN
O 7 O
O O HN
7 7 7 7 O
7
Hydrophillic Hydrophobic 10
head tail
10
9 9 9 10
O
9
N
8 H
O
9
N
8 H
O
9
N
8 H
As commented before, fully inorganic systems found strong limitations for biological use,
due to possible compositional toxicity and severe interaction with serum proteins. For
this reason, inorganic nanocarriers are better handled surface modified or combined with
organic components, which provide good stability in physiological fluids, low toxicity and
immunogenicity, and minimized reactions with serum proteins, to control their fate in
25
distributed in the nanoscale or molecular range are hybrid materials. Typically, the hybrid
nanocontainers are made of inorganic and organic building blocks, presenting unique
properties depending on the spatial and size distribution of their components [135-136].
Actually, efforts have been done to design hybrid devices combining different functions
the weaknesses of one another [137]. Whilst organic component provides good
biocompatibility and better solubility in physiological fluids, the inorganic moiety offers
almost unlimited external surface area to bind different functionalities, improving the
capabilities of the DDS. On the other hand, the main handicap of these systems is the
lack of long-term toxicity studies about the possible interactions within the body of the
different components.
A myriad of compositions and models have been reported with potential use in
section we only present a brief description of some of the most representative examples
of the different systems proposed, as most of these hybrid nanocarriers for CPT delivery
triggers for selective release activation at the targeting tissues and cells. These last are
attention in the last decade. In the case of biomedical applications, their complex
chemical composition, their variable structure, but mostly the inherent multifunctional
property, offer plenty of room for the development of novel systems that are able to carry
26
Mesoporous silica nanoparticles (MSNs) modified with organic ligands are probably the
most widely used inorganic material for hybrid DDS development [138]. These particles
have large surface areas and porous interiors that can be used as reservoirs for storing
hydrophobic drugs as CPT. Moreover, particle surface and pore size can be easily
tailored to selectively store and deliver different molecules. In this context, MSNs were
which reduced aggregation [139]. Then CPT was absorbed within the 2 nm diameter
pores and the particles were incubated with different cells lines (PANC-1 human
pancreatic carcinoma cell line, Capan-1 human pancreatic ductal adenocarcinoma cell
adenocarcinoma cell line and MKN 45 human gastric cancer cell line) to determine if they
were able to transport the hydrophobic drug into the cancer cell. In all cases, results
indicated that the cytotoxic efficacy of CPT-loaded MSNs was very similar to that of the
dissolved CPT. In an ulterior study, folic acid was incorporated to the surface of MSNs
BALB/cAnNCrj-nu nude mice bearing MCF-7 human breast carcinoma tumors. These
authors showed that nanoparticles were able to repress subcutaneous tumor growth with
a minimum therapeutic dose of about 0.5 mg/kg CPT-loaded FA-MSN for complete tumor
growth inhibition. Inhibition of tumor growth was also complete in the CPT-treated group
but with clear signs of toxicity in blood analysis, as high concentration of liver enzymes,
levels of calcium and phosphorous. Moreover, one mouse in the group treated with CPT
was euthanized on the 51st day due to manifestation of severe toxicity. Conversely, the
group treated with CPT-loaded FA-MSN only presented mildly elevated blood
27
In another MSN-based model, MSNs were functionalized with hyaluronic acid (MSN-HA)
capable of selectively targeting specific cancer cells over-expressing the CD44 protein,
leading to rapid and concentration dependent uptake by the cancer cells through the
in the pores, and the system showed enhanced cytotoxicity to HeLa cells compared to
both free CPT and CPT-loaded MSNs-HA in the presence of excess free HA.
The integration of functional organic fragments into the framework of mesoporous silica
hybrid compositions where the organic R moiety homogeneously distributes within the
whole framework without blocking pore channels. The recent development shift from bulk
strategy the chemical nature of the silica framework can be significantly altered, offering
a number of specific biological effects and functions suitable for biomedicine, as tunable
vivo evaluation of the biological effects and the biosafety of MONs is still in its infancy,
and complete studies of biodistribution, excretion, biodegradation, and toxicity are still
pending [147]. In this context, in a singular study, monodisperse MONs co-doped with
anthracene (Anth), naphthalimide (Naph) and biphenyl (Biph) at various ratios were
prepared by a modified Stöber method, and CPT was loaded inside the pores [148].
Next, release experiments in PBS demonstrated that sample containing the three
fluorophores (Biph/Anth/Naph) presented quicker release profile than samples with two
fluorophores (Biph/Anth) or with only Biph. Actually, this CPT tunable retention ability is
28
governed by the quantity of incorporated fluorophores and their polarities, which follows
the order of Naph > Anth ~ Biph. Moreover, although the release rate can be more finely
tuned in an asynchronous way (a single drug) by varying the ratio between fluorophores,
it also possible monitoring the synchronous discharge of several drugs. In this sense, it
can be said that drug releasing performance of MONs can be well correlated with its
color. Studies on the cellular uptake efficiency were carried out on HeLa human cervix
carcinoma cell line, resulting that by varying the ratio between the different fluorophores,
a colorful bioimaging system benefited from FRET should be obtained. Also, the cell
viability evaluated using MTT assay indicated similar growth inhibition potential for MONs
Fig. 9. Illustrations (A, D, and G), emission spectra (B, E and H) and TEM images (C, F
and I) of monodisperse mesoporous organosilica nanoparticles co-doped with FRET
cascades composed of triple fluorophores: anthracene (Anth), naphtalimide (Naph) and
biphenyl (Biph). The red plus sings indicate the maximum absorption wavelength of
different fluorophores. (Reprinted from ref. 148. Copyright © 2012 American Chemical
Society)
Closely related to MSNs, mesoporous bioactive glass nanoparticles, with higher specific
surface area and pore volume than conventional bioactive glasses, have been recently
29
proposed for hybrid DDS development [149]. In a significant study, the mesoporous
within the pores (MBG-CPT) [57]. In vitro testing over HeLa human cervix carcinoma
cells, which are overexpressed folate receptors, and L929 murine fibroblast cells, which
are deficient folate receptors, suggested that while loading efficiency was better for FA
free nanoparticles, MBG-CPT showed higher cytotoxicity in HeLa cell line than MBG-OH
because of the increased cell uptake of anticancer drug delivery vehicles mediated by
the cells by endocytosis and that cell uptake was mediated by the folate receptor.
solution led to drug loaded nanovesicles with stable polygonal geometry, consisting of
encapsulation and release profile for CPT, which are strongly conditioned by chemical
In another organic-inorganic model CPT was first incorporated into micelles derived from
aluminum layered double hydroxides (LDHs) by an ion exchange process [151]. The
30
When administered to 9L Glioma cells, the nanobiohybrid containing CPT resulted in
significantly lower survival times compared to untreated cells, or to cells incubated with
the surfactant, the pristine LDH, or water (delivery medium). In addition, LDH surface
targeting biomolecules, which may direct these nanohybrids towards specific cells or
nanocomplexes showed slightly lower cytotoxic activity than free CPT, they look like a
promising biocompatible model for CPT delivery, allowing drug administration in a dose-
provide high resolution imaging and sensing (e.g., using fluorescent molecules,
and controlled drug delivery and release, combining organic and supramolecular
chemistry with colloidal nanoscience. Here, different models for delivery and controlled
release of CPT have been reported, the most significant of which are described below.
Core-shell organic nanocapsules for oral delivery of camptothecin were formulated with
a core made of amphiphilic cyclodextrins (CDs) and an external coating of chitosan (CS)
[153]. These formulations were in the range of 180-220 nm, with narrow size distribution.
CPT was loaded in the inner polymer and then positively charged chitosan was added
to cover the particles. The resulting nanocapsules were found to be stable in simulated
31
gastrointestinal media and proved favoring CPT permeability through an artificial mucus
gel layer and through Caco-2 (human colorectal adenocarcinoma cells) membrane. In
vivo animal studies in CD1 female mice demonstrated that CPT-loaded CD@CS
nanocapsules promoted drug absorption at the intestinal tract vs stomach uptake. This
could be an interesting vehicle for CPT oral administration, although it does not solve the
One of the most studied core-shell systems is that obtained by protecting a metal
sulfated polysaccharide outer arm) as shell, were synthesized for CPT targeted delivery
(Fig. 10) [154]. These nanocarriers presented a release pattern curve at different pH in
two stages, and an initial rapid release (1 h) was followed by a sustained release period
Fig. 10. Schematic illustration of gold nanoclusters as core and folate (FA)-conjugated
amphiphilic hyperbranched block copolymer as shell. (Reprinted with permission from
ref. 154. Copyright © 2012 American Chemical Society)
32
(up to 15 h), and then reached a plateau. Cytotoxicity studies over HeLa cells showed
that, with regards to the free drug, CPT-loaded nanocarrier provided slightly higher
(about 10%) anticancer activity, also gaining specificity to cancer cells with
Fully inorganic core-shell hybrids can also be used for CPT delivery. In a most usual
outer layer, CPT being encapsulated inside the silica pores. In this sense, recently, our
group carried out combined chemo and photothermal therapy in in vitro testing by means
protecting shell of porous silica that contains CPT (GNC@SiO2) [155]. This
surface plasmonresonance peak, LSPR, at 660 nm) to the cytotoxic activity of CPT
simultaneously released for the efficient destruction of cancer cells. A method was used
for the controlled assembly of 15 nm diameter gold colloids into stable clusters with a
silica coating that incorporates CPT molecules within the pores. After internalization in
photothermolysis under femtosecond pulse laser irradiation of 790 nm. Moreover, the
simultaneous release of CPT during the process increased cell death up to 70%. On the
other hand, control experiments with no irradiation reported 45% cell death for CPT-
loaded GNC@SiO2 nanoparticles and 30% cell death for the free drug. This therapeutic
model could be of interest for application in the treatment and suppression of non-solid
applications in cancer therapy due to their optical properties (LSPR at 770 nm) and
cytotoxic activity (CPT loading in the mesopores). Moreover, tLyP-1, a kind of tumor
33
homing and penetrating peptide, was engrafted to GNR@SiO2 [156]. The obtained
GNR@SiO2−tLyP-1 was loaded with CPT and tested over hMSC human mesenchymal
stem cells, HeLa human cervix carcinoma and MCF-7 human breast cancer cell lines.
