+ All Categories
Home > Documents > Inductively coupled plasma mass spectrometry: a unique ...Inductively coupled plasma mass...

Inductively coupled plasma mass spectrometry: a unique ...Inductively coupled plasma mass...

Date post: 21-Feb-2020
Category:
Upload: others
View: 18 times
Download: 1 times
Share this document with a friend
264
Inductively coupled plasma mass spectrometry: a unique, ultrasensitive tool for exploring the pharmacology of metal-based anticancer agents
Transcript

Inductively coupled plasma mass spectrometry:

a unique, ultrasensitive tool for exploring the

pharmacology of metal-based anticancer agents

ISBN/EAN: 978-90-393-4683-9

© 2007 Elke Brouwers, Amsterdam

Cover design: Beeldhouwer Jos Reniers, www.josreniers.nl

Printed by: Ponsen & Looijen BV, Wageningen, The Netherlands

Inductively coupled plasma mass spectrometry:

a unique, ultrasensitive tool for exploring the

pharmacology of metal-based anticancer agents

Inductief gekoppelde plasma-massa-spectrometrie:

een uniek, zeer gevoelig hulpmiddel voor het onderzoeken van de

farmacologie van metaal bevattende antikanker middelen

(met een samenvatting in het Nederlands)

PROEFSCHRIFT

ter verkrijging van de graad van doctor

aan de Universiteit Utrecht

op gezag van de rector magnificus, prof. dr J.C. Stoof,

ingevolge het besluit van het college voor promoties

in het openbaar te verdedigen

op vrijdag 16 november 2007 des middags te 12.45 uur

door

Elke Elvira Margaretha Brouwers

geboren op 8 november 1977 te Helmond

Promotoren:

Prof. Dr J.H. Beijnen

Prof. Dr J.H.M. Schellens

The research described in this thesis was performed at the Department of

Pharmacy & Pharmacology, Slotervaart Hospital/The Netherlands Cancer

Institute, Amsterdam, The Netherlands

This research was supported financially by The Netherlands Organisation for

Health Research and Development, ZonMw (OND1307436)

Publication of this thesis was financially supported by:

Varian, inc., Mulgrave, Victoria, Australia;

The Netherlands Laboratory for Anticancer Drug Formulation, Amsterdam, The

Netherlands.

Het mooiste van een droom,

is de mogelijkheid om hem te verwezenlijken

Voor Remco

Contents

Chapter 1 Introduction 1.1 Aim and outline 13 1.2 The application of inductively coupled plasma mass

spectrometry in clinical pharmacological oncology research 17

Chapter 2 Determination of platinum and ruthenium in biological fluids

2.1 Determination of oxaliplatin in human plasma and plasma

ultrafiltrate by graphite-furnace atomic-absorption- spectrometry 73

2.2 Sensitive inductively coupled plasma mass spectrometry

assay for the determination of platinum originating from cisplatin, carboplatin, and oxaliplatin in human plasma ultrafiltrate 89

2.3 Determination of ruthenium originating from the

investigational anti-cancer drug NAMI-A in human plasma ultrafiltrate, plasma, and urine by inductively coupled plasma mass spectrometry 109

Chapter 3 Determination of platinum-DNA adducts 3.1 Inductively coupled plasma mass spectrometric analysis of

the total amount of Pt-DNA adducts in peripheral blood mononuclear cells and tissue from patients treated with

cisplatin 133

3.2 The effects of sulfur-containing compounds and gemcitabine on the binding of cisplatin to plasma proteins and DNA determined by ICP-MS and HPLC-ICP-MS 151

Chapter 4 Persistent effects of platinum agents 4.1 Long-term platinum retention after treatment

with cisplatin and oxaliplatin 175

4.2 Persistent neuropathy after treatment with cisplatin and oxaliplatin 193

Chapter 5 Environmental monitoring of platinum agents 5.1 Monitoring of platinum surface contamination in seven

Dutch hospital pharmacies using inductively coupled plasma mass spectrometry 215

Conclusions and perspectives 238 Summary 244 Samenvatting 248 Dankwoord 254 Curriculum Vitae 259 List of Publications 260

Chapter 1

Introduction

Chapter 1.1

Aim and outline

Chapter 1.1

14

Aim

Research to unravel the pharmacokinetics of metal-based anticancer agents is required

to understand the clinical behaviour of the drugs and to further optimise treatment

regimens. Accurate and sensitive methods for the quantitative determination of metal-

based anticancer agents are indispensable to investigate these aspects. Until recently,

many studies relied on atomic absorption spectrometry (AAS) for the analysis of

platinum (Pt) and ruthenium (Ru). The sensitivity of this technique, however, only allows

the investigation of pharmacokinetics during or shortly after therapy. The sensitivity is

insufficient to answer research questions, which are of current interest. Inductively

coupled plasma mass spectrometry (ICP-MS) does provide this high sensitivity.

For this thesis project, the major aim was to develop and validate analytical ICP-MS

methods for the analysis of metal-based anticancer agents. These methods were applied

to answer research questions concerning long-term pharmacokinetics, Pt-induced side

effects, the effects of antidotes on Pt-induced side effects, and environmental

monitoring.

Outline

Chapter 1.2 provides background information on the mechanism of metal-based

anticancer agents. Furthermore, it provides an extensive overview of publications

describing the analysis of Pt and Ru using ICP-MS in the field of oncology. The focus is on

the determination of the total metal concentration and on the speciation of Pt and Ru

compounds in human biological fluids, DNA- and protein-adducts, and environmental

samples. Chapter 2 describes the development and validation of assays for the analysis

of Pt in plasma and plasma ultrafiltrate (pUF) using atomic absorption spectrometry

(Chapter 2.1), of Pt in pUF using ICP-MS (Chapter 2.2), and of Ru in plasma, pUF, and urine

using ICP-MS (Chapter 2.3). Chapter 3 describes the use of ICP-MS for the determination

of Pt adducts. In Chapter 3.1, the development, optimisation, and validation of an ICP-MS

method for the determination of Pt bound to DNA in peripheral blood mononuclear

cells (PBMCS) and tissue was described. The method was applied to study Pt-DNA

adduct levels in PBMCs and tissue from patients treated with cisplatin. In Chapter 3.2, the

effect of sulfur-containing compounds and gemcitabine on Pt-protein and Pt-DNA

adduct levels was quantified. Chapter 4 describes the long-term effects of cisplatin and

oxaliplatin treatment. Chapter 4.1 illustrates the long-term pharmacokinetics of Pt. Pt

levels in plasma of 45 patients treated with cisplatin or oxaliplatin were monitored 8-75

months after the end of their treatment. To evaluate whether the remaining Pt was still

reactive, the Pt-DNA and Pt-protein binding characteristics of the Pt from the patients’

samples were quantified. In addition, the relationships between several determinants

Aim and outline

15

and Pt levels were evaluated. The same group of patients as described in Chapter 4.1 was

used to evaluate persistent Pt induced neuropathy (Chapter 4.2). Again, relationships

between several determinants and neuropathy were investigated, among which were

the plasma Pt levels.

Chapter 5 describes the development of a method to monitor Pt surface contamination

within hospital pharmacies. The method was applied to assess surface contamination in

seven Dutch hospital pharmacies.

In Chapter 6, the presented results are evaluated and placed in a broader perspective

and future research is discussed.

Chapter 1.2

The application of inductively coupled plasma mass spectrometry in clinical

pharmacological oncology research

Elke E.M. Brouwers Matthijs M. Tibben

Hilde Rosing Jan H.M. Schellens

Jos H. Beijnen

Submitted for publication

Chapter 1.2

18

Contents

1 Introduction 2 Analytical ICP-MS assays: general aspects

2.1 Technique 2.2 Interferences 2.3 Combination of ICP-MS detection with speciation techniques 2.4 Method validation

3 Analytical ICP-MS assays: total metal determination 3.1 Application assays

3.1.1 Metal-based anticancer agents in biological fluids/cells 3.1.2 Metal-based anticancer agents bound to DNA 3.1.3 Metal-based anticancer agents in environmental samples

3.2 Assay development 3.2.1 Sample pretreatment

3.2.1.1 Metal-based anticancer agents in biological fluids/cells 3.2.1.2 Metal-based anticancer agents bound to DNA 3.2.1.3 Metal-based anticancer agents in environmental samples

3.2.2 Calibration 3.2.3 Instrumental adjustments

4 Analytical ICP-MS assays: speciation of metal-based anticancer agents 4.1 Application assays

4.1.1 Speciation of metal-based compounds and metabolites 4.1.2 Speciation of reaction products of metal-based anticancer compounds

with DNA and proteins 4.1.3 Speciation of metal-based anticancer compounds in environmental

samples 4.2 Assay development

4.2.1 Reversed phase chromatography (RP) 4.2.2 Reversed phase ion-pairing chromatography (RPIP) 4.2.3 Size exclusion chromatography (SEC) 4.2.4 Ion-exchange chromatography (IEC) 4.2.5 Speciation techniques other than liquid chromatography

5 Conclusions and perspectives

ICP-MS in oncology

19

Abstract

Metal-based anticancer agents are frequently used in the treatment of a wide variety of

cancer types. The monitoring of these anticancer agents in biological samples is

important to understand their pharmacokinetics, pharmacodynamics, and metabolism.

In addition, determination of metals originating from anticancer agents is relevant to

assess occupational exposure of health care personnel working with these drugs. The

high sensitivity of inductively coupled plasma mass spectrometry (ICP-MS) has resulted

in an increased popularity of this technique for the analysis of metal-based anticancer

drugs. In addition to the quantitative analysis of the metal of interest in a sample, ICP-MS

can be used as an ultrasensitive metal selective detector in combination with speciation

techniques such as liquid chromatography. In the current review we provide a

systematic survey of publications describing the analysis of platinum- and ruthenium-

containing anticancer agents using ICP-MS, focused on the determination of total metal

concentrations and on the speciation of metal compounds in biological fluids, DNA- and

protein-adducts, and environmental samples. We conclude that ICP-MS is a powerful

tool for the quantitative analysis of metal-based anticancer agents from multiple sample

sources.

Chapter 1.2

20

1 Introduction

Many heavy metals are considered to be harmful to humans. However, the toxic effects

of some metals can be positively used to treat patients suffering from cancer. The first

metal-containing anticancer agent was discovered in the 1960s by Rosenberg et al [1].

While investigating the possible effects of an electric field on growth processes in

bacteria, these authors discovered that electrolysis products from platinum (Pt)

electrodes produced an inhibition of the cell division process. After the identification of

cisplatin (cis-diamminedichloridoplatinum(II)) (Figure 1) as one of the species

responsible for this anti-proliferative effect, the compound was successfully developed

into one of the most widely used anticancer agents.

Unfortunately, the use of cisplatin is hampered by severe side effects, such as ototoxicity,

nephrotoxicity and neurotoxicity and by the intrinsic and acquired resistance of several

tumour types. These limitations have stimulated the search for other metal-containing

cytotoxic compounds with better safety profiles and enhanced antitumour

characteristics. Thousands of compounds have been synthesised and evaluated in the

past 40 years and only few of these agents have entered clinical trials. Besides cisplatin,

nowadays, carboplatin (cis-diammine(1,1-cyclobutanedicarboxylato)platinum(II)) and

oxaliplatin ([(1R,-2R)-1,2-cyclohexanediamine-N,N´][oxalato(2-)-O,O´]platinum) (Figure 1)

have found important clinical applications in the treatment of cancer. Currently, also the

orally administered satraplatin (platinum(IV) cis-dichloro-trans-bis(acetato-O)ammine

(cyclohexanamine)) is under consideration for approval for the treatment of hormone

refractory prostate cancer [2]. Furthermore, ruthenium (Ru) complexes (Figure 2) are

regarded as promising alternatives for Pt complexes. NAMI-A [Imidazolium-

trans(imidazole)(dimethylsulfoxide) tetrachloro ruthenate(III)] [3], and KP1019 or FFC14A

(Indazolium trans-[tetrachlorobis(1H-indazole)ruthenate(III)] [4,5] are the first Ru

complexes that have finished phase I studies.

As the interest in metal-based anticancer agents grows, there is an increasing need for

accurate and sensitive methods for the quantitative determination of these compounds.

The determination of Pt or Ru in clinical samples of patients treated with these

compounds is required to understand the pharmacokinetics and pharmacodynamics of

these drugs. The clinical matrices of interest are usually either tissue or biological fluids

such as blood, serum, plasma, plasma ultrafiltrate, and urine. Also adducts of the

compounds and metabolites with DNA, proteins, and small molecules are of importance.

In addition to the analysis of clinical samples, methods for determination of metal-based

anticancer agents can be employed to assess occupational exposure of health care

personnel working with these drugs. This can be done by monitoring e.g. blood or urine

of personnel to measure the physical uptake of the drugs or by surface sampling to

assess contamination of environments where the drugs are processed.