After cell internalization, core-shell nanorods were irradiated by NIR illumination. Then,
GNR@SiO2−tLyP-1 surface plasma resonance and CPT cytotoxic activity. Moreover, the
systematic toxicity of CPT on human mesenchymal stem cells was minimized, because
the GNR@SiO2−tLyP-1 selectively targeted and penetrated into the tumor cells, and little
Fig. 11. (A) Schematic of synthesis and controlled aggregation of gold colloids protected
with a thin silica layer into gold nanoclusters (GNC@SiO2). (B) UV-vis/NIR absorbance
spectra of gold nanoclusters obtained at different NaOH concentration. (C) Photograph
of the different colloidal gold dispersions. (D) Fluorescent microscope images of 42-MG-
BA human glioma cells incubated with GNC@SiO2 (with and without CPT) after laser
irradiation. A control with no nanoparticles (NP) is also shown. (Adapted from ref. 155
with permission of The Royal Society of Chemistry)
34
3.2.3. Metal organic framework nanoparticles
Due to their porous ordered structure and virtually infinite combinations of metals and
ligands, metal-organic frameworks (MOFs) can exhibit exceptionally high surface areas
with large pore sizes. Consequently, they have been recently proposed for applications
in loading and controlled release of several drug molecules [157]. However, MOFs need
(NMOFs) to be used as delivery vehicles for imaging agents and drug molecules [158].
NMOFs possess some potential advantages over existing nanocarriers given their
compositional and structural diversity, which allows for the synthesis of different shapes,
sizes and chemical properties, and their biodegradability, a result of relatively labile
metal-ligand bonds. Surprisingly, despite the huge potential of these nanomaterials for
of potential hazards associated to their components are still a matter to investigate [159].
Very few toxicological studies concern NMOFs, usually limited to specific cell lines, which
makes it difficult to compare the results obtained. In this context, there are very few
examples of MOF application to CPT delivery yet, most of them focused on the structure
of ZIF-8 [160,161].
cell line demonstrated cell internalization and minimal cytotoxicity. The 70 nm particle
size facilitates cellular uptake, whereas ZIF-8 framework easy dissociation results in
quick endosomal release of the small-molecule cargo. Here, it was shown that
encapsulated CPT ZIF-8 particles show enhanced cell death, which involves cell uptake
35
Fig. 12. Schematic illustration of MOF application for small molecule delivery (fluorescein
in green (A) and CPT in blue (B)) inside of ZIF-8. CPT molecule fits well within the
microporous framework, which loads a maximum of 2 % drug. (Reprinted from ref. 160.
Copyright © 2014 American Chemical Society)
efficacy. All vehicles provide a clear increase of drug stability, as the encapsulated or
bonded molecule is protected from enzymatic degradation and lactone ring cleavage
during its transit through blood stream and tissues to tumor location. Moreover, colloidal
stability of these formulations can boost drug solubility several orders (e.g., β-
cyclodextrins), which gathered to the stability rise improves the bioavailability. However,
the most significant effect of incorporating CPT into a nanocontainer is the dramatic
toxicity reduction and therapeutic window extension, as the transported molecule has
very little activity before it is discharged at the tumor place. As a consequence, the in
vivo antitumor activity increases with regards to the free drug, due to an improved
bioavailability and the possibility to increase the dose with no side effects. In addition,
many of these nanomedicines offer the possibility of delivering several drugs in a row,
which is very important for the treatment of MDR cancer, and allow for the incorporation
of targeting molecules that favor selective therapy. Finally, inorganic and hybrid materials
can present unique properties (e.g., optical activity, magnetism, semiconductor activity),
36
4. Stimuli responsive systems
These are currently the most representative DDSs in Nanomedicine. The main
characteristic is their ability to deliver a therapeutic agent into the target cells with no
previous release. This is due to the accurate control of the release process through a
the stimulus which activates the drug release mechanism. Examples of endogenous
stimuli are pH, redox potential and enzyme activity. On the other hand, classical
exogenous stimuli include magnetic fields, ultrasounds, light and temperature [162-165].
Several pH-sensitive DDSs have been developed based on the pH difference between
normal and tumor tissue. Here, some of these platforms take advantage of the fact that,
in some cases, tumor tissue pH profile is slightly lower than normal tissue. However, the
that do not distribute readily within the tumour interstitium. This factor, together with the
lower pH values inside endosomes, suggests that enhancing endosomal escape is the
most realistic objective for pH-sensitive nanocarriers in the short term [166].
method. For this purpose, hydrophilic methyl ether poly(ethylene glycol) (MPEG) and
copolymer (Fig. 13). Then, CPT was encapsulated into the micelles (CPT/MPEG-PAE)
PAE were demicellized and the encapsulated CPT was rapidly released from these
nanocarriers, although under pH 7.4 the release rate was clearly slower. In vitro
cytotoxicity assays in MDA-MB231 human breast tumor cells confirmed that these
37
micelles provoked severe cytotoxicity within an acidic environment due to rapid CPT
Fig. 13. (A) Chemical structure of methyl ether poly (ethylene glycol) - poly(β-aminoester)
block copolymer (MPEG-PAE). Letters x and y indicate the number of times that the
block polymer is repeated. (B) Schematic diagram of pH-responsive CPT/MPEG-PAE
micelles at weakly pH condition. The figure shows CPT in blue, and in red and yellow
the MPEG-PAE block copolymer. (C) Size distribution of MPEG-PAE micelles in PBS at
pH 7.4 by dynamic light scattering (inset: TEM image). (Reprinted from ref. 167,
Copyright © 2010, with permission from Elsevier)
Another pH-sensitive nanoscaled system for CPT delivery has been obtained by
glycol (MAP). When this polymer-drug conjugate was placed in water, it self-assembled
release rate with pH was observed but, unlike the standard behavior, as pH increased
from 6.5 to 7.4, the release half-live dropped from 338 to 58 h for MAP−CPT
nanoparticles. In vitro testing over BT-474 and HER2 human breast cancer cell lines
showed that cellular uptake of nanoparticles was enhanced by 70%, compared to non-
and nanoparticle disruption by fat. The targeted MAP-CPT nanoparticle system carried
38
ca. 60 CPT molecules per nanoparticle and showed prolonged plasma circulation with
an elimination half-life of 21.2 h and AUC value of 2766 μg h/mL at a 10 mg CPT/kg tail
core of poly(p-dioxanone) has been also used to control drug release on the basis of
environment pH changes [169]. CPT was incorporated in the core of this system and its
release was triggered at pH 7.4, 6.2 and 5.0. In vitro drug release studies in HeLa human
cervix carcinoma and L929 mouse fibroblast cells demonstrated that the micelles
presented a much faster release of CPT at pH 5.0 than at pH 7.4, whereas blank micelles
without CPT were found to be nontoxic in both cell lines. The good internalization effect
This nanoplatform was based on hollow silica nanoparticles sealed with ZnO quantum
dots (QDs). Subsequently, CPT was encapsulated in the hollow core through
hydrophobic interactions. On the other hand, doxorubicin (DOX) was deposited on the
the mesoporous surface [170]. This cooperative drug loading largely increased drug
encapsulation efficacy of both CPT and DOX up to 89.28% and 44.98%, respectively.
Drug release was compared at pH 7.4 and pH 5.0, concluding that in acid medium the
space and reducing intracellular space, which is related with glutathione (GSH)
39
concentration. This fact has found application in the building of specific DDSs able to
selectively discharge their therapeutic load after cell uptake. Here, drug covalent linking
to nanocarriers is mostly carried out by disulfide bridge, which makes possible to design
of different sensitive systems. The main concern for in vivo application of these systems
is the presence of GSH in serum which, despite the fact that its concentration is less than
one order below that of cytosol, may provoke some premature release during CPT transit
linking CPT-loaded organic units with disulfide bonds [172]. These micelles were
core via click chemistry (Fig. 14). In this case, CPT was rapidly released due to the
disulfide bond cleavage, which is present both in the CPT prodrug and the cross-linker.
In vitro cytotoxic assays against MCF-7 human breast cancer cells suggested high
anticancer activity due to the disassembly of the cross-linked micelle and rapid CPT
release.
Also recently, our group implemented a new synthetic strategy by direct coupling of as-
of silica hybrid nanoparticles containing a non-porous core and a mesoporous shell [173].
Upon reaction with thiols in physiological conditions, disulfide bridge cleavage occurs,
releasing the naked drug after an intramolecular cyclization mechanism (Fig. 15).
subcellular level for the monitoring of uptake and delivery. Confocal microscopy
experiments in HeLa human cervix carcinoma cells confirmed that nanoparticles enter
40
Fig. 14. Schematic procedure for preparation of redox-responsive uncross-linked (UCL)
and core-cross-linked (CCL) hybrid micelles. (Reprinted partially with permission from
ref. 172. Copyright © 2013 American Chemical Society)
the cells by endocytosis but are able to escape from endo-lysosomes and enter the
in a separated study [174], it was shown that prodrug side chain carbon number (n)
aqueous medium and cell uptake kinetics. When n value increases, the negative surface
Furthermore, n value also determines the type of products released and, subsequently,
the cytotoxic activity. Although full disulfide bridge reduction occurs in all cases within
the cell, the subsequent intramolecular cyclization releasing CPT, CO2 and the
corresponding thiolactone only happened quantitatively for n=1 (Fig. 15), whereas higher
homologues released low CPT quantities (5% for n=2 and traces for n=3).
41
mesoporous silica
amorphous silica core
shell with CPT
with Rhodamine B
O O
Si S O O O
S
n
O N
n=1,2,3
HS N
GSH
Si
O O
O O O O
HS Intramolecular
n cyclization N
O N N
O CPT
O
N O n=1: 100 %
O HO
n=2: ~5 %
S
n=3: <1 %
n
Fig. 15. Redox-responsive nanoplatform for CPT delivery by direct coupling of (pyridin-
2-yldisulfanyl)alkyl carbonate CPT derivatives with thiol groups of silica hybrid
nanoparticles containing a non-porous core and a mesoporous shell. Upon reaction with
thiols in physiological conditions, disulfide bridge cleavage occurs, releasing the naked
drug after an intramolecular cyclization mechanism. (Adapted from ref. 173, with
permission of The Royal Society of Chemistry)
au protein through the reducible disulfylbutyrate (buSS) linker, obtaining discrete, stable,
well-defined nanostructures with a high and quantitative drug loading [175]. Drug content
was precisely controlled using the two amine functionalities of the amino acid lysine to
create branching points that allow the attachment of one, two or four CPT molecules,
corresponding to respective drug loadings of 23%, 31% and 38%. Depending on the
can be either nanofibers or nanotubes. Such nanostructures provide CPT protection from
the external environment. Release profiles showed that glutathione degraded the
designed buSS linker, releasing a mixture of intermediates and CPT. Under tumor-
relevant conditions, these drug nanostructures released the active form of CPT and
showed in vitro efficacy similar to free CPT against MCF-7 human breast cancer cells
and 9L and F98L rat gliosarcoma cell lines. In addition, to avoid the formation of
42
intermediates, the same researchers prepared these CPT-nanostructures using a
followed a self-immolative process to release free CPT, improving cytotoxicity over MCF-
7 cell line [176]. Furthermore, in vivo preliminary experiments, after intratumoral injection
with 36% CPT loaded sample over 6-8 week female BALB/c mice bearing subcutaneous
CPT nanotubes in the tumor site than the free molecule, which allows a sustainable
release over a long period [177]. Conversely, no improvement in tumor targeting was
detected with the nanomedicine in case of administration by tail vein, so the potential of
release via enzymatic digestion of the nanocarrier or linking moiety [178]. The biological
recognition of the substrate by the enzyme leads to the selective and specific triggering
event to release the cargo, which minimizes toxicological issues related to CPT. In this
context, due to the high concentration of carboxylesterases in the cytosol, most of these
models have been developed over the basis of an ester bond at 20-OH position between
CPT and the nanostructured surface. On the other hand, there are two main
shortcomings in the use of enzyme-responsive devices for CPT delivery: i) the presence
tumor cells, and ii) the binding of CPT to a nanoparticle surface can limit the accessibility
to enzyme active site due to steric effect, slowing down the release process, which
43
In a singular approach, CPT molecules were directly grafted to carboxyl groups of poly(L-
derivatives (Fig. 16). This system released CPT under physiological conditions by means
cells indicated nanoparticle delivery system IC50 was 4 times that of the free drug,
probably due to the slow release of CPT from polymer. However, a positron emission
64
tomography study with Cu-labeled nanoparticles in C57Bl/6 mice bearing Lewis lung
than the free drug, delaying tumor growth with no visible body weight or any other side
effect.
steps. First, a prodrug is prepared by esterification with a short linker and then, this
44
prodrug is covalently linked to a supramolecular structure. For instance, to incorporate
CPT into PAMAM, CPT was firstly substituted at 20-OH position with glycine through
ester bond and, afterwards, the prodrug was reacted with PAMAM G4, to give PAMAM-
CPT [181]. This material was stable in plasma and the presence of proteolytic enzymes
such as cytosolic esterases promoted conjugate cleavage and CPT release. Cell
cancer cells indicated a high cytotoxic effect due to nuclear fragmentation and formation
of apoptotic bodies.
nanoplatform for CPT delivery based on enzyme-triggered release. For this sake, CPT
was covalently linked to nanoparticles through an ester bond with the 20-hydroxy moiety,
in order to stabilize its lactone ring and to avoid unspecific release of the drug [179]. The
obtained material was highly stable in plasma, with less than 5% release of the cargo at
physiological pH. Cell internalization and in vitro efficacy assays against HeLa, U87-MG
human primary glioblastoma cell line, HCT-116 human colon carcinoma and HT-29
(SNP-CPT) entered cells via endocytosis and the intracellular release of the cargo
induced cell death with half maximal inhibitory concentration (IC50) values and cell cycle
distribution profiles similar to those observed for the naked drug. Furthermore, in vivo
45
a
A 1200
PBS Tumor Volume
CPT
1000 SNP-CPT
600
400
200
0
1 3 5 7 9 11 13 15 17 19 21 23
Time of Treatment (days)
Vehicle
3.0
2.0
CPT 1.5
1.0
0.5
SNP-CPT
Fig. 17. Tumor growth delay in hsd:athymic nude-Foxn1 mice bearing HT-29.Fluc
subcutaneous (s.c.) tumors. (A) Comparative analysis of the localized s.c. growth of the
HT-29.Fluc tumors untreated (PBS) and treated (CPT or SNP-CPT) in hsd:athymic nude-
Foxn1 mice by external measurement (caliper) of tumor volume. (B) In vivo monitoring
of a representative mouse from each treatment group shows radical tumor growth
reduction with regards CPT and the control group. (Reprinted partially from ref. 179,
Copyright © 2011, with permission from Elsevier)
by light for biomedical applications has become an important field of research [182]. This
46
(UV), visible (Vis) and/or near infrared (NIR) light which is able to activate drug release
drug release. In this sense, light is not only noninvasive but also controllable both
spatially and temporally, allowing for greater safety and specificity [183]. Conversely, low
penetration ability and predominant use of UV and visible light limit the clinical potential
of these systems. So far, many different systems have been described, involving
developed to deliver and release CPT into living cells upon light activation [185]. These
light-triggered MSNs were functionalized with azobenzene moieties in the pore interiors,
with one end attached to the pore walls and the other end free to undergo
photoisomerization. The impeller in the nanopores trapped drug molecules in the dark,
and release them in response to UV light (413 nm) excitation (Fig. 18). In vitro studies in
PANC-1 human pancreatic and SW480 human colon cancer cell lines showed the
dark, the cells were irradiated with ~0.1 W cm-2, 413 nm light, for various excitation times
(0 to 10 min). Cell death was induced under photocontrol. In the absence of light
excitation, the CPT remained in the particles and cells were not damaged. However,
illumination promptly expelled CPT from the particles, causing cancer cell apoptosis,
proof-of-concept, because irradiation at this wavelength can damage cells and does not
penetrate deep inside the tissues. However, using two-photon excitation (TPE) in the
NIR region instead of UV/Vis light leads to better tissue penetration, lower scattering
47
Fig. 18. Schematic illustration of the light-activated mesostructured silica nanoparticles
functionalized with azobenzene derivatives in the pore interiors. (Adapted with
permission from ref. 185. Copyright © 2008 John Wiley and Sons)
losses and three-dimensional spatial resolution. For this sake, the former nanoimpeller-
controlled MSNs device was functionalized with a high absorption cross-section two-
photoizomerize azobenzene moieties in the NIR region [184]. The two-photon absorption
properties of the fluorophore were retained in the materials and no decrease of cross-
sections σ2 was noticed after encapsulation. The nanoimpeller groups pending in the
porous framework allow the physical entrapment of CPT, which is then kicked out of the
loaded with CPT and then screened under TPE in MCF-7 human breast cancer cells.
Cells were incubated with nanoparticles and further irradiated in at very short time with
a focused laser beam (input 3 W, output 900 mW cm-2). The MTT assay was performed
two days after irradiation. Nanoimpellers loaded with CPT and high
comparison to no irradiation.
payload molecules [186]. To tackle this strategy, three CPT molecules were attached to
48
a phenol group connected to a photolabile 2-nitrobenzyl ether moiety. This photolabile
group was linked to a DNA strand to obtain an amphiphile, which self-assembled in DNA-
CPT nanostructure. Under UV irradiation (365 nm), the 2-nitrobenzyl was cleaved,
releasing the DNA from the conjugate and leaving a decapped self-immolative drug core,
to release free CPT molecules. In vitro efficacy assays in SK-BR-3 human breast cancer
light triggering, achieving IC50 values similar to those observed with the naked drug.
hydrophobic layer able to monitor drug delivery by adjusting the wetting of particle
surface. In this context, a light-sensitive release system based on a hybrid with MSN
core and spiropyran and perfluorodecyltriethoxysilane coating was built [187]. Particles
are protected from being wetted by water at the optimal ratio of spiropyran to fluorinated
silane (0.249:1), which successfully inhibits drug release. Upon irradiation with UV light
(365 nm), the conformational conversion of spiropyran from a “closed” state to an “open”
49
state caused the surface to be wetted, leading to the release of encapsulated CPT from
the pores (Fig. 19). In vitro experiments in EA.hy926 human somatic cell hybrid and HeLa
human cervix carcinoma cells demonstrated an accurate drug controlled release. Here,
CPT loaded nanoparticles (10 mg/g) were incubated in cells for 24 h and then, upon UV
light irradiation a noticeable loss of cell viability was observed with regards to non-
materials which are embedded into organic polymeric moieties, allowing the release of
the drug under an external magnetic stimulus. The inorganic component is designed to
impart responsivity to external magnetic fields, both static and dynamic, in order to
impose a remote magnetic control of drug cargo release. Furthermore, soft matter
principles [188]. Nevertheless, as commented before (section 3.2), the most serious
concern about the use of magnetic nanoparticles within therapeutic formulations is their
biocompatibility, especially in the long term. In particular, their stability when introduced
within the body is a crucial aspect, as the proper dispersion and circulation of
nanoparticles strongly affect drug release kinetic, accumulation, and toxicity. Actually,
SPIONs are the only engineered inorganic nanoparticles approved for human use.