ICP-MS in oncology

21

Pt

Cl

Cl

NH3

NH3

O

O

O

O

Pt

NH3

NH3

Cisplatin Carboplatin

Pt

Cl

ClNH3

N

CH3

Pt

O

O

O

O

NH

2

NH2

ZD0473 Oxaliplatin

PtCl

O

O

ClNH3

NH2

CCH3

O

CCH3

O

Pt

ClCl

ClCl

NH

2

NH2

Satraplatin Ormaplatin

NH3

Pt

Cl NH3

Pt

NH3 NH2(CH2)6H2N

NH3

Pt

Cl

NH3

NH3NH2(CH2)6H2N

BBR3464

Figure 1. Structural formula of platinum anticancer agents

4+

Chapter 1.2

22

N

NH

-

NH+

NH

RuClCl

Cl Cl

SO(CH3)2

NAMI-A

KP1019

Figure 2. Structural formula of ruthenium anticancer agents

Because of the limited sample availability and the low drug concentrations present in

these matrices, sensitive and specific methods are needed. Up to now, numerous

techniques have been used for the study of Pt and Ru anticancer agents. The assays can

be roughly divided into two groups. The first group comprises methods for the

determination of total metal concentrations utilising techniques such as atomic

absorption spectrometry (AAS) [6-8], voltammetry [9-13], differential pulse polarography

(DPP) [14], neutron activation analysis (NAA) [15] [16], x-ray spectrometry [17], x-ray

fluorescence (XRF) [18-20], inductively coupled plasma atomic emission spectrometry

(ICP-AES) [21], and inductively coupled plasma mass spectrometry (ICP-MS). The second

group includes methods for the speciation of the various Pt or Ru species. Usually, a

speciation technique such as high performance liquid chromatography (HPLC) is

coupled to a diode array detector [22], electrochemical detector [23], UV detector

[24,25], or to an element specific detector such as ICP-MS. By combining speciation

techniques with electrospray ionisation mass spectrometry (ESI-MS) [26-28] information

on structural composition can be achieved.

In this review we will focus on ICP-MS. Since 1990, when the first ICP-MS assay for the

analysis of Pt anticancer agents was published [29], ICP-MS has acquired increasing

popularity in the field of analysis of metal-based anticancer drugs. It has been applied for

the analysis of various Pt and Ru compounds (Figure 1 and 2). The technique is highly

-

NH+NHCl

NNH

NNH

RuCl

Cl

Cl

ICP-MS in oncology

23

sensitive and is applicable to a wide range of sample matrices including those of

biological and environmental origin. As a result of the successful application of ICP-MS in

the field of oncology, the number of publications on the quantitative analysis of Pt and

Ru using ICP-MS and speciation techniques coupled to ICP-MS has increased

tremendously over the last twenty years. The papers have appeared in a large range of

scientific journals, covering the many disciplines this research comprises, e.g. medicine,

pharmacy, and chemistry.

The purpose of the current review is to provide a selected, systematic survey of

publications describing the analysis of Pt and Ru using ICP-MS in the field of oncology.

The focus is on the determination of the total metal concentration and on the speciation

of Pt and Ru compounds in human biological fluids, DNA- and protein-adducts, and

environmental samples. Problems encountered when developing an ICP-MS assay with

or without combination with a speciation technique are discussed.

2 Analytical ICP-MS assays: general aspects

2.1 Technique

As the name implies, ICP-MS is a combination of an inductively coupled plasma (ICP)

with a mass spectrometer (MS) (Figure 3). Typically, the sample is introduced into the ICP

by a sample introduction system consisting of a peristaltic pump and a nebuliser, which

generates a fine aerosol in a spray chamber. The spray chamber separates the small

droplets from the large droplets. Large droplets fall out by gravity and exit through the

drain tube at the end of the spray chamber, while the small droplets pass between the

outer wall and the central tube and are eventually transported into the sample injector

of the plasma torch using a flow of argon gas. The aerosol is then transported to the ICP,

which is a plasma ion source. This plasma is formed by the application of a high voltage

spark to a tangential flow of argon gas, which causes electrons to be stripped from their

argon atoms. These electrons are caught up and accelerated into a magnetic field,

formed by a radio frequency (RF) energy which is applied on a RF coil surrounding the

plasma torch. This process causes a chain reaction of collision-induced ionisation leading

to an ICP discharge. The ICP reaches temperatures of 6,000-8,000°K. As the aerosol

transits the plasma, the droplets undergo numerous processes which include

desolvation, dissociation, atomisation, and ionisation [30]. Ions produced by the argon

ICP are principally atomic and singly charged, making it an ideal source for atomic MS.

Since the ICP works at atmospheric pressure and the MS requires a vacuum, an interface

typically consisting of a coaxial assembly of two cones (sampler and skimmer cone) and

a series of pressured differentials to allow efficient sampling of the atmospheric pressure

plasma gases while minimally perturbing the composition of the sample gases.

Chapter 1.2

24

After passing through the sampler and skimmer cones, several electrostatic lenses or ion

optics focus the ions into the MS, where the ions are separated based on their mass-to-

charge (m/z) ratios. Three main MS principles are used in ICP-MS systems: quadrupole,

magnetic sector, and time of flight (TOF). The quadrupole is the most commonly used

type in ICP-MS. It comprises of two pairs of parallel cylindrical rods. The voltages applied

to these rods give a dynamic hyperbolic electric field, in which any ion above or below

the set mass enters an unstable trajectory and is lost from the ion beam. By varying the

voltages applied to these rods, a full mass spectrum can be obtained [30]. While the

quadrupole MS is used in the majority of ICP-MS instruments, some systems utilise a

magnetic sector or high resolution (HR) analyser, typically employed when higher mass

resolution is required [31]. This analyser uses a magnetic field, which is dispersive with

respect to ion energy and mass and deflects different masses through different angles.

The ions subsequently enter an electrostatic analyser, which is dispersive with respect to

ion energy and focuses the ions to the detector. In a TOF MS [32], a uniform electrostatic

pulse is applied to all ions at the same time, causing them to be accelerated down a

flight tube. Because lighter ions achieve higher velocities and arrive at the detector

earlier than heavier elements, the arrival times of the ions are determined by their m/z

ratios.

After passing the MS, the ions strike the active surface of the detector, typically an

electron multiplier. The electron multiplier subsequently generates a cascade of

electrons or discrete pulses, which is proportional to the number of ions that initially

struck the front of the detector.

Figure 3. Schematic figure of an inductively coupled plasma mass spectrometer (obtained from Varian,

Mulgrave, Victoria, Australia)

ICP-MS in oncology

25

2.2 Interferences

One of the main limitations of ICP-MS is the appearance of interferences, which can be

classified into two major groups. The first group comprises the spectral interferences,

which arise from other elements (isobaric interferences), polyatomic ions (e.g. oxides), or

doubly charged ions with the same m/z ratio as the analyte isotope. Elemental isobaric

interferences can usually be avoided by choosing an interference free analyte isotope

when the analyte of interest is not monoisotopic. Alternatively, because of the constant

nature of isotope ratios for most of the naturally occurring elements, elemental isobaric

interferences can be easiliy corrected mathematically by monitoring the intensity of an

isotope of the interfering element which is free from spectral interferences [30]. For the

three most abundant Pt isotopes (194Pt: abundance 33.0%, 195Pt: 33.8%, 196Pt: 25.2%), only 196Pt is subject to an isobaric interference (196Hg). However, this interference can be

corrected online by monitoring 202Hg signals. For the most abundant isotopes of Ru (99Ru:

12.7%, 101Ru: 17.0%, 102Ru: 31.6%, 104Ru, 18.7%), 99Ru, 102Ru, and, 104Ru are subject to isobaric

interferences of respectively 99Tc, 102Pd, and 104Pd, which can also be corrected online.

Polyatomic or molecular interferences can be produced by the combination of two or

more atomic ions and are usually associated with either the argon plasma, atmospheric

gases, or matrix components of the solvent or the sample. These interferences can be

overcome by choosing an interference free isotope, removing the matrix [33], using

alternative sample introduction systems [33], using mathematical corrections equations

[34], employing cool plasma conditions [35], using a collision reaction cell [36], or by

using a high resolution mass analyser [37]. Elements with high masses, such as Pt, are

less susceptible to molecular interferences than lower masses, such as Ru [38,39].

However, metal oxide interferences, which can occur as a result of incomplete

dissociation of the sample matrix or from recombination within the plasma or the

interface, can interfere with the analysis of Pt and Ru. Pt isotopes may be subject to

interferences from hafnium oxides [40,41] and tungsten oxides [41]. Ru isotopes can be

subject to oxide interferences from krypton, bromine, selenium, strontium, and

rubidium. Oxide formation, though, can be minimised by optimising the gas flow rate,

pump rate, and ionisation conditions of the plasma. Since metal oxide formation is

typically controlled, via the plasma conditions, to be less than 2% and because hafnium

background concentrations in biological samples are typically lower than Pt

backgrounds [42], hafniumoxides will not interfere significantly with Pt signals [43]. Prior

to the development of an ICP-MS assay for Pt or Ru, however, background

concentrations of the elements of potentially interfering metal oxides in the biological

matrix should be investigated.

The last type of spectral interferences are the doubly charged ions, which are analysed at

half the mass of the element, since the mass spectrometer measures m/z ratios. Pt

isotopes are not susceptible to interference of doubly charged ions as no element with a

Chapter 1.2

26

mass two times the mass of Pt exists. Ru isotopes, however, might be interfered by

numerous doubly charged ions (e.g [Au]2+, [Pt]2+, [Hg]2+, and [Pb]2+). The formation of

doubly charged ions can, however, be minimised by optimising ICP-MS parameters such

as lens voltages and plasma conditions.

The second group of interferences are the non-spectral interferences which can be

broadly divided into two categories: first the physical signal suppression resulting from

(un)dissolved solids or organics present in the matrix. Matrix components may have an

impact on the droplet formation in the nebuliser or droplet size selection in the spray

chamber, which can affect the transport efficiency and thus the signal intensity [44]. In

the case of organic matrices, the viscosity of the sample that is aspirated is modified. In

addition, the solids present in the matrix might lead to a deposition of solids on the

cones and subsequently result in an altered ion transmission. Furthermore, undissolved

solids can clog the nebuliser and torch. A decrease in these physical effects is possible by

an adapted sample pretreatment (e.g. dilution), the use of proper calibration techniques

[30] preferably combined with the use of an internal standard (IS), or by adjustment of

the sample introduction system.

The second category of non-spectral interferences are the matrix interferences [44]

which are caused by changes in the loading of the plasma or space-charge effects and

result in signal alteration. An extensive loading of the plasma may effect the ionisation

efficiency of the analyte ions. High concentrations of easily ionisable matrix elements,

such as sodium, might result in a decreased ionisation efficiency of elements with higher

ionisation energies and thus a decreased signal of these elements. In general, the lower

the degree of ionisation of the analyte in the plasma, the greater the effect of a matrix

component on the ion count rate of the element will be.

Space-charge effects are frequently seen in the analysis of light elements. The

magnitude of signal suppression generally increases with decreasing atomic mass of the

analyte ion. This is the result of a poor transmission of ions through the ion optics due to

matrix induced space-charge effects. The high-mass matrix element will dominate the

ion beam and pushes lighter elements out of the way resulting in a suppression of the

signal.

It is difficult to measure and quantify non-spectral matrix interferences. Again,

separation of the analytes from the matrix or dilution of samples may reduce this type of

non-spectral interferences. Furthermore, internal standardisation may be successful in

reducing the interferences. The IS, however, must be closely matched in both mass and

ionisation energy because they are to behave equal to the analyte. Also, the use of

matrix matched calibration standards or standard addition might correct the matrix

interferences. Although the signal suppression of the analyte will be corrected by proper

calibration methods, the actual space-charge effects will not be solved. The most

common approach to reduce space-charge effects is to apply voltages to the individual

ICP-MS in oncology

27

ion lens components. This will steer the analyte ions through the mass analyser while

rejecting a maximum number of matrix ions [44].

2.3 Combination of ICP-MS detection with speciation techniques

ICP-MS can be used as a Pt or Ru specific detector for several speciation technologies.

ICP-MS has several advantages over other methods of detection including a wide linear

dynamic range, low detection limits, potential for isotope determinations, and multi-

element capability. Moreover, the signal intensities are independent of the chemical

structure of the analyte incorporating Pt or Ru and hence the method does not require

standards of each analyte/metabolite/adduct. ICP-MS can provide quantitative

information for structurally non-correlated metal compounds.

2.4 Method validation

Following development of an ICP-MS assay and before implementation into routine use,

the assay needs to be validated to demonstrate that it is suitable for its intended use.

Validation is required to ensure the performance of the method. As chromatography is

widely used in bioanalysis, validation guidelines have already been extensively

described for speciation methods [45]. In contrast, no such guidelines are available for

ICP-MS. This has led to some discrepancies concerning the definition of validation

parameters in literature describing ICP-MS based bioanalytical assays. No stringent

procedure is followed for the assessment of limit of detections (LODs), lower limit of

quantifications (LLOQs), precision, accuracy, and linearity in the field of ICP-MS. The LOD

and LLOQ for instance can be obtained by several approaches such as; signal-to-noise

ratios, the standard deviation of the noise, or the standard deviation of the noise and

slope of the calibration curve [46]. For reported ICP-MS assays it is not always defined

which approach has been used. Furthermore, the LOD, LLOQ, and calibration range are

reported either in the processed sample matrix (the final matrix entering the ICP-MS) or

in the unprocessed sample matrix. The difficulty is, that the matrix in question is not

always clearly defined. Another intricacy is that concentrations of compounds are

commonly reported in weight per volume (w/v) instead of molar concentrations

(moles/v). In case of an elemental detection technique like ICP-MS, it therefore is pivotal

to report whether the metal or the metal-containing compound is used for calculation of

the concentrations. Unfortunately, this is not always clear from the reported data.

Because of these issues it is difficult to compare assays based on their detection limits

and other validation parameters.

In our opinion, procedures followed in e.g. the FDA guidelines could, as far as applicable

for ICP-MS, serve as an example for the development of a guideline for the validation of

Chapter 1.2

28

ICP-MS assays in biological matrices [45]. Validation parameters could include

assessment of the LLOQ, carry-over, linearity, specificity, accuracy, precision,

crossanalyte/IS interference, and stability.

3 Analytical ICP-MS assays: total metal determination

3.1 Application assays

3.1.1 Metal-based anticancer agents in biological fluids/cells

After an intravenous infusion, metal-based anticancer compounds form a variety of

hydrolysed intermediates in the blood [47,48]. These reactive species become rapidly

partitioned into plasma protein-bound metal, free plasma metal, tissue metal, white

blood cell metal, and erythrocyte-sequestered metal. The free metal fraction is generally

considered as the pharmacologically active metal fraction [49,50]. Because of the rapid

biotransformation and reactivity of the biotransformation products, investigation of the

pharmacokinetics of the intact parent compounds or the metabolites is technically

difficult and not feasible in routine analysis. Consequently, the assessment of total Pt or

Ru, rather than the analysis of the parent compound (e.g. cisplatin) and its metabolites

(e.g. aquated cisplatin), is a generally accepted approach for the analysis of the

pharmacokinetics of metal-based anticancer agents [51,52] in different biological

matrices. The analysis of the total metal content by an elemental technique such as ICP-

MS gives insight into the distribution of the drug irrespective of the molecular

composition of the drug and its metabolites. Since the first application of ICP-MS for an

oncology research question in 1990 [29], ICP-MS has become an accepted and

commonly used technique for the analysis of Pt anticancer agents. Table 1 summarises

the literature in which ICP-MS is used as the analytical technique to analyse total Pt and

Ru in biological fluids and tissue. Biological fluids predominantly studied are plasma or

serum, which contain the protein-bound and free metal fraction, ultrafiltered plasma

(pUF), ultracentrifuged plasma (pUC), or protein precipitated plasma (pP), which contain

the free metal fraction, and urine which contains metal eliminated by the kidney. The

tissues that are primarily studied in addition to tumour cells include renal and nerve

tissue, which are of interest due to the renal and neurotoxicity of Pt agents. The

capability of ICP-MS to measure ultra-trace Pt levels, allows the evaluation of long-term

Pt retention after treatment with Pt agents [53,53-56] as well as the determination of Pt

levels in small amounts of tissue samples.

Tabl

e 1:

Tot

al m

etal

det

erm

inat

ion

in b

iolo

gica

l flu

ids

and

cells

(NS

= no

t spe

cifie

d, N

A =

not

app

licab

le)

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Sam

plin

g

per

iod

V

alid

atio

n

des

crib

ed

Rem

arks

R

ef.

Plas

ma

Red

blo

od

cel

ls

Tiss

ue

Bo

ne

Uri

ne

Hu

man

Rat

Cis

pla

tin

A

cid

dig

esti

on

: dilu

tio

n

Aci

d d

iges

tio

n: d

iluti

on

Aci

d d

iges

tio

n: d

iluti

on

Aci

d d

iges

tio

n: d

iluti

on

Dilu

tio

n

194 Pt

19

5 Pt

196 Pt

Inte

rnal

sta

nd

ard

isat

ion

11

5 In

197 A

u

VG

Iso

top

es

Plas

maq

uad

PQ

1 U

p t

o 3

wee

ks

po

st

trea

tmen

t

No

C

om

par

iso

n to

GF-

AA

S: m

eth

od

s w

ere

in g

oo

d a

gre

emen

t

[29]

Blo

od

H

um

an

Cis

pla

tin

Ino

rgan

ic

pla

tin

um

Aci

d d

iges

tio

n: d

iluti

on

195 Pt

Exte

rnal

cal

ibra

tio

n in

sa

line

Inte

rnal

sta

nd

ard

isat

ion

197 A

u

VG

Iso

top

es

Plas

maq

uad

PQ

2

NS

Yes

C

om

par

iso

n to

vo

ltam

met

ry:

met

ho

ds

wer

e in

g

oo

d a

gre

emen

t

Co

mp

aris

on

wet

as

hin

g-d

ry a

shin

g

[9]

Seru

m

Cer

ebro

spin

al

fluid

Hu

man

C

isp

lati

n

1:20

dilu

tio

n

195 Pt

Exte

rnal

cal

ibra

tio

n in

1%

H

NO

3

Inte

rnal

sta

nd

ard

isat

ion

197 A

u

Perk

in-E

lmer

Sci

ex

ELA

N 5

000

Cro

ss fl

ow

neb

ulis

er

NA

N

o

[8

3]

Plas

ma

pU

C

Hu

man

C

isp

lati

n

1:20

dilu

tio

n

1:7

dilu

tio

n

194 Pt

Exte

rnal

cal

ibra

tio

n in

ar

tific

ial p

lasm

a

Inte

rnal

sta

nd

ard

isat

ion

153 Eu

Ner

mag

ICP-

MS

Co

nce

ntr

ic n

ebu

liser

NA

Y

es

[8

4]

Plas

ma

Red

blo

od

cel

ls

Uri

ne

Tiss

ue

Hu

man

Cis

pla

tin

Car

bo

pla

tin

Aci

d d

iges

tio

n: d

iluti

on

N

S N

S N

S V

G P

lasm

aqu

ad P

Q1

25 m

on

ths

po

st

trea

tmen

t

No

[53]

Plas

ma

Tiss

ue

Rats

Mic

e

Pig

s

Cis

pla

tin

A

cid

dig

esti

on

: dilu

tio

n

NS

Inte

rnal

sta

nd

ard

isat

ion

11

5 In

VG

Pla

smaq

uad

PQ

1 84

day

s p

ost

tr

eatm

ent

No

[96]

Kid

ney

Pi

gs

Cis

pla

tin

N

S N

S N

S N

S N

S 48

wee

ks p

ost

tr

eatm

ent

No

[54]

Inte

rver

teb

ral

dis

cs a

nd

ve

rteb

rae

Hu

man

C

isp

lati

n

Aci

d d

iges

tio

n: d

iluti

on

19

5 Pt

NS

NS

Shim

adzu

PIM

S-30

00

NS

No

[97]

Cel

l lin

es

Mic

e C

isp

lati

n

Aci

d d

iges

tio

n: d

iluti

on

NS

Stan

dar

d a

dd

itio

n

NS

Perk

in-E

lmer

Sci

ex

ELA

N 2

50

Co

nce

ntr

ic n

ebu

liser

NA

Y

es

[1

02]

Tabl

e 1.

Con

tinue

d

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Sam

plin

g

per

iod

V

alid

atio

n

des

crib

ed

Rem

arks

R

ef.

Inte

rver

teb

ral

dis

cs a

nd

ve

rteb

rae

Hu

man

C

isp

lati

n

Aci

d d

iges

tio

n: d

iluti

on

19

5 Pt

Exte

rnal

cal

ibra

tio

n

NS

Shim

adzu

PIM

S-30

00

NS

No

C

om

par

iso

n w

ith

IC

P-A

ES a

nd

AA

S [9

8]

pU

F

Hu

man

C

isp

lati

n

1:4

dilu

tio

n

195 Pt

N

S N

S Pe

rkin

-Elm

er S

ciex

EL

AN

Cro

ss fl

ow

neb

ulis

er

3 d

ays

afte

r st

op

infu

sio

n

No

[78]

Plas

ma

pU

C

Hu

man

C

isp

lati

n

1:20

dilu

tio

n

1:7

dilu

tio

n

194 Pt

Inte

rnal

sta

nd

ard

isat

ion

15

3 Eu

Ner

mag

Co

nce

ntr

ic n

ebu

liser

Up

to 1

8 d

ays

po

st d

ose

Y

es

[8

5]

Plas

ma

pU

C

Red

blo

od

cel

ls

Hu

man

O

xalip

lati

n

1:20

dilu

tio

n

1:10

dilu

tio

n

1:20

dilu

tio

n

NS

Exte

rnal

cal

ibra

tio

n in

sa

line

Inte

rnal

sta

nd

ard

isat

ion

153 Eu

Perk

in-E

lmer

Sci

ex

ELA

N 5

000

Up

to

3 w

eeks

p

ost

do

se

Yes

[86]

Plas

ma

pU

F

PUC

Hu

man

Oxa

lipla

tin

1:20

dilu

tio

n

1:10

dilu

tio

n

1:10

dilu

tio

n

NS

Exte

rnal

cal

ibra

tio

n in

sa

line

Inte

rnal

sta

nd

ard

isat

ion

153 Eu

Pe

rkin

-Elm

er S

ciex

EL

AN

500

0

21 d

ays

po

st

trea

tmen

t N

o

[1

12]

pU

F

Hu

man

N

S 1:

4 d

iluti

on

N

S St

and

ard

ad

dit

ion

N

S N

S N

S N

o

[1

49]

Do

rsal

roo

t g

ang

lia

Rats

O

xalip

lati

n

Cis

pla

tin

Orm

apla

tin

Aci

d d

iges

tio

n: d

iluti

on

N

S In

tern

al s

tan

dar

dis

atio

n

Ir

Bi

VG

OQ

-XR

ICP-

MS

con

cen

trat

ic n

ebu

liser

8 w

eeks

po

st

trea

tmen

t N

o

[1

50]

Plas

ma

pU

F

Uri

ne

NS

Satr

apla

tin

1:

10 to

1:1

00 d

iluti

on

Inte

rnal

sta

nd

ard

isat

ion

19

3 Ir

Perk

in-E

lmer

Sci

ex

ELA

N 5

000

14 d

ays

po

st

trea

tmen

t N

o

[8

0]

Lun

g c

ance

r cel

l lin

es

Cel

l lin

es

Cis

pla

tin

Vac

uu

m o

ven

dig

esti

on

: dilu

tio

n

NS

Exte

rnal

cal

ibra

tio

n in

1%

H

NO

3 N

S Se

iko

Inst

rum

ents

SP

Q65

00

1-2

h a

fter

ex

po

sure

N

o

[1

03,1

04]

pU

F

pP

Hu

man

C

isp

lati

n

1:16

7 d

iluti

on

1:50

dilu

tio

n

194 Pt

Inte

rnal

sta

nd

ard

isat

ion

11

5 In

203 Tl

Fiso

ns

Elem

enta

l VG

PQ

2+

NS

No

[93]

Ner

ve t

issu

e e.

g.

do

rsal

roo

t g

ang

lia

Live

r

Rat

Cis

pla

tin

A

cid

dig

esti

on

: dilu

tio

n

195 Pt

Exte

rnal

cal

ibra

tio

n in

1%

H

NO

3

NA

H

ewle

tt P

acka

rd H

P 45

00

V-g

roo

ve n

ebu

liser

Y

es

[1

51]

Tabl

e 1.

Con

tinue

d

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Sam

plin

g

per

iod

V

alid

atio

n

des

crib

ed

Rem

arks

R

ef.

Do

rsal

roo

t g

ang

lia

Rat

Cis

pla

tin

A

cid

dig

esti

on

: dilu

tio

n

195 Pt

Exte

rnal

cal

ibra

tio

n in

7%

H

NO

3/3%

H2O

2

Inte

rnal

sta

nd

ard

isat

ion

115 In

19

3 Ir

209 Bi

VG

Pla

smaQ

uad

(PQ

)-X

R

Flo

w in

ject

ion

an

alys

is

and

co

nti

nu

ou

s n

ebu

lisat

ion

.