One of the most illustrative drug delivery models of hybrid magnetic materials is based
on MSN pore capping with iron oxide nanoparticles [189]. This may be preferentially
done through chemical bonding. For this sake, MSNs surface was funtionalized with 3-
50
aminopropyltrimethoxy silane groups. Afterwards, CPT was filled into the pores and
these were covalently capped through amidation of the 3-aminopropyl groups at surface
nanoparticles [190]. The Fe3O4 nano-caps formed a dense and uniform layer tightly
bound to the MSN surface (MSN@Fe3O4). Under magnetic stimulus, the cumulative drug
release was very quick, and prolonged even when the stimulus was removed, due to
removal of the nano-caps implies a cleavage of the chemical bond between the Fe3O4
nanoparticles and silanized MSNs. The energy induced from the magnetic stimulus over
effective removal of the nano-caps. Actually, the rate of nano-cap removal and drug
diffusion are a function of the magnetic energy. Then, cytotoxicity studies in A549 human
lung adenocarcinoma epithelial cell line by MTT assay, indicated that nano-Fe3O4-
Fig 20. Schematic illustration of the synthesis and structure of mesoporous silica
nanoparticles capped with iron oxide nanoparticles. The structure was obtained by
capping mesoporous silica nanoparticles (MSN) with monodispersed Fe3O4
nanoparticles through chemical bonding. Under magnetic stimulus, some of the Fe3O4
nanoparticles were removed from the surface of the MSN, and CPT was released. On
the right, a brief illustration of all different ligands used in the synthesis of this material.
(Reprinted from ref. 190 with permission of The Royal Society of Chemistry)
51
capped MSN effectively constrained the CPT from free elution and were well tolerated,
resulting in a cell viability of more than 80%. However, upon magnetic stimulus, about
42% of the cells were killed, and the major cause of cell death was the CPT released
Magnetic stimulus can also be used to accelerate drug release from core-shell hybrid
alcohol. Then, CPT was incorporated to the organic component of the core, followed by
coating with a thin layer of silica to eliminate undesirable drug leakage [191]. Here, the
accuracy. The intracellular release of CPT from nanocarriers in A549 cells was
stimulated by high frequency magnetic field (HFMF, 2.5 kA m-1, 50 kHz) for 30-120 s.
Upon magnetic stimulus, CPT release profile for the nanocarriers in different eluting
media indicated that the rate of drug diffusion was considerably increased, which is due
incubated for 18 h, and the anticancer effect was measured by MTT assay. It was found
that CPT IC50 value measured on A549 human lung adenocarcinoma epithelial cells was
around five times lower than that of the free drug when CPT was loaded into magnetically
CPT, protected by a lipid–polymer hybrid shell [193]. In vitro testing of these drug delivery
capsules in MT2 mouse breast cancer cells showed significantly suppressed cancer cell
52
4.6. Other stimuli-responsive systems for camptothecin delivery
other analogue entities have been proposed for CPT delivery and controlled release
advanced approaches.
therapeutic payload. Under pathological local high temperature records (for instance,
those reported in inflammation and cancer) these systems undergo a phase transition
and conformational change that compromise the integrity of the nanostructure and
activate payload release. At this point, polymers can be designed in such a manner so
that this phase transition, which is termed as the lower critical solution temperature
(LCST), takes place at around 37 ºC. Under LCST the hydrogel nanopreparation is
soluble for in vivo injection, but becomes insoluble and aggregated in the tumor regions
heated above its LCST, leading to local drug discharge [194]. In this context, CPT
regular spherical core–shell nanoscale micelles of about 100-120 nm diameter and LCST
drug release above the LCST is correlated with the enhanced hydrophobicity of PNIPAM
shells, which leads to micelle shrinkage, and makes CPT diffuse out quickly. In vitro
cytotoxicity studies carry out by MTT assay over MDA-MB231 human breast cancer cell
53
line, demonstrated slightly lower activity for the CPT-loaded micelles than for the free
release at the target cells, preventing side effects to healthy tissues. Ultrasound waves
can trigger drug through the thermal and/or mechanical effects generated by cavitation
phenomena or radiation forces. Here, physical forces associated to cavitation can induce
and coconut oil as the core of the inner phase, which incorporates CPT molecules. Then,
a stabilizing coating made of phospholipids and/or Pluronic F68 was applied, obtaining
a mean droplet diameter of 220-420 nm. For the study of ultrasound influence on CPT
release the nanomulsion was exposed to ultrasound using a 1 MHz probe with a 1.5
W/cm2 intensity for 2 h. However, CPT released from the nanoemulsions was limited,
with less than 15% of the drug being discharged over 48 h. Moreover, it was found that
droplet diameter had an influence in release rate, as the amount of exposed surface
decreased for bigger droplets, which slows down CPT liberation. Furthermore, in order
nanoemulsions and, next, sonication with ultrasound energy was used to stimulate the
they began to oscillate or resonate and, eventually, bubbles were destroyed and
provoked a local drug release. Unfortunately, no biological study is reported with this
potential application as DDSs. Most of these systems have been developed as platforms
54
for glycemia control in diabetes mellitus patients. Among them, the most studied model
is that based on phenylboronic acid [198]. In aqueous medium, phenylboronic acid holds
an equilibrium between an uncharged form and a charged form. The first one is
hydrophobic but the second form is hydrophilic. This last is able to form a stable
phenylborate with cis-diol compounds, like glucose, which shifts the equilibrium in the
direction of increasing the hydrophilic form. Usually, this shift is enough to increase the
causing the entrapped payload to be released into solution. This provides a way for
designing polymeric carriers for self-regulated delivery not only of insulin but also of other
therapeutic agents. For the purpose of CPT delivery, hybrid nanoparticles with iron oxide
and mesoporous silica were loaded with the drug and further capped with a
mesopores followed by incubation with dextran (Mw 6000 and 100000), which bounded
with phenylborate linker via glucose monomer 1,2-diol groups (Fig. 21). Consequently,
dextran closes the surface of pores, preventing drug release. CPT release mechanism
is based on the competition between soluble glucose and dextran for binding
55
phenylboronic acid. Glucose has one 1,2-cis diol group that binds strongly to
phenylborate linker, removing the larger size dextran from pore surface. In vitro results
over HeLa human cervix carcinoma cell line showed a significant reduction of cell viability
in glucose containing RPMI cell culture medium, but no effect in the absence of glucose,
showing that drug loaded particle offers glucose concentration-dependent cell viability.
To end this section, we can summarize that stimuli-responsive systems are able to
deliver CPT to tumor tissue and cells with minimum (or meaningless) premature release,
imposing an efficient drug discharge inside cells under a specific stimulus. This is
particularly important in the case of very cytotoxic molecules like CPT, as these DDSs
increase anti-tumor activity and minimize side effects. It must be taken into account that
for most DDSs, less than ~5% of the injected dose reaches the tumor site [161]. In this
represent a safe alternative to regular chemotherapy with CPT structural analogues. The
main concern with most of these systems, for translation from the bench to the bedside,
is related to their chemical complexity, as most of them are hybrid materials with organic
and inorganic components, and several-step synthesis protocols. This hampers not only
the scaling-up of their synthesis, but also their degradability and biocompatibility.
Moreover, the ability of these models to be sensitive to discrete variations of pH, redox
potential or enzyme activity is not always straightforward to achieve, whereas the scope
Nevertheless, some of these systems are on the way towards achieving clinical
5. Clinical testing
One of the most disputable arguments again nanomedicines is that, despite the strong
publication statistics over the last decade, very few drugs have reached the commercial
market yet [79]. Although it is accepted that nanocarriers improve solubility, efficacy and
56
biodistribution, while decreasing adverse side effects of very cytotoxic drugs as CPT,
strict regulations concerning their use in human beings have hampered, so far, a quicker
translation of these therapeutic platforms into the clinical stage. At this point, it must be
taken into account that nanocarriers are evaluated for each of their components within
the formulation, and aspects like colloid stability in biological fluids, low immunogenicity
and long-term toxicity may not always be assessed. Moreover, the necessary refining of
the delivery process performance is yet to be achieved [200]. In this context, only a few
organic polymers have been approved by the Food and Drug Administration (FDA) as
CPT delivery systems for cancer therapy. Below we describe briefly some of the most
representative examples as well as their current position at the pipeline. All these cases
have been compiled in Table 2. We accept that some of these DDSs might not form solid
nanoparticles in real in vivo conditions [201], but in all cases, the methods followed for
linked to CPT, and was designed to enhance drug solubility and biodistribution [117].
The antitumor drug was released by hydrolysis of the ester bond between the
cyclodextrin and CPT hydroxyl group at C20 position. With regards to irinotecan, this
cells, Panc-1 human pancreatic cells and HT29 human colorectal adenocarcinoma cell
line [117,202], entering Phase Ia trial for the treatment of solid tumors over 10 patients
in 2006. Also, a combined Phase Ib/IIa trial over 62 patients with advanced solid
This preparation was tolerated with no evident side effects. Consequently, a Phase II
trial was started in 2008 for therapy of 29 ovarian cancer patients and, later on, in 2011,
a Phase II trial was launched over 157 patients with non-small cell lung cancer, in order
57
to assess the activity and safety of CRLX101. In addition, comparative preclinical studies
suggested that CRLX101 behavior in animal models could be the same than in humans
tumors by EPR effect is quite rationalized for animal models of human disease, these
models poorly represent human tumors. Nevertheless, in a very recent study over 9
gastric tumor patients consisting of collecting tumor and nonneoplastic tissue biopsies
nanoparticles in tumor tissue in five of these 9 patients, whereas no CRLX101 trace was
IX [205].
prevent CPT lactone ring cleavage. In this supramolecular system, CPT binds the
showed antitumor activity against multiple tumor cell lines, such as B16 murine
melanoma cells and H322, H460 and H1299 human lung cancer cells, and retained
substantial anti-tumor activity in syngeneic (over C57BL/6 mice) and xenogeneic (over
female nude mice) tumor models [206]. CT-2106 Phase I trial was started by Cell
Therapeutics, Inc. in 2003. A dose escalation study was carried out over 24 patients with
different advanced solid tumors [207]. The only serious related toxicity was grade 4
Cell Therapeutics, Inc. extended this Phase I trial in 2006. For this sake, 26 patients with
advanced solid tumors (50% melanoma), were intravenously weekly administered for
three consecutive weeks during a 28-day cycle [208]. Dose limiting toxicities were
rapid urinary excretion, and similar terminal half-life than free drug. 25 patients were
assessable for response, but no objective responses (complete or partial) were noted.
58
Only 3 patients had stable disease, one breast cancer patient, one histiocytoma and one
melanoma, whereas the rest exhibited progression of the disease. Overall, this trial
provided a strong indication that CT-2106 may have the ability to substantially improve
PHF or Fleximer® [209]. This compound was designed to improve efficacy and toxicity
with regards to CPT, and showed higher antitumor efficacy compared with irinotecan
against HT-29 human colon carcinoma xenografts [210]. CPT discharge involves a dual-
hydrolyze to yield CPT. A Phase I study with XMT-1001 was lauched in 2007 by Mersana
Therapeutics, Inc. to determine the maximum tolerable dose (MTD), and to assess the
safety and pharmacokinetics of XMT-1001 and its release products. Then, a total of 24
patients with small cell lung cancer (SCLC) and non-small cell lung cancer (NSCLC)
received XMT-1001 by intravenous infusion every 21 days [211]. No drug related serious
adverse events were observed during treatment, whereas several patients showed
promising evidence of clinical activity, including tumor shrinkage and stable disease for
extension study of XMT-1001 was initiated in 2011 over 11 patients (10 with NSCLC and
1 with gastric cancer), aiming at fully exploring the broad anti-tumor potential of this
59
Table 2. Nanomedicines of CPT under clinical investigation.a
Clinical Number of
Compound Formulation Pathology Dosis regimen MTDb Toxicity Ref.
trial patients
Advanced Gastric,
60 minute IV infusions on
Gastroesophageal, or
CRLX-101 Cyclodextrin-PEG copolymer Phase Ia 10 days 1, 8, and 15 of a 28 30 mg/m2/month Pancytopenia [202]
Esophageal Squamous or
day cycle
Adenocarcinoma
60 min IV infusion for 3 Myelosuppression,
Phase 15 mg/m2 IV every
62 Solid malignances consecutive weeks every grade 3/4 neutropenia [203]
Ib/IIa 2 wks
28 days and fatigue.
Ovarian Cancer IV infusion once every 14 G3 vasovagal reaction
15 mg/m2 IV every
Phase II 29 Fallopian Tube Cancer days. Each cycle is 28 G3 pneumonia G3 [117]
2 wks
Primary Peritoneal Cancer days pulmonary embolism
15mg/m2 IV every other 15mg/m2 IV every Fatigue/asthenia,
Phase II 157 Non-small cell lung cancer [117]
week 2 wks dypsnea
Neutropenia and
thrombocytopenia,
Poly(L-glutamic acid) polymer 10 min IV infusion every 3 75 mg CPT-
CT-2106 Phase I 24 Adult solid tumors febrile neutropenia, [207]
conjugate wks 2106/m2
grade 2 hematuria,
stomatitis
Grade 3 fatigue and
Advanced solid tumors (50% 10 min IV infusion weekly 25 mg CPT- grade 3 and 4
Phase I 26 [208]
melanoma) x 3 every 4 wks 2106/m2/week thrombocytopenia,
grade 2 hematuria
Polyacetal poly(1-
Small and non-small cell 1-20.5 mg CPT eq/m2 IV 113 mg CPT Myelosuppresion,
XMT-1001 hydroxymethylethylene Phase I 24 [211]
lung cancer infusión once every 3 wks (eq)/m2 diarrhea
hydroxymethylformal)
10 with non-small cell lung
IV infusion once every 3 initially 113 mg
Phase Ib 11 cancer and 1 with gastric --- [209]
wks CPT (eq)/m2
cancer
7000 mg/m2 Myelosuppression,
Solid malignances (15 with 1 h IV infusion once every
EZ-246 Pegylated CPT conjugate Phase I 37 (1.67% CPT bladder toxicity at [213]
colorectal cancer) 3 wks
weight load) 4800 mg/m2
Advanced and metastatic
Neutropenia,
adenocarcinomas of the 1 h IV infusion once every
Phase II 35 3240 mg/m2 thrombocytopenia, [214]
stomach gastroesophageal 3 wks
anaemia
junction
Metastatic, refractory solid 100 mg Myelosuppression,
N-(hydroxypropyl)methacrylamide 30 min IV infusion once
MAG-CPT Phase I 23 tumors (6 with colorectal CPT/equiv./m2 grade 2 [218]
polymer every week
cancer) every 4 wks dysuria/hematuria
a Additionalinformation may be found at www.clinicaltrials.gov.
b MTD: Maximun tolerated dose
Conversely, EZ-246 (also known as pegamotecan) is a covalent conjugate of CPT with
in vivo studies, this compound showed higher activity than irinotecan and topotecan
against human tumor xenografts (including colon, lung, breast and pancreatic origin)
[212]. Then, in a Phase I trial of this formulation, 37 patients with different solid
malignancies (15 with colorectal cancer) were administered by intravenous infusion for
toxicity was associated to the treatment, mostly neutropenia and genitourinary toxicity
[213]. Actually, free CPT accumulated slowly in plasma, with a dose proportional
Subsequently, a Phase II trial was started at 2004 over 35 patients with advanced and
Partial antitumor response was observed in 5 subjects, with a median time to progression
of about 12 weeks, and median overall survival of 38 weeks. However, severe toxicity,
appeared again in some patients, showing the need of further studies, maybe combined
Finally, MAG-CPT (also known as PNU 166148 and mureletecan) comprises CPT linked
C20 α-hydroxy with glycine [216]. These conjugates were designed primarily to improve
the clinical delivery of CPT, and utilized a hydrolytically labile spacer able to liberate drug
showed that this compound was active in colon, stomach, pancreas, ovary, breast, lung,
and melanoma xenograft models [217]. Then, a Phase I study over 23 patients with
metastatic, refractory solid tumors (6 with colorectal cancer) was carried out in 1999
[218]. Unfortunately, no objective antitumor responses were seen, and severe drug-
diarrhea. Also, one patient died after the first treatment cycle at the maximum dose.
Based on the results of the Phase I studies and the intratumoural kinetic study,
CPT [219].
CPT is one of the most active molecules for cancer treatment, but its particular limitations
as poor solubility, lack of stability of the lactone ring and strong toxicity, have encouraged
the development of different delivery forms in order to achieve a safe approach. Here, it
is a fact that current cancer treatment cannot be tackled as the administration of a single
therapeutic molecule but, instead, the new target is personalized medicine, which selects
the appropriate drug combination and dosing schedules for individual patients.
In this context, despite the fact that irinotecan and topotecan present improved
pharmacokinetic parameters, and that they are the only commercial CPT derivatives
approved by the FDA, therapies merely relying on single, small molecules seem now out
of step, mostly due to their limited efficacy and unacceptable toxicities. For this reason,
huge emphasis is being placed on developing new DDSs which are able to
simultaneously deliver CPT and other active agents, as well as incorporating imaging
moieties that may ease the diagnosis process. Most specifically, future seems good for
those systems which are able to selectively discharge the drug at the target cells under
a dramatically better spatial and temporal control over drug release, which is crucial for
62
Although no CPT nanomedicine has achieved commercial form, there are several
organic polymer compositions currently under different phases of clinical trials. In all
cases, there is a significant increase of the MTD with regards to CPT structural
positioned to achieve regular clinical use, according to the promising efficacy and
tolerability results, as well as plasma half-life values obtained in Phase II trials. However,
most of these systems show limited functionalization ability to bind covalently the drug,
which limits the loading capacity, as well as the possibility to combine CPT with other
therapeutic molecules and/or imaging agents. In this sense, next generation of CPT
delivery vehicles should merge properties of organic and inorganic materials. Whilst
organic components will provide good stability in physiological fluids, low toxicity and
(which alone could present strong limitations for biological use, due to possible
compositional toxicity and severe interaction with serum proteins), may offer almost
unlimited external surface area to bind different active molecules and stimuli-sensitive
linkers and devices. Obviously, from the regulatory point of view it is much more
challenging to develop hybrid nanocarriers, which require the evaluation of every single
component and, possibly, their clinical translation will not be as smooth as that of organic
polymers, but the potential of these systems points out to solving the strong interpatient
variability that currently hampers the widespread use of CPT nanopreparations. In this
review, we have tried to reflect some of the most interesting candidates to achieve the
clinics soon.