8 w

eeks

po

st

do

se

Yes

[55]

Blo

od

Plas

ma

pU

F

Hu

man

O

xalip

lati

n

Aci

d d

iges

tio

n: d

iluti

on

Aci

d d

iges

tio

n: d

iluti

on

1:5

or 1

:18

dilu

tio

n

195 Pt

Exte

rnal

cal

ibra

tio

n in

m

atri

x

Inte

rnal

sta

nd

ard

isat

ion

193 Ir

Fin

nig

an M

AT

SOLA

IC

P-M

S

Co

nce

ntr

ic n

ebu

liser

Ult

raso

nic

neb

ulis

er

3 w

eeks

po

st

do

se

Yes

[87]

Blo

od

Plas

ma

Bra

in t

issu

e

Rats

C

isp

lati

n

Car

bo

pla

tin

Oxa

lipla

tin

1:24

dilu

tio

n

1:24

dilu

tio

n

Aci

d d

iges

tio

n: d

iluti

on

195 Pt

Exte

rnal

cal

ibra

tio

n m

atri

x fr

ee

NS

Hew

lett

Pac

kard

450

0

V g

roo

ve n

ebu

liser

NS

No

[151

]

0.9

% N

aCl

solu

tio

n

n-o

ctan

ol

solu

tio

n

Plas

ma

pP

Bra

in t

issu

e e.

g.

do

rsal

roo

t g

ang

lia

Rats

C

isp

lati

n

Oxa

lipat

in

Car

bo

pla

tin

Satr

apla

tin

Orm

apla

tin

NS

NS

1:24

dilu

tio

n

1:10

dilu

tio

n

Aci

d d

iges

tio

n: d

iluti

on

195 Pt

Ex

tern

al c

alib

rati

on

in 1

0%

HC

l N

S

Hew

lett

Pac

kard

450

0

56 d

ays

po

st

trea

tmen

t N

o

[5

6]

Ova

rian

an

d

mel

ano

ma

can

cer c

ells

Cel

l lin

es

Cis

pla

tin

Oxa

lipla

tin

Aci

d d

iges

tio

n

194 Pt

19

5 Pt

sum

med

Exte

rnal

cal

ibra

tio

n

Inte

rnal

sta

nd

ard

isat

ion

103 Rh

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

NS

No

[69]

Plas

ma

pU

F

Uri

ne

Tiss

ue

Bird

s C

isp

lati

n

1:50

dilu

tio

n

1:10

0 d

iluti

on

1:50

dilu

tio

n

Aci

d d

iges

tio

n: d

iluti

on

195 Pt

Ex

tern

al c

alib

rati

on

in

mat

rix

Inte

rnal

sta

nd

ard

isat

ion

115 In

V

G E

lem

enta

l VG

Pl

asm

a Q

uad

Co

nce

ntr

ic n

ebu

liser

96 h

aft

er

star

t in

fusi

on

N

o

[1

52]

Plas

ma

pU

F

Uri

ne

Hu

man

BB

R346

4 1:

10 to

1:1

00 d

iluti

on

N

S Ex

tern

al c

alib

rati

on

in

mat

rix

N

S N

S 10

day

s af

ter

infu

sio

n

No

[81]

Tabl

e 1.

Con

tinue

d

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Sam

plin

g

per

iod

V

alid

atio

n

des

crib

ed

Rem

arks

R

ef.

Plas

ma

pU

F

Uri

ne

Hu

man

O

xalip

lati

n

1:10

0 d

iluti

on

1:30

dilu

tio

n

1:20

0 d

iluti

on

194 Pt

19

5 Pt

196 Pt

Exte

rnal

cal

ibra

tio

n

NS

Perk

in-E

lmer

ELA

N 6

000

Cro

ss fl

ow

neb

ulis

er

4 d

ays

po

st

do

se

No

[64]

Blo

od

Plas

ma

pP

pU

F

cyto

sol

mem

bra

ne

pre

par

atio

ns

Hu

man

Sa

trap

lati

n

1:25

dilu

tio

n

194 Pt

19

5 Pt

Exte

rnal

cal

ibra

tio

n

NS

Hew

lett

Pac

kard

450

0

V-g

roo

ve n

ebu

liser

NS

No

[82]

Plas

ma

pU

F

Uri

ne

Red

blo

od

cel

ls

Tiss

ue

Rat

Oxa

lipla

tin

1:

250

dilu

tio

n

1:60

dilu

tio

n

1:40

0 d

iluti

on

Aci

d d

iges

tio

n: d

iluti

on

Aci

d d

iges

tio

n: d

iluti

on

NS

Inte

rnal

sta

nd

ard

isat

ion

Ir

Pe

rkin

-Elm

er S

ciex

EL

AN

600

0

NS

Yes

[79]

Lun

g c

ance

r ce

lls

Cel

l lin

es

Cis

pla

tin

A

cid

dig

esti

on

: dilu

tio

n

NS

Exte

rnal

cal

ibra

tio

n

NS

Fiso

ns

Plas

ma

Qu

ad 2

tu

rbo

N

S N

o

[9

9]

Uri

ne

Seru

m

Lun

gs

Mic

rod

ialy

sate

s o

f tu

mo

ur t

issu

e

Hu

man

C

arb

op

lati

n

Aci

d d

iges

tio

n: d

iluti

on

Mir

cow

ave

dig

esti

on

: dilu

tio

n

Mir

cow

ave

dig

esti

on

: dilu

tio

n

1:54

dilu

tio

n

194 Pt

19

6 Pt

enri

ched

Exte

rnal

cal

ibra

tio

n in

w

ater

Inte

rnal

sta

nd

ard

isat

ion

IDM

S

115 In

18

7 Re

Elem

ent 1

Hig

h

reso

luti

on

ICP-

SFM

S

Mic

roco

nce

ntr

ic

neb

ulis

er

Ult

raso

nic

neb

ulis

er

4 h

po

st d

ose

[4

3]

pP

Do

rsal

roo

t g

ang

lia

Rats

O

xalip

lati

n

1:10

dilu

tio

n

Aci

d d

iges

tio

n: d

iluti

on

NS

Exte

rnal

cal

ibra

tio

n

NS

Hew

lett

Pac

kard

450

0

24 h

po

st

trea

tmen

t N

o

[1

53]

pP

Tum

ou

r tis

sue

Mic

e

Car

bo

pla

tin

D

iluti

on

Aci

d d

iges

tio

n: d

iluti

on

195 Pt

Ex

tern

al c

alib

rati

on

N

S A

gile

nt 4

500

V-g

roo

ve n

ebu

liser

NS

No

[92]

Ren

al tu

bu

lar

cells

Ra

bb

its

Cis

pla

tin

A

cid

dig

esti

on

N

S In

tern

al s

tan

dar

dis

atio

n

113 In

N

S N

S N

o

[1

54]

Blo

od

Uri

ne

Bile

Hu

man

C

isp

lati

n

1:20

dilu

tio

n

NS

NS

195 Pt

Ex

tern

al c

alib

rati

on

Inte

rnal

sta

nd

ard

isat

ion

103 Rh

A

gile

nt 7

500i

2

wee

ks p

ost

tr

eatm

ent

Yes

[155

]

Tabl

e 1.

Con

tinue

d

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Sam

plin

g

per

iod

V

alid

atio

n

des

crib

ed

Rem

arks

R

ef.

Plas

ma

pU

F

Tiss

ue

of 1

1 o

rgan

s

Bird

C

arb

op

lati

n

NS

NS

NS

NS

NS

96 h

aft

er s

tart

in

fusi

on

N

o

[1

56]

pU

F

NS

Cis

pla

tin

N

S N

S N

S N

S N

S 2

h p

ost

tr

eatm

ent

No

[157

]

Bre

ast

can

cer

cells

C

ell

lines

C

isp

lati

n

Oxa

lipla

tin

Car

bo

pla

tin

Plat

inu

m

com

ple

xes

Aci

d d

iges

tio

n

195 Pt

In

tern

al s

tan

dar

dis

atio

n

113 In

Th

erm

o O

pte

k X

5 Se

ries

48

h a

fter

sta

rt

incu

bat

ion

sa

mp

ling

No

[100

]

Plas

ma

pU

F

Uri

ne

Hu

man

Sa

trap

lati

n

1:30

dilu

tio

n

1:30

dilu

tio

n

1:50

dilu

tio

n

195 Pt

Exte

rnal

cal

ibra

tio

n in

m

atri

x

Inte

rnal

sta

nd

ard

isat

ion

193 Ir

Pe

rkin

-Elm

er S

ciex

EL

AN

500

0

14 d

ays

po

st

do

se

Yes

[90]

Do

rsal

roo

t g

ang

lia

Rats

C

isp

lati

n

Aci

d d

iges

tio

n

195 Pt

In

tern

al s

tan

dar

dis

atio

n

102 Rh

Pe

rkin

-Elm

er S

ciex

EL

AN

600

0 N

S N

o

[6

6]

Seru

m m

ice

Live

r, ki

dn

ey

Cel

ls

Mic

e C

isp

lati

n

2 p

oly

mer

s w

ith

cis

pla

tin

NS

NS

NS

NS

NS

15 m

in a

fter

ad

min

stra

tio

n

No

[158

]

Cel

l su

rfac

e

Cel

ls

Cu

ltu

re m

ediu

m

Jurk

at

cell

lines

Cis

pla

tin

an

d

cisp

lati

n

carb

on

ato

co

mp

lex

Aci

d d

iges

tio

n: d

iluti

on

NS

NS

NS

Perk

in-E

lmer

Sci

ex

ELA

N61

00

NA

N

o

[6

2]

Seru

m

pU

F

Peri

ton

eal l

iqu

id

Ova

rian

can

cer

cells

Hu

man

C

isp

lati

n

Car

bo

pla

tin

1:10

0 d

iluti

on

19

5 Pt

Exte

rnal

cal

ibra

tio

n in

m

atri

x

209 Bi

H

ewle

tt P

acka

rd 4

500

NA

Y

es

[1

59]

Mel

ano

ma

cells

M

ice

AP5

346

Aci

d d

iges

tio

n: d

iluti

on

N

S Ex

tern

al c

alib

rati

on

in 0

.5%

Tr

ito

n-X

in w

ater

Inte

rnal

sta

nd

ard

isat

ion

115 In

Th

erm

o F

inn

igan

El

emen

t 2 IC

P-M

S N

S N

o

[6

7]

Plas

ma

Tiss

ue

Rats

N

S N

S N

S Ex

tern

al c

alib

rati

on

in

mat

rix

NS

NS

7 d

ays

po

st

trea

tmen

t N

o

[1

60]

Tabl

e 1.

Con

tinue

d

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Sam

plin

g

per

iod

V

alid

atio

n

des

crib

ed

Rem

arks

R

ef.

pU

F

Hu

man

C

isp

lati

n

Oxa

lipla

tin

Car

bo

pla

tin

1:10

or 1

:100

dilu

tio

n

194 Pt

Exte

rnal

cal

ibra

tio

n in

m

atri

x

Inte

rnal

sta

nd

ard

isat

ion

191 Ir

V

aria

n 8

10-M

S

Mic

ro n

ebu

liser

3 w

eeks

po

st

do

se

Yes

[6]

Cel

l fra

ctio

ns

Cel

l lin

es

Cis

pla

tin

N

S 19

5 Pt

194 Pt

en

rich

ed

Exte

rnal

cal

ibra

tio

n in

m

atri

x

IDM

S

NS

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

Flo

w in

ject

ion

an

alys

is

Pare

llel p

ath

neb

ulis

er

Mic

ron

ebu

liser

NA

Y

es

[7

5]

Cel

ls

Cel

l lin

e C

isp

lati

n

Satr

apla

tin

Aci

d d

iges

tio

n: d

iluti

on

NS

Inte

rnal

sta

nd

ard

isat

ion

11

3 In

Ther

mo

-Fin

nig

an

ELEM

ENT

II N

A

No

[101

]

Kid

ney

M

ice

Cis

pla

tin

Ti

ssu

e w

as s

liced

N

S Ex

tern

al c

alib

rati

on

in

mat

rix

NS

Elem

ent t

her

mo

el

ectr

on

ICP-

SFM

S

Co

up

led

to la

ser

abla

tio

n s

yste

m

1 h

aft

er

inje

ctio

n

Yes

[105

]

Plas

ma

pU

F

Uri

ne

Hu

man

N

AM

I-A

1:

10 to

1:1

00 d

iluti

on

10

1 Ru

Exte

rnal

cal

ibra

tio

n in

m

atri

x

Inte

rnal

sta

nd

ard

isat

ion

98Y

24

h a

fter

sta

rt

infu

sio

n

Yes

[39]

Jurk

at c

ells

H

um

an

Car

bo

pla

tin

A

cid

dig

esti

on

NS

NS

NS

NS

NA

N

o

[1

61]

Seru

m

Lun

g ti

ssu

e

Lun

g tu

mo

ur

tiss

ue

Rats

C

arb

op

lati

n

Aci

d d

iges

tio

n: d

iluti

on

195 Pt

NS

159 Tb

A

gile

nt 7

500c

e 12

0 m

in a

fter

st

art i

nfu

sio

n

No

[162

]

ICP-MS in oncology

35

3.1.2 Metal-based anticancer agents bound to DNA

The mechanism of action of Pt compounds is still not completely understood. It is,

however, generally accepted that DNA platination is the ultimate event in the cytotoxic

activity of Pt anticancer agents. The hydrolysed products of the Pt compounds are

believed to primary attack the nucleophilic N7 positions from guanine (G) and adenine

(A) leading to the formation of monofunctional adducts and bifunctional intra- and

interstrand crosslinks [57,58] (Figure 4). The four major cisplatin-DNA adducts are: Pt-G

(monofunctionally bound cisplatin), Pt-GG (intrastrand crosslink on pGpG sequences),

Pt-AG (intrastrand crosslink on pApG sequences), and G-Pt-G (intrastrand crosslinks on

pG(pN)pG and interstrand crosslinks) [59,60]. Pt-GG and Pt-AG represent respectively 60-

65% and 20-25% of the total amount of adducts formed. platinum-DNA (Pt-DNA)

adducts affect the DNA replication and transcription and, thereby, inhibit tumour

growth. As a consequence, in addition to the analysis of Pt in biological fluids and cells,

the quantification of Pt-DNA adducts is of major interest. For cisplatin, only 1% of the Pt

molecules that enter the cells actually bind to nuclear DNA [61,62]. This issue illustrates

the need for sensitive techniques to quantify the level of Pt bound to DNA. The high

sensitivity of ICP-MS allows the determination of Pt-DNA adducts in a small number of

cells. Table 2 summarises literature in which ICP-MS was used for quantification of the

total amount of Pt-DNA adducts in peripheral blood mononuclear cells (PBMCs) or

tissues from patients [63-65] or rodents [66,67] after treatment with Pt agents. This Table

also summarises the quantification of Pt-DNA adducts in various cell types [62,66,68-75]

after in vitro incubation with Pt.

3.1.3 Metal-based anticancer agents in environmental samples

Another application of ICP-MS for the analysis of total Pt is the monitoring of personnel

working with Pt anticancer drugs and the monitoring of the contamination of

environments where these drugs are prepared and administered. Because Pt agents play

a major role in the treatment of cancer, large amounts of these agents are processed,

e.g. in hospital pharmacies. Considering the numerous publications regarding the

monitoring of the potential exposure of personnel, apparently, the potential health risks

for persons manipulating cytotoxic drugs are a concern. Another source of

contamination of the environment, which might effect the health of individuals is the

release of metal-based anticancer agents by hospitals into waste water. Considerable

portions of Pt drugs are eliminated via the patients urine [76] into the waste water. The

low concentrations present in biological samples from personnel and in environmental

samples such as surface wipes, air filters, and waste water, make ICP-MS to a commonly

used method for the quantification of Pt in these samples. Table 3 summarises the

literature published in this field.

Chapter 1.2

36

Figure 4 [171]. Cisplatin-DNA adducts (a) Pt-G (monofunctionally cisplatin to guanine (X may be the

original chloride, or a hydroxyl group)); (b) G-Pt-G (interstrand crosslink); (c) cisplatin guanine–protein

crosslink; (d) Pt-GG (intrastrand crosslink on pGpG-sequences); (e) G-Pt-G (intrastrand crosslink on

GpNpG sequences (N represents a base)); (f) Pt-AG (intrastrand crosslink on pApG-sequences).

60-65 %

1 % 10 %

3 %

20-25 %

Pt H3N

NH3

NH3

NH3

NH3

PtG

H3N

G

NH3

X

PtNH3

G

G

PtNH3 G

N

Pt

NH3

H3N G PtNH3 A

G

G

G

G

a.

b.

c.

d.

e.

f.

Tabl

e 2:

Det

erm

inat

ion

of to

tal a

mou

nt o

f met

al b

ound

to D

NA

(NS

= no

t spe

cifie

d)

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Tim

e sc

ale

Val

idat

ion

d

escr

ibed

R

emar

ks

Ref

.

DN

A fr

om

PB

MC

s H

um

an

Cis

pla

tin

Car

bo

pla

tin

Dilu

tio

n in

0.1

% T

rito

n in

wat

er

NS

Exte

rnal

cal

ibra

tio

n

NS

Perk

in-E

lmer

Sci

ex

ELA

N 5

000

Y

es

In v

ivo

and

in v

itro

sa

mp

les

[6

3]

DN

A fr

om

le

uke

mia

an

d

lun

g c

ance

r cel

ls

Cel

l lin

es

Cis

pla

tin

Dilu

tio

n in

0.6

M H

Cl a

nd

hea

tin

g

194 Pt

+19

5 Pt

Inte

rnal

sta

nd

ard

isat

ion

10

3 Rh

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

N

o

In v

itro

[6

8]

DN

A fr

om

o

vari

an a

nd

m

elan

om

a ca

nce

r cel

ls

Cel

l lin

es

Cis

pla

tin

Oxa

lipla

tin

NS

194 Pt

+19

5 Pt

Exte

rnal

cal

ibra

tio

n

Inte

rnal

sta

nd

ard

isat

ion

103 Rh

Pe

rkin

-Elm

er S

ciex

EL

AN

600

0

No

In

vit

ro

[69]

DN

A fr

om

lun

g

can

cer a

nd

ad

eno

carc

ino

ma

cells

Cel

l lin

es

Cis

pla

tin

1:

20 d

iluti

on

in 3

.5 %

HN

O3 a

nd

h

eati

ng

194 Pt

19

5 Pt

196 Pt

19

8 Pt

Exte

rnal

cal

ibra

tio

n

NS

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

Cro

ss fl

ow

neb

ulis

er

N

o

In v

itro

[7

0]

DN

A fr

om

wh

ite

blo

od

cel

ls

Hu

man

O

xalip

lati

n

NS

194 Pt

19

5 Pt

196 Pt

NS

NS

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

Cro

ss fl

ow

neb

ulis

er

3 d

ays

po

st

do

se

No

In

viv

o [6

4]

DN

A fr

om

wh

ite

blo

od

cel

ls

Hu

man

BB

R346

4 D

iluti

on

in

1.7

5% H

NO

3 an

d

hea

tin

g

194 Pt

19

5 Pt

196 Pt

19

8 Pt

Exte

rnal

cal

ibra

tio

n

Inte

rnal

sta

nd

ard

isat

ion

Tl

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

Cro

ss fl

ow

neb

ulis

er

Ther

mo

finn

igan

N

eptu

ne

pla

sma

ion

isat

ion

mu

lti-

colle

cto

r mas

s sp

ectr

om

eter

(PIM

MS)

Mic

ron

ebu

liser

No

In

vit

ro

Co

mp

are

Qu

adru

po

le IC

P an

d

PIM

MS

[71]

DN

A fr

om

bo

ne

mar

row

as

pir

ates

Hu

man

C

arb

op

lati

n

Dilu

tio

n in

3.5

% H

NO

3 an

d

hea

tin

g

Th

erm

oFi

nn

igan

N

eptu

ne

pla

sma

ion

isat

ion

m

ulit

colle

cto

r MS

2 d

ays

afte

r st

op

tr

eatm

ent

No

In

viv

o [6

5]

DN

A is

ola

ted

fr

om

cer

vica

l an

d c

olo

rect

al

can

cer c

ells

Cel

l lin

es

Cis

pla

tin

Oxa

lipla

tin

Aci

d d

iges

tio

n: d

iluti

on

Exte

rnal

cal

ibra

tio

n i

n

0.1M

HN

O3

Inte

rnal

sta

nd

ard

isat

ion

103 Rh

Pe

rkin

-Elm

er E

LAN

DRC

p

lus

Y

es

In v

itro

[7

2]

DN

A fr

om

do

rsal

ro

ot

gan

glia

Ra

ts

Cis

pla

tin

1:

25 d

iluti

on

in 2

% H

Cl

195 Pt

In

tern

al s

tan

dar

dis

atio

n

102 Rh

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

N

o

In v

itro

In v

ivo

[66]

Tabl

e 2:

Con

tinue

d

Sam

ple

Sp

ecie

s C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Tim

e sc

ale

Val

idat

ion

d

escr

ibed

R

emar

ks

Ref

.

DN

A fr

om

so

lid

tum

ou

r cel

ls

Mic

e A

P534

6 A

cid

dig

esti

on

: dilu

tio

n

NS

Exte

rnal

sta

nd

ard

s in

0.5

%

trit

on

-X

Inte

rnal

sta

nd

ard

isat

ion

115 In

Th

erm

o F

inn

igan

El

emen

t 2

N

o

In v

ivo

[67]

DN

A fr

om

do

rsal

ro

ot

gan

glia

Ra

t C

isp

lati

n

Oxa

lipla

tin

NS

195 Pt

In

tern

al s

tan

dar

dis

atio

n

102 Rh

Pe

rkin

-Elm

er S

ciex

EL

AN

600

0

N

o

In v

itro

[7

3]

DN

A fr

om

jurk

at

cells

Ju

rkat

ce

ll lin

es

Cis

pla

tin

an

d

cisp

lati

n

carb

on

ato

co

mp

lex

NS

NS

NS

N

S Pe

rkin

-Elm

er S

ciex

EL

AN

6100

No

In

vit

ro

[62]

DN

A fr

om

se

vera

l ca

rcin

om

a ce

lls

Cel

l lin

es

Seve

n

amm

ine/

pro

pyl

amin

e Pt

(II)

com

ple

xes

wit

h

carb

oxy

late

s

Son

icat

ion

N

S N

S N

S Pe

rkin

-Elm

er S

ciex

EL

AN

600

0

No

In

vit

ro

[74]

DN

A o

f ova

rian

an

d m

elan

oo

ma

can

cer c

ells

Cel

l lin

es

Cis

pla

tin

N

S 19

5 Pt

194 Pt

en

rich

ed

Exte

rnal

cal

ibra

tio

n in

1%

H

NO

3

IDM

S

NS

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

Flo

w in

ject

ion

an

alys

is

Pare

llal f

low

neb

ulis

er

Mic

ron

ebu

liser

Y

es

In v

itro

[7

5]

Tabl

e 3:

Det

erm

inat

ion

of to

tal a

mou

nt o

f met

al in

env

ironm

enta

l sam

ples

(NS

= no

t spe

cifie

d)

Sam

ple

C

om

po

un

d

Sam

ple

pre

par

atio

n

Iso

top

es

anal

ysed

C

alib

rati

on

tec

hn

iqu

e IS

Ty

pe

of i

nst

rum

ent

Val

idat

ion

des

crib

ed

Ref

.

Surf

ace

wip

es

Glo

ves

Uri

ne

Plat

inu

m

Extr

acti

on

wit

h w

ater

Extr

acti

on

wit

h w

ater

Dilu

tio

n

NS

Exte

rnal

cal

ibra

tio

n in

0.7

% H

NO

3

Inte

rnal

sta

nd

ard

isat

ion

Ir

Elec

tro

ther

mal

va

po

risa

tio

n IC

P-M

S Y

es

[88]

Uri

ne

Plat

inu

m

1:4

dilu

tio

n

NS

Inte

rnal

sta

nd

ard

isat

ion

19

3 Ir

Perk

in-E

lmer

Sci

ex E

LAN

50

00

Yes

[1

63]

Ou

tsid

e o

f via

ls

Surf

ace

wip

es

Glo

ves

Cis

pla

tin

Car

bo

pla

tin

Extr

acti

on

wit

h w

ater

194 Pt

19

5 Pt

Inte

rnal

sta

nd

ard

isat

ion

19

3 Ir

Perk

in-E

lmer

ELA

N 6

100

Dir

ect n

ebu

lisat

ion

No

[1

06]

Was

te w

ater

Solid

resi

du

es

Cis

pla

tin

Oxa

lipla

tin

Car

bo

pla

tin

Aci

difi

cati

on

Dig

esti

on

: dilu

tio

n

195 Pt

Ex

tern

al c

alib

rati

on

Inte

rnal

sta

nd

ard

isat

ion

115 In

Pe

rkin

-Elm

er S

ciex

ELA

N

DRC

II

No

[7

6]

Air

filt

ers

Surf

ace

wip

es

Glo

ves

Uri

ne

Cis

pla

tin

Car

bo

pla

tin

Extr

acti

on

wit

h w

ater

Extr

acti

on

wit

h w

ater

Extr

acti

on

wit

h w

ater

NS

194 Pt

19

5 Pt

Y

Filt

ers/

wip

es/g

love

s: IC

P-M

S

Uri

ne:

ele

ctro

ther

mal

va

po

risa

tio

n IC

P-M

S

Yes

[9

1]

Filt

ers

Cis

pla

tin

Ex

trac

tio

n w

ith

0.9

% N

aCl i

n n

-p

rop

ano

l/w

ater

(75:

25) v

/v)

NS

Exte

rnal

cal

ibra

tio

n w

ith

ex

trac

ted

filt

ers

NS

NS

Yes

[1

08]

Surf

ace

sam

ple

s C

isp

lati

n

Car

bo

pla

tin

Oxa

lipla

tin

Extr

acti

on

wit

h 1

% H

Cl

Son

icat

ion

194 Pt

Ex

tern

al c

alib

rati

on

wit

h

extr

acte

d t

issu

es

Inte

rnal

sta

nd

ard

isat

ion

191 Ir

V

aria

n 8

10-M

S Y

es

[107

]

Air

sam

ple

s

Oxa

lipla

tin

V

apo

ur w

as t

rap

ped

in w

ater

NS

NS

NS

Ther

mo

ele

men

tal λ

7

CC

T b

ench

to

p s

erie

wit

h

DRC

(Th

erm

o o

pte

k)

No

[1

64]

Chapter 1.2

40

3.2 Assay development

For ICP-MS analysis, biological samples cannot be analysed directly, but require a

pretreatment to reduce the matrix effects of endogenous compounds, such as cell

constituents, proteins, salts, and lipids. The development of ICP-MS methods for metals

in biological matrices is generally focused on the selection of an appropriate sample

pretreatment and the selection of calibration procedures to avoid and compensate matrix

effects. Additionally, instrumental modifications can be used to further optimise the

assay. For the analysis of low concentrations of metal it is important to consider that,

whatever sample pretreatment procedure is used, special care has to be taken to avoid

contamination of samples. A careful selection of pretreatment devices and reagents

should be performed. Glassware should be avoided as it may contain considerable

amounts of Pt. Moreover, sample pretreatment needs to be performed in a dedicated

area to prevent environmental Pt or Ru originating from e.g. pollution by car exhaust

catalysts [77], from interfering with the analysis.