Acknowledgements
63
Rivero thanks the Cursol Foundation for a post-doctoral scholarship. Prof. Eduardo
Fernández and Dr. Ibane Abasolo are acknowledged for the useful discussion.
References
[1] M.E. Wall, M.C. Wani, C.E. Cook, K.H. Palmer, A.T. McPhail, G.A. Sim, Plant
Leukemia and Tumor Inhibitor from Camptotheca acuminata, J. Am. Chem. Soc. 88
(1966) 3888-3890.
[2] D.R. Beidler, J.Y. Chang, B.S. Zhou, Y.C. Cheng, Camptothecin resistance involving
[3] B.C. Giovanella, H.R. Hinz, A.J. Kozielski, J.S. Stehlin Jr., R. Silber, M. Potmesil,
Complete growth Inhibition of Human cancer xenografts in nude mice by treatment with
[5] L.H. Li, T.J. Fraser, E.J. Olin, B.K. Bhuyan, Action of Camptothecin on Mammalian
[6] A.Y. Chen, L.F. Liu, DNA topoisomerases: essential enzymes and lethal targets.
[7] N.H. Oberlies, D.J. Kroll, Camptothecin and taxol: historic achievements in natural
64
[8] M.K. Clements, S. Wasi, S.S. Daoud, Camptothecin exhibits selective cytotoxicity
towards human breast carcinoma as compared to normal bovine endothelial cells in vitro,
[9] Y.Q. Liu, W. Dai, J. Tian, L. Yang, G. Feng, X.W. Zhou, L. Kou, Y.L. Zhao, W.Q. Li,
L.H. Li, H.Y. Li, Synthesis and insecticidal activities of novel spin-labeled derivatives of
[10] L. Zhang, D. Ma, Y. Zhang, W. He, J. Yang, C. Li, H. Jiang, Characterization of DNA
[12] C.Y. Chiao, H.S. Li, Effect of topical use of camptothecine-dimethyl sulfoxide
[13] A.L. Bodley, T.A. Shapiro, Molecular and cytotoxic effects of camptothecin, a
topoisomerase I inhibitor, on trypanosomes and Leishmania, Proc. Natl. Acad. Sci. USA
92 (1995) 3726–3730.
[14] M. del Poeta, S.F. Chen, D. Von Hoff, C.C. Dykstra, M.C. Wani, G. Manikumar, J.
Heitman, M.E. Wall, J.R. Perfect, Comparison of in vitro activities of camptothecin and
nitidine derivatives against fungal and cancer cells, Antimicrob. Agents Chemother. 43
(1999) 2862–2868.
[15] A.N.M.A. Alaghaz, B.A. El-Sayad, S.A.H. Albohy, Synthesis, spectroscopic, and
[16] Y.Y. Li, S.W. Chen, L.M. Yang, R.R. Wang, W. Pang, Y.T. Zheng, The anti-HIV
65
[17] S.B. Horwitz, C.K. Chang, A.P. Grollman, Antiviral action of camptothecin.
[18] E. Boven, A.H. Van Hattum, I. Hoogsteen, H.M.M. Schlüper, H.M. Pinedo, New
Analogues of Camptothecins. Activity and resistance. Ann. NY. Acad. Sci. 922 (2000)
175-177.
[19] Y.H. Hsiang, R. Hertzberg, S. Hecht, L.F. Liu, Camptothecin induces protein-linked
DNA breaks via mammalian DNA topoisomerase I, J. Biol. Chem. 260 (1985) 14873-
14878.
[20] P.J. Tobin, L.P. Rivory, Camptothecins and Key Lessons in Drug Design, Drug
[21] W. Muqeet, B. Qudsia, Camptothecin and its analogs antitumor activity by poisoning
(2002) 641-661.
[23] S.W. Carrigan, P.C. Fox, M.E. Wall, M.C. Wani, J.P. Bowen, Comparative molecular
[26] A. Hatefi, B. Amsden, Camptothecin Delivery Methods, Pharm. Res. 19 (2002) 1389-
1399.
66
[27] J. Zhi, Q. Liu, T. Li, B. Ren, Measurement and correlation of the solubility for
[30] Y.Q. Liu, W.Q. Li, S.L. Morris-Natschke, K. Qian, L. Yang, G.X. Zhu, X.B. Wu, A.L.
Chen, S.Y. Zhang, X. Nan, K.H. Lee, Perspectives on biologically active camptothecin
[31] B.A. Hanson, R.L. Schowen, V.J. Stella, A mechanistic and kinetic study of the E-
[33] C.E. Müller, Prodrug approaches for enhancing the bioavailability of drugs with low
[34] C.J. Thomas, N.J. Rahier, S.M. Hecht, Camptothecin: current perspectives, Bioorg.
[35] R.P. Verma, C. Hansch, Camptothecins: a SAR/QSAR study, Chem. Rev. 109
(2009) 213-235.
[36] V.J. Venditto, E.E. Simanek, Cancer therapies utilizing the camptothecins: a review
67
[37] R.P. Hertzberg, M.J. Caranfa, K.G. Holden, D.R. Jakas, G. Gallagher, M.R Mattern,
S.M. Mong, J.O. Bartus, R.K. Johnson, W.D. Kingsbury, Modification of the hydroxy
[38] V. Srivastava, A.S. Negi, J.K. Kumar, M.M. Gupta, P.S. Khanuja, Plant-based
anticancer molecules: a chemical and biological profile of some important leads, Bioorg.
[39] S. Wang, Y. Li, Y. Liu, A. Lu, Q. You, Novel hexacyclic camptothecin derivatives.
(2007) 656-664.
[41] Y. Jiao, H. Liu, M. Geng, W. Duan, Synthesis and antitumor activity of 10-
[43] F. Xiao, Y. Xue, Y. Luo, B. Zhang, W. Lu, B. Yang, Synthesis and cytotoxic activity
[44] M.C. Wani, A.W. Nicholas, M.E. Wall, Plant antitumor agents. 23. Synthesis and
68
cytotoxic activity, and oral availability of new lipophilic 9-substituted camptothecin
[46] M.E. Wall, M.C. Wani, A.W. Nicholas, G. Manikumar, C. Tele, L. Moore, A.
Truesdale, P. Leitner, J.M. Besterman, Plant antitumor agents. 30. Synthesis and
[47] D.L. Emerson, J.M. Besterman, H.R. Brown, M.G. Evans, P.P. Leitner, M.J. Luzzio,
J.E. Shaffer, D.D. Sternbach, D. Uehling, A. Vuong, In vivo antitumor activity of two new
609.
3033-3039.
[49] T. G. Burke, Z. Mi, Ethyl substitution at the 7 position extends the half-life of 10-
2580-2582.
sensitive LC-MS/MS method for simultaneous quantification of sinotecan and its active
metabolite in human blood, J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 951-952
(2014) 62-68.
analogues: a review on their chemotherapeutic potential, Nat. Prod. Res. 19 (2005) 393-
412.
69
and in vitro cytotoxicity of 5-C-substituted 20(S)-camptothecin analogues, Bioorg. Med.
[53] J.X. Duan, X. Cai, F. Meng, J.D. Sun, Q. Liu, D. Jung, H. Jiao, J. Matteucci, B. Jung,
Their synthesis, preclinical activity, and potential use for cancer treatment, J. Med.
[54] A. Chatterjee, R. Digumarti, R.N. Mamidi, K. Katneni, V.V. Upreti, A. Surath, M.L.
[55] C.Y. Wang, X.D. Pan, H.Y. Liu, Z.D. Fu, X.Y. Wei, L.X. Yang, Synthesis and
[56] X. Hu, J. Hu, J. Tian, Z. Ge, G. Zhang, K. Luo, S. Liu, Polyprodrug Amphiphiles:
[57] H.M. Lin, H.Y. Lin, M.H. Chan, Preparation, characterization, and in vitro evaluation
[58] W. Ha, J. Yu, X.Y. Song, J. Chen, Y.P. Shi, Tunable temperature-responsive
70
[61] Y. Kawato, m. Aonuma, Y. Hirota, H. Kuga, K. Sato, Intracellular roles of SN- 38, a
[62] Q. Tian, J. Zhang, T.M. Tan, E. Chan, W Duan, S.Y. Chan, U.A. Boelsterli, P.C. Ho,
H. Yang, J.S. Bian, M. Huang, Y.Z. Zhu, W. Xiong, X. Li, S. Zhou, Human Multidrug
(2005)1837-1853.
[63] V. Bala, S. Rao, B.J. Boyd, C.A. Prestidge, Prodrug and nanomedicine approaches
for the delivery of the camptothecin analogue SN38, J. Control. Release 172 (2013) 48-
61.
[64] G.J. Creemers, G. Bolis, M. Gore, G. Scarfone, A.J. Lacave, J.P. Guastalla, R.
Despax, G. Favalli, R. Kreinberg, S. Van Belle, I. Hudson, J. Verweij, W.W. Ten Bokkel
cancer: results of a large European phase II study, J. Clin. Oncol. 14 (1996) 3056-3061.
[65] P. Pantazis, J.A. Early, A.J. Kozielski, J.T. Mendoza, H.R. Hinz, B.C. Giovanella,
[66] J.W. Clark, Rubitecan, Expert. Opin Inv. Drugs 15 (2006) 71-79.