3.2.1 Sample pretreatment

3.2.1.1 Metal-based anticancer agents in biological fluids/cells

As was mentioned before, the direct determination of metals in biological matrices is

problematic. High protein contents can easily block the nebuliser and torch and deposit

on the cones and thereby affect the performance of the method. The most commonly

used pretreatment method for liquid biological samples is dilution. This method is

employed in order to lower the concentration of dissolved solids. For the analysis of Pt

and Ru in liquid biological samples, dilution with water [64,78-82], diluted HCl [29],

diluted HNO3 [6,39,83-92], and reagents such as a mixture of edta diammonium salt and

Triton X-100 in water [39,56,93] have been frequently used to reduce the solid content

below the for ICP-MS required 0.2% (g/v) [94]. Dilution factors are a compromise

between a minimum sample dilution to assure low quantification limits and a maximum

reduction of total dissolved solid content. A simple dilution, however, could be

problematic in samples with high protein contents such as whole blood, plasma, or

serum because the acids present in most ICP-MS diluents might precipitate proteins

which may clog up the nebuliser. Another possibility to reduce the matrix effects is by

protein precipitation. Using this method, however, protein-bound elements such as Pt,

will not be analysed which can be an undesired effect.

Although various studies used dilution as sample pretreatment procedure for protein

containing fluids without reporting precipitation problems, Tothill et al. showed that

addition of a small amount of acid to plasma which was diluted with water resulted in

protein precipitation [29]. Hence, instead of dilution, several applications based their

ICP-MS in oncology

41

pretreatment on acid digestion where after, prior to analysis, samples were diluted with

a proper diluent. In bioanalysis of metal-based anticancer drugs, there are several

methods commonly used for digestion. In general, the sample is heated with several

combinations of concentrated acids. Tothill et al. digested plasma by using 70% HNO3

(v/v) and heated the mixture at 100 °C until dryness [29]. Another study used a similar

procedure for the digestion of whole blood and plasma [87]. Nygren et al digested

samples by dry ashing and wet ashing [9]. For the first procedure, a combination of

concentrated HNO3 and HCl (aqua regia), and temperatures up to 800 °C were used to

obtain a dried and ashed sample. For wet ashing, the samples were heated with

concentrated acids (HNO3/HClO4) resulting in a solution of ashed sample and acid. The

methods were in agreement, indicating that no sample was lost due to volatility after

dry ashing. An open vessel wet ashing procedure (HNO3/H2O2/HCl) was used in another

study to digest urine [43]. HCl was used to provide chloro-complexes of Pt and thereby

reduce the memory effects in the sample introduction system that were observed when

diluting in 1% HNO3 (v/v). A more aggressive microwave digestion procedure was used

for serum using concentrated HNO3 in combination with elevated pressure and

temperature levels [95]. The authors used digestion with subsequent dilution instead of

a simple dilution because the total dilution factor was reduced by using the first method.

The pretreatment of tissue samples always includes digestion followed by dilution. The

digestion procedure can be performed in several ways, varying from heating at high

temperatures in concentrated HNO3 [29,53,56,62,66,67,92,96-101] to more complex

procedures such as digestion with concentrated HNO3 and 30% H2O2 [55,69] at elevated

temperatures or microwave digestion with HNO3, H2O2, and HCl [43]. Perry et al.

compared the digestion of cells with sodium hydroxide with the sonication of cells

followed by digestion with concentrated HNO3 [102]. Digestion with sodium hydroxide,

however, was not adequate because it resulted in significant signal suppression. The

addition of 0.9% (g/v) sodium chloride to the cell lines, with subsequent incubation of

the suspension in a vacuum oven at 120 °C also appeared to be an appropriate digestion

technique for cells [103,104], indicating that the addition of acid is not a requisite. The

matrix effect of the high level of dissolved sodium chloride was reduced by dilution with

water. An alternative for digestion of solid biological samples is the use of laser ablation

ICP-MS (LA-ICP-MS), which can directly analyse solid samples, requiring no further

sample pretreatment. Kidney sections from mice were analysed for Pt distribution using

this technique [105]. The advantage of this technique is that information can be

obtained on the distribution of metal among the tissue. The use of this technique,

however, requires the availability of an advanced laser ablation system, which is not

present in most clinical laboratories.

Chapter 1.2

42

3.2.1.2 Metal-based anticancer agents bound to DNA

The analysis of Pt in DNA samples is preceded by the isolation of DNA from PBMCs or

tissue samples. For this purpose, commercially DNA kits are available. However, instead

of using commercially available DNA kits, a standard isolation procedure can also be

used. In general, cells are washed with phosphate buffered saline, suspended in a buffer

containing Tris-HCl, NaCl, and edta disodium salt, and lysed with sodium dodecylsulfate

and protease. After removing the proteins, the DNA is ethanol or isopropanol

precipitated, where after it is dissolved in a water containing diluent. Subsequently, in

some applications, samples were diluted with triton in water [63] or with a HCl solution

[66] and analysed directly. Others diluted samples in a HCl [68] or HNO3 [65,70,71]

solution and heated it prior to analysis. A more extensive wet digestion procedure with

concentrated HNO3 and overnight heating was performed by Rice et al. [67]. Yamada et

al. digested DNA with concentrated HNO3 and 30% H2O2 [72]. Zhang et al. have described

an assay for which the DNA was sonicated prior to analysis [74]. To decide which

technique is suitable, experiments which assess the matrix effect and recovery of the

metals from the resulting solution should be performed. The lack of certified reference

compounds makes a proper validation difficult. Even though, Yamada et al.

demonstrated that matrix effects of various amounts of DNA did not effect Pt

quantification. None of the other studies described validation procedures.

3.2.1.3 Metal-based anticancer agents in environmental samples

Processing of environmental samples such as surface wipes is focused on the ability to

remove all the Pt present on the surface and on the efficiency to extract all the Pt from

the swab. In most assays, swabs were wetted with a diluent such as 30 mM NaOH

[88,91,106] or water [106,107] which was then used to wipe the surface. Subsequently, Pt

was extracted from the swab with water [88,91,106] or diluted HCl [107] under,

respectively, constant mixing or sonication. Extraction efficiencies of Pt from the swabs

were assessed in all studies and were higher than 87%. Only one study reported the

actual Pt recovery from the spiked surface, which exceeded 50% for all Pt compounds

tested [107]. The extraction of Pt from gloves was done using water [88,91,106], leading

to Pt recoveries of 67%. Air filters, which were used in a demistifier [108] and in a clean

room [91], were extracted with respectively 0.9% NaCl in a n-propanol water mixture

(75:25 v/v) and water.

ICP-MS in oncology

43

3.2.2 Calibration

In addition to an adequate pretreatment procedure, the calibration method should be

optimised during the development of a reliable ICP-MS assay. In general, calibration

procedures are focused on the compensation of matrix effects. As mentioned before, a

frequently applied method to circumvent matrix interference problems is internal

standardisation. An IS can be used to normalise the analyte signal and thereby correct

for matrix effects and instrumental signal drifts. It was shown that to correct

appropriately for non-spectral matrix effects, the mass number and ionisation potential

of the IS should be close to that of the analyte [35,109,110]. For the analysis of Pt,

internal standardisation is the most commonly used method to overcome matrix

interferences. Several IS are described for Pt in literature. In order of decreasing

popularity the following IS are used: iridium (191Ir or 193Ir, IP1; 9.1), indium (113In or 115In, IP1;

5.8), europium (153Eu, IP1; 5.7), bismuth (203Bi, IP1; 7.3), gold (197Au, IP1; 9.2), thallium (203Tl,

IP1; 6.1), rhodium (103Rh, IP1; 7.5), and rhenium (187Re, IP1; 7.9). Considering the mass and

first ionisation potential of Pt (195Pt, IP1; 9.0), iridium and gold seem to be the most

adequate ISs. Gold, however, tends to suffer from severe memory effects [83].

Furthermore, Tothill et al. encountered differences in chemistry between gold and Pt

[29]. These issues point out that gold is a less suitable IS than iridium. Even though the

resemblance between Pt and the IS is mentioned all over in literature, the frequent use

of indium suggests that stable results are achieved using this IS and that a deviating

mass and ionisation potential is not always critical. Ding et al. compared several ISs for

the quantification of Pt in cells [55]. They, however, concluded that indium did not

properly correct for matrix effects leading to a Pt recovery of 87-92% in the reliable

quantification range. Iridium and bismuth, in contrast, did show acceptable results (101-

108%). Hann et al. showed that rhenium and indium did not correct for matrix

suppression of Pt in digested urine [43]. The inconsistent results point out that, when

developing a new assay, it is advisable to test several ISs for each matrix. For the analysis

of Ru (IP: 7.4), yttrium (98Y, IP; 6.4) and germanium (72Ge, IP; 7.9) have been used as IS

[39,111]. They both corrected well for matrix effects.

Internal standardisation is often used in combination with external calibration,

preferably with matrix matched calibration standards. In addition to spiking the analyte

in the biological matrix [6,39,87,90,105] some authors use alternative matrices

containing the major components of the biological matrix, such as saline solutions

[9,86,112] and artificial plasma [84]. Others use calibrants in water [43] or in diluted acid

[55,56,72,83,103,104]. When significant matrix effects are observed in the samples, it is,

however, advisable to use matrix matched standards instead of solely water or diluted

acid. An alternative method for external standardisation is standard addition, whereby

the standard is added to the sample at multiple levels. The advantage is that the spiked

sample undergoes the same matrix effect as the analyte present in the unknown sample.

Chapter 1.2

44

The disadvantage is that a large amount of sample is needed and that the method is

time consuming. Perry et al. used this method for the determination of Pt in cell lines

[102]. Unfortunately, no validation results were shown. Yamada et al. obtained

equivalent results for external standardisation in diluted acid and standard addition for

the determination of DNA-bound Pt [72]. A last solution to overcome the matrix effects

is the isotope dilution mass spectrometry (IDMS) technique. With this technique, the

sample is spiked with an enriched isotope of the element of interest. The analysis of

isotope ratios in the unspiked and spiked sample, as well as in the spike itself, lead to the

quantification of the analyte. Hann et al. compared IDMS with enriched 196Pt to external

quantification with aqueous standards and standard addition for the analysis of three Pt

levels in urine [43]. Quantification with external calibrants revealed incorrect results for

all spiked levels, regardless whether internal standardisation with rhenium or indium

was used or not. This could be due to the aqueous matrix used for calibration that might

have resulted in different matrix effects of calibrants and samples. Standard addition was

only found to be suitable for samples spiked at higher levels, whereas IDMS resulted in a

correct quantification of all levels.

3.2.3 Instrumental adjustments

The majority of instruments used in the determination of the total amount of metals

originating from anticancer agents in biological matrices are quadrupole based

instruments. Few applications used double focussing ICP-MS for Pt analysis

[43,65,71,105]. Hann et al. examined that for digested serum and microdialysates, the

low and high resolution mode resulted in similar concentrations [43]. This indicates that

in these two matrices, no spectral interferes were present and that the use of the high

resolution mode in routine analysis was not required. John et al. compared a double

focussing ICP-MS with a quadrupole ICP-MS for the determination of Pt bound to DNA

[71]. The problem of oxide interference was minor for the quadrupole ICP-MS because

oxide formation was low. For the double focussing ICP-MS, however, oxide formation

was higher (>5%) leading to a higher contribution of hafnium interferences to the Pt

signal. The double focussing ICP-MS was more sensitive than the quadrupole

instrument. Analysis times for the first instrument, on the other hand, were significantly

longer.

In addition to the mass spectrometer, the type of sample introduction system can also

greatly affect the sensitivity of the instrument. Furthermore, it can determine the

amount of sample needed. Conventional pneumatic nebulisers (concentric or cross

flow), which are often used for Pt or Ru analysis [9,64,70,71,78,84,85,102] operate at a

flow rate of 0.5-1 ml/min [113]. The high flow rate in combination with the time required

to carry out a complete signal reading (1-5 min), requires a high sample volume (>1 mL).

ICP-MS in oncology

45

Obviously, this volume is much larger than the available sample volume in many clinical

applications. Therefore, undigested samples need to be diluted with a large volume of

diluent, which reduced the detection limit of the method. In addition to this limitation,

the transport efficiency of common sample introduction systems is low. With common

pneumatic nebulisers, only about 1% of the total sample volume is actually transported

into the plasma [30]. To surmount these issues, specially designed systems (micro

nebulisers) have been developed [75], which operate at low sample uptake rates (0.1-0.2

ml/min) and thereby only use a limited sample volume. In addition, ultrasonic nebulisers

have been developed, which have a high efficiency, independent of the gas flow. Thus

more analyte can be transported to the ICP resulting in lower detection limits. A 25-fold

improve of the detection limit for ultrasonic nebulisation (USN) compared to standard

concentric nebulisation was observed by Morrison et al for the determination of Pt in

pUF. USN was also used by Turci et al. [89]. Hann et al. tested two introduction systems,

USN and micro concentric nebulisation (MCN), for the determination of Pt in urine [43].

USN and MCN revealed signal intensities in the same range, probably because of

reduced nebulisation efficiency of the USN due to matrix effects of the digested urine

sample. LODs for USN were slightly better than for MCN. MCN based nebulisers were

also applied by Brouwers et al. and John et al. [6,39,71]. An alternative method for

continuous sample introduction is flow injection analysis (FIA) [55], in which a discrete

sample volume is injected into a continuously flowing carrier stream. A tenfold

improvement of the detection limit was observed compared to continuous nebulisation

because no sample dilution was needed using FIA. Another method to decrease the

required sample volume and increase the transport efficiency is electrothermal

vaporisation (ETV) [88,91]. The sample is deposited into an electrically conductive

vaporisation cell where the sample is dried and vaporised. An argon gas flow

subsequently carries the sample vapour to the plasma.

4 Analytical ICP-MS assays: speciation of metal-based anticancer

agents

4.1 Application assays

Because of the complex nature of Pt and Ru solution chemistry and the lack of suitable

certified reference materials to identify the species, often, no priority was given to the

speciation of the compounds in clinical samples. However, in addition to the

determination of total metal concentrations, the determination of the parent metal

compounds and their metabolites is valuable. Besides, the speciation of metal

compounds can be applied to investigate the metabolism of the compounds in the

Chapter 1.2

46

body and to characterise adducts of the metals with endogenous species, such as

proteins and DNA, to improve knowledge on the mechanism of action of metal-based

anticancer agents. Furthermore, it may be used to study the stability of metal-based

anticancer agents. In environmental samples, speciation can be practical to study the

composition of the molecules in e.g. surface samples and waste water. While most of the

work has involved the speciation of cisplatin and cisplatin-adducts, oxaliplatin,

carboplatin, satraplatin, ormaplatin, ZD0473, BBR3464, and the Ru compounds NAMI-A

and KP1019 have received attention too.

4.1.1 Speciation of metal-based compounds and metabolites

After administration of Pt drugs, Pt compounds rapidly form a variety of reactive

intermediates in the blood stream, including hydrolysed products. Speciation analysis of

these intermediates is gaining interest. Methods applying ICP-MS have been developed

for the speciation and quantification of the parent drugs and metabolites of cisplatin

[114,115], BBR3463 [116], satraplatin [82,117-119], oxaliplatin [120], and ZD0473 [26] in

matrices such as whole blood, red blood cells, plasma, pUF, pP, and urine. A summary of

the assays is presented in Table 4. In addition to the investigation and quantification of

the individual Pt agents and their metabolites, speciation analysis with ICP-MS detection

has been used to assess the hydrofobicity of several Pt compounds [56].

4.1.2 Speciation of reaction products of metal-based anticancer compounds with

DNA and proteins

As mentioned before, it is generally accepted that DNA platination is the ultimate event

in the cytotoxic mechanism of action of Pt anticancer agents. Speciation of the various

Pt-DNA adducts formed can be used to gain more insight into the exact cytotoxic

mechanism of Pt compounds. Pt-DNA adducts, however, are not the only reaction

products of Pt compounds that are interesting to study. Before Pt compounds can bind

to DNA, they must pass from the blood through the cytosol of the cell. Reactive Pt

complexes can bind to various constituents in the blood or in cells. Among the potential

Pt-binding sites are proteins and other compounds containing thiol donor ligands such

as cysteine, glutathione, and methionine [61,121]. The exact role that Pt binding to

proteins such as albumin and transferrin plays in the mechanism of drug action, remains

unclear. Interactions with proteins, however, might play an important role in drug

efficacy and side effects [122]. Methionine is important because of its large

concentrations and reactivity. There is evidence that the nephrotoxicity of cisplatin is

increased in the presence of its reaction products with methionine [123]. Glutathione is

believed to be involved in the cellular detoxification of Pt compounds. There is an

ICP-MS in oncology

47

indication that the polymorphisms of genes encoding glutathione transferase are

relevant for clinical response and development of toxicity [124-126]. Thus, the detection

and identification of reaction products of Pt-containing drugs with thiol compounds is

important in studies of toxicities.

The cytotoxic mechanism of Ru compounds is still largely unknown. In contrast to the

view that DNA is the main target of these agents, DNA-independent mechanism are also

suggested [4,127]. The binding of NAMI-A to DNA is far weaker than that of Pt complexes

[128]. Conversely, NAMI-A was shown to bind tightly to serum proteins [129,130],

suggesting that the binding of Ru compounds to plasma proteins is of utmost

importance for its cytotoxic effect. There is experimental evidence that the Ru moiety is

transferred into the tumour predominantly via the transferrin pathway [131]. Hence, the

investigation of the interaction of Ru compounds with serum proteins is relevant.

Research has utilised methods for speciation and ICP-MS quantification of

biotransformation and reaction products of Pt and Ru agents with DNA, proteins and

thiol compounds. These methods are summarised in Table 5.

4.1.3 Speciation of metal-based anticancer compounds in environmental samples

Environmental samples, such as hospital waste water may contain Pt agents and their

metabolites. For the determination of the toxicity of Pt in the environment, the

determination of total Pt is not sufficient. It is of interest to investigate the molecular

form in which Pt is present and, furthermore, to assess the ability of cleaning systems to

remove these molecules from the water circulation. Falter et al. investigated the

chromatographic behaviour of cisplatin and carboplatin in water in the presence of

several anions [132]. The results of this study are used for the development of an

extraction procedure for cisplatin and carboplatin from environmental matrices. Hann et

al. studied low concentrations of cisplatin [95,133], oxaliplatin [133], and carboplatin

[133] and their degradation products in water containing varying concentrations of

chloride and in human urine in order to support the development of elimination

procedures as well as toxicological studies [43,133]. Lenz et al. investigated the

metabolism and adsorption of Pt compounds in biological waste water [134]. These

methods are summarised in Table 6.

Tabl

e 4:

Spe

ciat

ion

of m

etal

-bas

ed c

ompo

unds

and

met

abol

ites

(NS

= no

t spe

cifie

d, N

A =

not

app

licab

le)

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

wra

te

Det

ecti

on

IS

+m

ob

ile p

has

e R

emar

ks

Ref

.

Salin

e so

luti

on

s C

isp

lati

n

Met

abo

lites

Pr

ior t

o s

pec

iati

on

: D

iluti

on

Io

n p

airi

ng

O

D5

C18

co

lum

n

150

x 4.

6mm

, 5µ

m

1 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 2

50

SDS

wit

h 3

% n

-p

rop

ano

lol p

H 2

.6

On

line

In v

itro

[114

]

Aq

ueo

us

solu

tio

n

Satr

apla

tin

Im

pu

riti

es a

nd

d

egra

dat

ion

p

rod

uct

s

Prio

r to

sp

ecia

tio

n:

1:10

dilu

tio

n

Reve

rsed

ph

ase

250

x 4.

6 m

m, 5

µ

m P

EEK

co

lum

n

pac

ked

wit

h

Hyp

ersi

l Ph

enyl

b

on

ded

sili

ca

1.3

ml/

min

Fi

son

s Pl

asm

aQu

ad 2

+

V-g

roo

ve

neb

ulis

er

Iso

crat

ic

1:3

v/v

acet

on

itri

le in

w

ater

On

line

Oxy

gen

ad

dit

ion

In v

itro

[117

]

pU

F

Satr

apla

tin

M

etab

olit

es

Prio

r to

sp

ecia

tio

n:

1:50

dilu

tio

n

Rece

rsed

ph

ase

PEEK

250

x 4

.6 m

m

pak

ced

wit

h

Hyp

ersi

l ph

enyl

5

µm

bo

nd

ed s

ilica

1 m

l/m

in

Fiso

ns

Plas

maQ

uad

2+

Co

nce

ntr

ic

neb

ulis

er

Gra

die

nt

15:8

5 (A

) to

90:

10

(B) a

ceto

nit

rile

: w

ater

On

line

Inte

rfac

e

In v

ivo

[118

]

pP

Sa

trap

lati

n

Met

abo

lites

N

S Re

vers

ed p

has

e Pr

od

igy

C8

150

x 4.

6 m

m

1 m

l/m

in

Hew

lett

Pac

kard

45

00

V-g

roo

ve

neb

ulis

er

Gra

die

nt

25%

met

han

ol-

0.01

%

ort

ho

ph

osp

oric

ac

id p

H 2

.5 (A

) 10

0% m

eth

ano

l (B

) (m

axim

al 2

0%

B)

On

line

In v

ivo

[119

]

Plas

ma

Red

blo

od

cel

ls

pU

F

Uri

ne

Oxa

lipla

tin

B

iotr

ansf

orm

atio

n

pro

du

cts

Pr

ior t

o s

pec

iati

on

: Pl

asm

a: d

iluti

on

Red

blo

od

cel

ls:

hem

oly

sis

pU

F: in

ject

ed

dir

ectl

y

Uri

ne:

dilu

tio

n

Reve

rsed

ph

ase

Hic

hro

m-C

IL

colu

mn

150

x

4.6m

m, 5

µm

1 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 5

000

Iso

crat

ic

15 m

M H

CO

OH

in

90:1

0 w

ater

/met

han

ol

On

line

In v

ivo

[120

]

Salin

e so

luti

on

C

isp

lati

n

Oxa

lipat

in

Car

bo

pla

tin

Satr

apla

tin

Orm

apla

tin

Hyd

rofo

bic

ity

of

pla

tin

um

co

mp

ou

nd

s

Prio

r to

sp

ecia

tio

n:

Inje

cted

dir

ectl

y Re

vers

ed p

has

e Pr

od

igy

C8

5 µ

m

1 m

l/m

in

Hew

lett

Pac

kard

45

00

Iso

crat

ic

100%

wat

er F

or

JM21

6 m

eth

ano

l w

as a

dd

ed

On

line

In v

itro

[56]

Salin

e an

d 0

.5%

HPM

C

solu

tio

n

Uri

ne

(do

g)

14C

lab

eled

Z

D04

73

Puri

ty fo

rmu

late

d

mat

eria

l

Met

abo

lites

Prio

r to

sp

ecia

tio

n:

Inje

cted

dir

ectl

y Re

vers

ed p

has

e In

erts

il O

DS-

3 15

0 m

m x

4.6

mm

5 µ

m

0.25

ml/

min

H

exap

ole

co

llisi

on

/rea

ctio

n

cell

ICP-

MS

Co

nce

ntr

ic

neb

ulis

er

Gra

die

nt

Trifl

uo

roac

etic

ac

id (A

) met

han

ol

(B) 2

-20%

B

On

line

In v

itro

In v

ivo

[165

]

Tabl

e 4:

Con

tinue

d

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

wra

te

Det

ecti

on

IS

+m

ob

ile p

has

e R

emar

ks

Ref

.

pP

pU

F 10

RBC

Satr

apla

tin

M

etab

olit

es

Stab

ility

Pr

ior t

o s

pec

iati

on

: In

ject

ed d

irec

tly

Prio

r to

ICP-

MS:

Fr

acti

on

s w

ere

dilu

ted

in 1

% H

NO

3

Reve

rsed

ph

ase

Pro

dig

y C

8 15

0 x

4.6

mm

N

A

Hew

lett

Pac

kard

45

00

V-g

roo

ve

neb

ulis

er

Gra

die

nt

25-4

0%

met

han

ol-0

.85%

p

ho

spor

ic a

cid

p

H 2

.5

Off

line

In v

itro

[82]

Plas

ma

(do

g)

ZD

0473

M

etab

olit

es

Prio

r to

sp

ecia

tio

n:

1:2

dilu

tio

n

Reve

rsed

ph

ase

Phen

om

enex

Sy

ner

gi P

ola

r RP

150

x 4.