[67] H.A. Burris III, S. Rivkin, R. Reynolds, J. Harris, A. Wax, H. Gerstein, K.L. Mettinger,
[68] J.Y. Jung, S. H. Song, T.Y. Kim, J.H. Park, H.S. Jong, S.A. Im, T.Y. Kim, Y.J. Bang,
N.K. Kim, The Synergism between Belotecan and Cisplatin in Gastric Cancer, Cancer
71
[69] S. Lim, B.C. Cho, J.Y. Jung, G.M. Kim; S.H. Kim, H.R. Kim, H.S. Kim, S.M. Lim, J.S.
Park, J.H. Lee, D. Kim, E.Y. Kim, M.S. Park, Y.S. Kim, S.K. Kim, J. Chang, J.H. Kim,
Phase II study of camtobell inj. (belotecan) in combination with cisplatin in patients with
previously untreated, extensive stage small cell lung cancer, Lung Cancer 80 (2013)
313-318.
[70] M.A. Garrison, L.A. Hammond, C.E.Jr. Geyer, G. Schwartz, A.W. Tolcher, L.
Smetzer, J.A. Figueroa, M. Ducharme, J. Coyle, C.H. Takimoto, R.L. De Jager, E.K.
protracted 21-day infusion in patients with advanced solid malignancies, Clin. Cancer
[71] C.J. Gerrits, J.H. Schellens, G.J. Creemers, P. Wissel, A.S. Planting, J.F. Pritchard,
Cancer nanomedicines: closing the translational gap, The Lancet 384 (2014) 2175-2176.
[75] N. Doshi, S. Mitragotri, Designer biomaterials for nanomedicine, Adv. Funct. Mater.
19 (2009) 3843-3854
[76] A. Mahapatro, D.K. Singh, Biodegradable nanoparticles are excellent vehicle for site
[77] Y. Shen, E. Jin, B. Zhang, C.J. Murphy, M. Sui, J. Zhao, J. Wang, J. Tang, M. Fan,
72
nanocapsules for intracellular cancer drug delivery, J. Am. Chem. Soc. 132 (2010) 4259-
4265.
[78] R. Duncan, Polymer therapeutics for tumor specific delivery, Chem. Ind.-London. 7
(1997) 262-264.
[79] V.J. Venditto, F.C. Szoka Jr., Cancer nanomedicines: So many papers and so few
[80] V.P. Torchilin, Multifunctional nanocarriers, Adv. Drug Deliver. Rev. 58 (2006) 1532-
1555.
[81] V. López-Dávila, A.M. Seifalian, M. Loizidou, Organic nanocarriers for cancer drug
[82] H.C. Huang, S. Barua, G. Sharma, S.K. Dey, K. Rege, Inorganic nanoparticles for
polymeric micelles, which is designed based on EPR effect. Adv. Drug Deliver. Rev. 63
(2011) 184-192.
copolymer design for camptothecin incorporation into polymeric micelles for passive
73
[88] M. Yokoyama, P. Opanasonit, T. Okano, K. Kawano, Y. Maitani, Polymer design
and incorporation methods for polymeric micelle carrier system containing water-
[90] N. Fan, K. Duan, C. Wang, S. Liu, S. Luo, J. Yu, J. Huang, Y. Li, D. Wang,
75 (2010) 543-549.
[91] W. Zhang, J. Huang, N. Fan, J. Yu, Y. Liu, S. Liu, D. Wang, Y. Li, Nanomicelle with
[92] H.C. Yen, H. Cabral, P. Mi, K. Toh, Y. Matsumoto, X. Liu, H. Koori, A. Kim, K.
[93] D.L. Emerson, Liposomal delivery of camptothecins, Pharm. Sci. Technol. To. 3
(2000) 205-209.
[94] A.M Sætern, G.E. Flaten, M. Brandl, A method to determine the incorporation
74
[96] M. Watanabe, K. Kawano, K. Toma, Y. Hattori, Y. Maitani, In vivo antitumor activity
(2011) 2101-2141.
1795.
[99] J. Lim, E.E. Simanek, Triazine dendrimers as drug delivery systems: from synthesis
[101] V.J. Venditto, K. Allred, D.C. Allred, E.E. Simanek, Intercepting the synthesis of
[104] L.W Seymour, Review: Synthetic polymers with intrinsinc anticancer activity, J.
[105] R. Duncan, The dawning era of polymer therapeutics, Nat. Rev. Drug Discov. 2
(2003) 347-360.
75
[106] S. Kunugi, T. Yamaoka, Polymers in Nanomedicine. S. Kunugi, T. Yamaoka (Ed.)
[107] R. Haag, F. Kratz, Polymer therapeutics: concepts and applications, Angew. Chem.
polymer-based prodrugs for cancer chemotherapy. Chapter 11. Y. Shen (Ed.) RSC
[110] M.J. Vicent, R. Duncan, Polymer conjugates: nanosized medicines for treating
[111] W.B. Liechty, D.R. Kryscio, B.V. Slaughter, N.A. Peppas, Polymers for Drug
[113] K. Cai, X. He, Z. Song, Q. Yin, Y. Zhang, F.M. Uckun, C. Jiang, J. Cheng, Dimeric
Drug Polymeric Nanoparticles with Exceptionally High Drug Loading and Quantitative
[114] J. Cheng, K.T. Khin, G.S. Jensen, A. Liu, W.E. Davis, Synthesis of linear, β-
(2003) 1007-1017.
[115] J. Cheng, K.T. Khin, M.E. Davis, Antitumor activity of β-cyclodextrin polymer-
76
[116] T. Schluep, J. Hwang, I.J. Hildebrandt, J. Czernin, C.H.J. Choi, C.A. Alabi, B.C.
Mack, M.E. Davis, Pharmacokinetics and tumor dynamics of the nanoparticle IT-101
from PET imaging and tumor histological measurements, P. Natl. Acad. Sci. USA 106
(2009) 11394-11399.
[118] T. Schluep, J. Hwang, J. Cheng, J.D. Heidel, D.W. Bartlett, B. Hollister, M.E. Davis,
[120] J. Zeng, H. Huang, S. Liu, H. Xu, J. Huang, J. Yu, Hollow nanosphere fabricated
tumor activity in human prostate cancer cells, Eur. J. Pharm. Sci. 47 (2012) 686-694.
[122] Y. Zou, Q.P. Wu, W. Tansey, D. Chow, M.C. Hung, C. Charnsangavej, S. Wallace,
resistant human lung cancer xenografted in nude mice, Int. J. Oncol. 18 (2001) 331-336.
[123] J.W. Singer, R. Bhatt, J. Tulinsky, K.R. Buhler, E. Heasley, P. Klein, P. de Vries,
77
[125] K.H. Min, K. Park, Y.S. Kim, S.M. Bae, S. Lee, H.G. Jo, R.W. Park, I.S. Kim, S.Y.
encapsulated camptothecin enhance the drug stability and tumor targeting in cancer
[126] S. Soukasene, D.J. Toft, T.J. Moyer, H. Lu, H.K. Lee, S.M. Standley, V.L. Cryns,
[127] H.C. Huang, S. Barua, G. Sharma, S.K. Dey, K. Rege, Inorganic nanoparticles for
[129] J. Nam, N. Won, J. Bang, H. Jin, J. Park, S. Jung, S. Jung, Y. Park, S. Kim, Surface
engineering of inorganic nanoparticles for imaging and therapy, Adv. Drug Deliv. Rev.
65 (2013) 622-648.
[130] Y. Zhu, J. Lei, Y. Tian, Uniform iron oxide hollow spheres for high-performance
[131] P.M. Castillo, M. Mata, M.F. Casula, J.A. Sánchez-Alcázar, A.P. Zaderenko,
[133] N.G. Sahoo, H. Bao, Y. Pan, M. Pal, M. Kakran, H.K.F. Cheng, L. Li, L.P. Tan,
5237.
78
[134] M. Assali, J.J. Cid, M. Pernía-Leal, M. Muñoz-Bravo, I. Fernández, R.E. Wellinger,
[137] Z. Li, E. Ye, D., R. Lakshminarayanan, X.J. Loh, Recent Advances of Using Hybrid
[139] J. Lu, M. Liong, J.I. Zink, F. Tamanoi, Mesoporous silica nanoparticles as a delivery
[140] J. Lu, M. Liong, Z. Li, J.I. Zink, F. Tamanoi, F, Biocompatibility, biodistribution, and
[141] J. Lu, Z. Li, J.I. Zink, F. Tamanoi, In vivo tumor suppresion efficacy of mesoporous
[142] M. Ma, H. Chen, Y. Chen, K. Zhang, X. Wang, X. Cui, J. Shi, Hyaluronic acid-
organosilicas with organic groups inside the channel walls, Nature 402 (1999) 867-871.
79
[144] S. Inagaki, S. Guan, Y. Fukushima, T. Ohsuna, O. Terasaki, Novel mesoporous
materials with a uniform distribution of organic groups and inorganic oxide in their
[145] B.J. Melde, B.T. Holland, C.F. Blanford, A. Stein, Mesoporous sieves with unified
3235-3272.
[147] Y. Chen, Q. Meng, M. Wu, S. Wang, P. Xu, H. Chen, Y. Li, L. Zhang, L. Wang, J.
Hybridization Approach for Biomedicine, J. Am. Chem. Soc. 136 (2014) 16326-16334.
[149] X. Yan, C. Yu, X. Zhou, J. Tang, D. Zhao, Highly ordered mesoporous bioactive
glasses with superior in vitro bone-forming bioactivities, Angew. Chem. Int. Edit. 43
(2004) 5980-5984
[150] M.H. Hsiao, T.H. Tung, C.S. Hsiao, D.M. Liu, Nano-hybrid carboxymethyl chitosan
[151] K.M. Tyner, S.R. Schiffman, E.P. Giannelis, Nanobiohybrids as delivery vehicles
(2010) 829-843.
80
[153] H. Ünal, I. d’Angelo, E. Pagano, F. Borrelli, A. Izzo, F. Ungaro, F. Quaglia, E.
[154] T. Chen, S. Xu, T. Zhao, L. Zhu, D. Wei, Y. Li, H. Zhang, C. Zhao, Gold
[156] B. Xu, Y. Ju, Y. Cui, G. Song, Y. Iwase, A. Hosoi, Y. Morita, tLyP-1-conjugated Au-
organic frameworks as efficient materials for drug delivery, Angew. Chem. Int. Edit. 45
(2006) 5974-5978.
[158] J.D. Rocca, D. Liu, W. Lin, Nanoscale metal-organic frameworks for biomedical
organic frameworks and their bio-related applications, Coord. Chem. Rev. 307 (2016)
342-360.
[160] J. Zhuang, C.H. Kuo, L.Y. Chou, D.Y. Liu, E. Weerapana, C.K. Tsung, Optimized
[161] X.H. Ji, X.B. Zhao, Y.H. Zhang, B.H. Lu, H.L. Ni, Solvothermal synthesis and
thermoelectric properties of lanthanum contained Bi-Te and Bi-Se-Te alloys, Mater. Lett.
59 (2005) 682-685.
81
[162] S. Ganta, H. Devalapally, A. Shahiwala, M. Amiji, A review of stimuli-responsive
nanocarriers for drug and gene delivery, J. Control. Release 126 (2008) 187-204.
[164] E. Fleige, M.A. Quiadir, R. Haag, Stimuli-responsive polymeric nanocarriers for the
controlled transport of active compounds: Concepts and applications, Adv. Drug Deliver.
[166] M. Kanamala, W.R. Wilson, M. Yang, B.D. Palmer, Z. Wu, Mechanisms and
(2016) 152-167.
[167] K.H. Min, J.H. Kim, S.M. Bae, H. Shin, M.S. Kim, S. Park, H. Lee, R.W. Par, I.S.
Kim, K. Kim, I.C. Kwon, S.Y. Jeong, D.S. Lee, Tumoral acidic pH-responsive MPEG-
poly(β-amino ester) polymeric micelles for cancer targeting therapy, J. Control. Release
[169] D.L. Tang, F. Song, C. Chen, X.L. Wang, Y.Z. Wang, A pH-responsive chitosan-b-
[170] Z. Li, H. Li, L. Liu, X. You, C. Zhang, Y. Wang, A pH-sensitive nanocarrier for co-
82
[171] G.K. Balendiran, R. Dabur, D. Fraser. The role of glutathione in cancer, Cell
[172] H. Wang, L. Tang, C. Tu, Z. Song, Q. Yin, L. Yin, Z. Zhang, J. Cheng, Redox
[175] A.G. Cheetman, P. Zhang, Y.A. Lin, L.L. Lock, H. Cui, Supramolecular
nanostructures formed by anticancer drug assembly, J. Am. Chem. Soc. 135 (2013)
2907-2910.
[176] A.G. Cheetham, Y.C. Ou, P. Zhang, H. Cui, Linker-determined drug release
[177] Y.A. Lin, A.G. Cheethman, P. Zhang, Y.C. Ou, Y. Li, G. Liu, D. Hermida-Merino,
delivery of camptothecin and its biological evaluation, J. Control. Release 156 (2011)
246-257.
83
[180] K.J. Chen, L. Tang, M.A. Garcia, H. Wang, H. Lu, W.Y. Lin, S. Hou, Q. Yin, C.K.
[182] R. Tong, D.S. Kohane, Shedding Light on Nanomedicine, Wiley Interdiscip. Rev.
[186] X.Tan, B.B. Li, X. Lu, F. Jia, C. Santori, P. Menon, H. Li, B. Zhang, J.J. Zhao, K.
[187] L. Chen, W. Wang, B. Su, Y. Wen, C. Li, Y. Zhou, M. Li, X. Shi, H. Du, Y. Song, L.
84
[189] S. Giri, B.G. Trewyn, M.P. Stellmaker, V.S. Lin, Stimuli-responsive controlled-
release delivery system based on mesoporous silica nanorods capped with magnetic
[190] P.J. Chen, S.H. Hu, C.S. Hsiao, Y.Y. Chen, D. Liu, S.Y. Chen, Multifunctional
(2011) 2535-2543.
[191] W.L. Tung, S.H. Hu, D.M. Liu, Synthesis of nanocarriers with remote magnetic drug
release control and enhanced drug delivery for intracellular targeting of cancer cells. Acta
[192] S.H. Hu, T.Y. Liu, H.Y. Huang, D.M. Liu, S.Y. Chen, Magnetic-sensitive silica
[193] S.D. Kong, M. Sartor, C.M.J. Hu, W. Zhang, L. Zhang, S. Jin, Magnetic field
[195] Y.L. Luo, X.L. Yang, F. Xu, Y.S. Chen, B. Zhang, Thermosensitive PNIPAM-b-
HTPB block copolymer micelles: Molecular architectures and camptothecin drug release,
[196] J.Y. Fang, C.F. Hung, S.C. Hua, T.L. Hwang, Acoustically active perfluorocarbon
nanoemulsions as drug delivery carriers for camptothecin: drug release and cytotoxicity
85
[198] R. Ma, L. Shi, Phenylboronic acid-based glucose-responsive polymeric
nanoparticles: synthesis and applications in drug delivery, Polym. Chem. 5 (2014) 1503-
1518.
nanoparticle for glucose-responsive and targeted drug delivery, ACS Appl. Mater.
[200] S. Wilhelm, A.J. Tavares, Q. Dai, S. Ohta, J. Audet, H.J. Dvorak, W.C.W. Chan,
[201] S. Svenson, What nanomedicine in the clinic right now really forms nanoparticles?,
[203] G.J. Weiss, J. Chao, J.D. Neidhart, R.K. Ramanathan, D. Bassett, J.A. Neidhart,
C.H. Choi, W. Chow, V. Chung, S.J. Forman, E. Garmey, J. Hwang, D.L. Kalinoski, M.
Koczywas, J. Longmate, R.J. Melton, R. Morgan, J. Oliver, J.J. Peterkin, J.L. Ryan, T.
Schluep, T.W. Synold, P. Twardowski, M.E. Davis, Y. Yen, First-in-human phase 1/2a
in patients with advanced solid tumor malignancies, Invest. New Drugs 4 (2013) 986-
1000.
[204] S. Eliasof, D. Lazarus, C.G. Peters, R.I. Case, R.O. Cole, J. Hwang, T. Schluep, J.
Chao, J. Lin, H. Han, D.T. Wiley, J.E. Zuckerman, M.E. Davis, Correlating preclinical
[205] A.J. Clark, D.T. Wiley, J.E. Zuckerman, P. Webster, J. Chao, J. Lin, Y. Yen, M.E.
86
nonneoplastic tissue after intravenous dosing, P. Natl. Acad. Sci. USA 113 (2016) 3850-
3854.
[206] J.W. Singer, P. De Vries, R. Bhatt, J. Tulinsky, P. Klein, C. Li, L. Milas, A.R. Lewis,
(2004) 3127.
[208] J. Homsi, G.R. Simon, C.R. Garrett, G. Springett, R. De Conti, A. Chiappori, P.N.
Munster, M.K. Burton, S. Stromatt, C. Allievi, P. Angiuli, A. Eisenfeld, D.M. Sullivan, A.I.
in Patients with Advanced Solid Malignancies. Clin. Cancer Res. 13 (2007) 5855-5861.
in clinical development for patients with advanced cancer. Adv. Drug Deliver. Rev. 61
(2009) 1193-1202.
[210] M.D. Walsh, S.K. Hanna, J. Sen, S. Rawal, C.B. Cabral, A.V. Yurkovetskiy, R.J.
Fram, T.B. Lowinger, W.C. Zamboni, Pharmacokinetics and Antitumor Efficacy of XMT-
of intravenous XMT-1001 in patients with advanced solid tumors, Mol. Cancer Ther. 6
(2007) A146.
[212] R.B. Greenwald, C.D. Conover, Y.H. Choe, Poly(ethylene glycol) conjugated drugs
and prodrugs: A comprehensive review. Crit. Rev. Ther. Drug 17 (2000) 101-161.
87
[213] E.K. Rowinsky, J. Rizzo, L. Ochoa, C.H. Takimoto, B. Forouzesh, G. Schwartz,
Every 3 Weeks in Patients With Advanced Solid Malignancies, J. Clin. Oncol. 21 (2003)
148-157.
[214] L.C. Scott, T. Evans, J.C. Yao, A.I. Benson, M. Mulcahy, A. Thomas, M. Decatris,
supplement 4030.
[215] L.C. Scott, J.C. Yao, A.B. Benson 3rd, A.L. Thomas, S. Falk, R.R. Mena J. Picus,
J. Wright, M.F. Mulcahy, J.A. Ajani, T.R. Evans, A phase II study of pegylated-
363-370.
experience and lessons learnt. Adv. Drug Deliver. Rev. 61 (2009) 1131-1148.
[217] C.D. Conover, R.B. Greenwald, A. Pendri, C.W. Gilbert, K.L. Shum, Camptothecin
its conjugation to polyethylene glycol via a glycine linker, Cancer Chemoth. Pharm. 42
(1998) 407-414.
Schellens, W.W. ten Bokkel Huinink, A phase I and pharmacokinetic study of MAG-CPT,
88
[219] D. Bissett, J. Cassidy, J.S. Bono, F. Muirhead, M. Main, L. Robson, D. Fraier, M.L.
89