6 m

m

1 m

l/m

in

Mic

rom

ass

Plat

form

ICP-

MS

wit

h h

eaxa

po

le

colli

sio

n/r

eact

ion

ce

ll

Ult

raso

nic

an

d

con

cen

tric

n

ebu

liser

Iso

crat

ic

met

han

ol/

wat

er

(20:

80 v

/v) w

ith

0.

1% fo

rmic

aci

d

and

0.1

5 m

M

amm

on

ium

ac

etat

e p

H 3

On

line

In v

ivo

[26]

pU

F

pP

BBR3

464

Imp

uri

ties

Met

abo

lites

NS

Cat

ion

exc

han

ge

Sup

elco

sil L

C-S

CX

25

0 x

4.6

mm

1

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

600

0 20

mM

pyr

idin

e in

w

ater

(A),

200

mM

pyr

idin

e in

w

ater

(B)

Gra

die

nt 0

-100

%

B

On

line

Ion

pai

rin

g n

o

succ

ess

In v

itro

[116

]

pU

F

UF

of c

ells

Cis

pla

tin

M

etab

olit

es

NS

Ion

pai

rin

g

µB

on

dap

ak C

18

colu

mn

300

x

3.9m

m, 5

µm

0.5

ml/

min

H

ewle

tt P

acka

rd

4500

IS

: 205 Tl

Iso

crat

ic

3% m

eth

ano

l, 0.

075

mM

SD

S,

pH

2.5

trifl

ic a

cid

On

line

In v

itro

[115

]

Tabl

e 5:

Spe

ciat

ion

of re

actio

n pr

oduc

ts o

f met

al-b

ased

ant

ican

cer c

ompo

unds

with

DN

A a

nd p

rote

ins

(NS

= no

t spe

cifie

d, N

A =

not

app

licab

le

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

wra

te

Det

ecti

on

IS

+m

ob

ile p

has

e R

emar

ks

Ref

.

Salin

e so

luti

on

C

isp

lati

n

Met

hio

nin

e,

glu

tath

ion

e,

cyst

ein

e b

ind

ing

Prio

r to

sp

ecia

tio

n:

Dilu

tio

n

Ion

pai

rin

g

OD

5 C

18 c

olu

mn

15

0 x

4.6m

m, 5

µm

1 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 2

50

10 m

M 1

-h

epta

nes

ulfo

nat

e 10

% m

eth

ano

l, 0.

1% fo

rmic

aci

d

(pH

2.6

)

On

line

In v

itro

[114

]

Cal

f th

ymu

s D

NA

C

isp

lati

n

Plat

inu

m b

ou

nd

to

AG

, GG

, an

d G

Prio

r to

sp

ecia

tio

n:

DN

A d

iges

tio

n w

ith

D

NA

se I

and

N

ucl

ease

P1

pro

tein

ase

K

Prio

r to

ICP-

MS:

Fr

acti

on

s d

ilute

d

An

ion

exc

han

ge

Mo

no

-Q a

nio

n

exch

ang

e 1

ml/

min

V

G P

lasm

aqu

ad

PQ1

IS:11

5 In

Gra

die

nt

12.5

mM

Tri

s (p

H

8.8)

(A),

12.5

mM

Tr

is, 1

M N

aCl (

pH

8.

8). (

5-25

%

bu

ffer

B)

Off

line

In v

itro

[146

]

Aq

ueo

us

solu

tio

n

Cis

pla

tin

M

eth

ion

ine

bin

din

g

NS

Reve

rsed

ph

ase

Sph

eris

orb

OD

S co

lum

n 1

5 cm

x

4.6

mm

, 5µ

m

1 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 5

000

Iso

crat

ic

15 m

M fo

rmic

ac

id in

wat

er

On

line

In v

itro

[123

]

Seru

m

Salin

e so

luti

on

Cis

pla

tin

Imid

azo

lium

tra

ns-

tetr

ach

loro

bis

(im

idaz

ole

) ru

then

ate

(III)

Sod

ium

tr

anst

etra

chlo

ro-b

is

(ind

azo

le)

ruth

enat

e (II

I)

Pro

tein

bin

din

g

Prio

r to

sp

ecia

tio

n:

1:10

0 d

iluti

on

Size

exc

lusi

on

Su

pel

co P

rog

el T

SK

4 cm

x 4

mm

, 6 µ

m

Sup

erd

ex75

-75

HR

10/3

0 SE

C 3

00 x

10

mm

,13

µm

Pro

gel

TSK

G30

00

PWX

L 30

cm x

7.8

m

m, 1

0 µ

m

0.9

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

600

0 30

mM

Tri

s-H

Cl

(pH

7 .2

) Iso

crat

ic

On

line

In v

itro

[142

]

Seru

m (r

abb

it)

pU

F

Pt(II

)Cl2-

4 Pr

ote

in b

ind

ing

Pr

ior t

o s

pec

iati

on

: D

irec

tly

inje

cted

Prio

r to

ICP-

MS:

Ex

trac

tio

n g

el w

ith

aq

ua

reg

ia

2-D

gel

el

ectr

op

ho

resi

s N

A

NA

El

emen

t HR-

ICP-

MS

Co

nce

ntr

ic

neb

ulis

er

IS:19

3 Ir

Off

lin

e

In v

itro

[148

]

Plas

ma

Red

blo

od

cel

ls

pU

F 30

kD

a

Uri

ne

Oxa

lipla

tin

Pr

ote

in b

ind

ing

N

S Si

ze e

xclu

sio

n

Sup

erd

ex 2

00 H

R 10

/30

0.4

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

500

0 50

mM

NaH

2PO

4 an

d 1

50 m

M N

aCl

On

line

In v

ivo

[120

]

Aq

ueo

us

solu

tio

n

Cis

pla

tin

G

luta

thio

ne,

cy

stei

ne

bin

din

g

NS

Reve

rsed

ph

ase

Hic

hro

m re

vers

ed

ph

ase

150

x 4.

6 m

m

1.0

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

500

0

15 m

M H

CO

OH

in

90/1

0 w

ater

/met

han

ol

On

line

In v

itro

[166

]

Tabl

e 5:

Con

tinue

d

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

wra

te

Det

ecti

on

IS

+m

ob

ile p

has

e R

emar

ks

Ref

.

Aq

ueo

us

solu

tio

n

Cis

pla

tin

Pl

atin

um

bo

un

d t

o

G

Prio

r to

sp

ecia

tio

n:

NS

Prio

r to

ICP-

MS:

O

nlin

e d

iluti

on

An

ion

exc

han

ge

Ion

pac

k A

S14

250

x 4

mm

wit

h

qu

ater

nar

y am

mo

niu

m

gro

up

s

1.2

ml/

min

IC

P-SF

MS

Fin

nig

an E

lem

ent

Co

nce

ntr

ic

neb

ulis

er

IS:11

5 In

Iso

crat

ic

wat

er+

17.5

mM

N

a 2CO

3+5m

M

NaH

CO

3+5%

C

H3C

N (p

H 1

0.7)

On

line

Split

of L

C

flow

. On

ly

72%

to IC

P-M

S w

ith

mak

e u

p fl

ow

In v

itro

[147

]

Gel

fro

m S

DS-

PAG

E Sa

trap

lati

n

Pro

tein

bin

din

g

Prio

r to

sp

ecia

tio

n:

1:3

dilu

tio

n

Prio

r to

ICP-

MS:

A

cid

dig

esti

on

d

iluti

on

Gel

ele

ctro

ph

ore

sis

SDS-

PAG

E N

A

Hew

lett

Pac

kard

45

00

V-g

roo

ve

neb

ulis

er

NA

O

fflin

e

In v

itro

[82]

DN

A

DN

A fr

om

cel

ls

Cis

pla

tin

Pl

atin

um

bo

un

d t

o

AG

, GG

, an

d G

Prio

r to

sp

ecia

tio

n:

DN

A d

iges

tio

n w

ith

D

NA

se I

and

n

ucl

ease

P1

Prio

r to

ICP-

MS:

Fr

acti

on

s d

ilute

d

An

ion

exc

han

ge

Mo

no

-Q H

R 5/

5 w

ith

ch

arg

ed

gro

up

of C

H2N

+

(CH

3) 3

1 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 6

000

Cro

ss-f

low

n

ebu

liser

Gra

die

nt

12.5

mM

Tri

s-H

Cl,

(pH

8.8

) (A

) an

d

12.5

mM

Tri

s-H

Cl,

pH

8.8

wit

h 1

M

NaC

l (B

)

Off

lin

e

In v

itro

[167

]

Aq

ueo

us

solu

tio

ns

Cis

pla

tin

H

aem

og

lob

in

bin

din

g

Prio

r to

sp

ecia

tio

n:

Dir

ectl

y in

ject

ed

Size

exc

lusi

on

Bi

oSE

p-S

EC 2

000

300

x 4.

6 m

m

0.8

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

610

0 D

RC p

lus

Iso

crat

ic

10 m

M

amm

on

ium

b

icar

bo

naa

t (p

H

7.4)

On

line

In v

itro

[139

]

Aq

ueo

us

solu

tio

ns

Cis

pla

tin

H

aem

og

lob

in

bin

din

g

Prio

r to

sp

ecia

tio

n:

Dir

ectl

y in

ject

ed

Size

exc

lusi

on

Bi

oSE

p-S

EC 2

000

300

x 4.

6 m

m

0.8

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

610

0 D

RC p

lus

Iso

crat

ic

5 m

M a

mm

on

ium

b

icar

bo

naa

t (p

H

7.4)

On

line

In v

itro

[168

]

Aq

ueo

us

solu

tio

ns

Cis

pla

tin

Ra

bb

it

met

allo

thio

nei

n

Prio

r to

sp

ecia

tio

n:

Dir

ectl

y in

ject

ed

Size

exc

lusi

on

Bi

oSE

p-S

EC 2

000

300

x 4.

6 m

m

0.8

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

610

0 D

RC p

lus

Iso

crat

ic

5 m

M s

od

ium

p

ho

sph

ate

(pH

7)

On

line

In v

itro

[169

]

Aq

ueo

us

solu

tio

ns

Red

blo

od

cel

ls

Oxa

lipla

tin

Cis

pla

tin

Car

bo

pla

tin

Hae

mo

glo

bin

b

ind

ing

Pr

ior t

o s

pec

iati

on

: A

qeo

us

solu

tio

n:

Dir

ectl

y in

ject

ed

Red

blo

od

cel

ls: 1

0 o

r 100

-fo

ld d

iluti

on

in

wat

er

Size

exc

lusi

on

Bi

oSE

p-S

EC 2

000

300

x 4.

6 m

m

0.8

ml/

min

Pe

rkin

-Elm

er

Scie

x EL

AN

610

0 D

RC p

lus

Iso

crat

ic

10 m

M

amm

on

ium

b

icar

bo

naa

t (p

H

7.4)

On

line

In v

itro

[140

]

Tabl

e 5:

Con

tinue

d

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

wra

te

Det

ecti

on

IS

+m

ob

ile p

has

e R

emar

ks

Ref

.

Salin

e so

luti

on

C

isp

lati

n

Met

hio

nin

e b

ind

ing

Pr

ior t

o s

pec

iati

on

: 1:

100

dilu

tio

n

Prio

r to

ICP-

MS:

O

nlin

e d

iluti

on

Cat

ion

exc

han

ge

Dio

nex

, C

S12A

, 25

0 x

2 m

m

0.25

ml/

min

via

T-

pie

ce w

ith

mak

e u

p

liqu

id

Perk

in-E

lmer

Sc

iex

ELA

N D

RC-I

I p

lus

Gra

die

nts

100

mM

HC

l (A

), w

ater

(B).

Max

imal

500

mM

H

Cl

On

line

IDM

S

enri

ched

19

6 Pt

In v

itro

[145

]

Pho

sph

ate

bu

ffer

ed

salin

e C

isp

lati

n a

nd

two

an

alo

gu

es

Alb

um

ine

bin

din

g

NS

Cap

illar

y el

ectr

op

ho

resi

s Fu

sed

-sili

ca

cap

illar

y (9

0 cm

x

75 µ

m)

NS

Via

T-p

iece

wit

h

mak

e u

p li

qu

id

Ag

ilen

t mo

del

75

00

Mic

roco

nce

ntr

ic

neb

ulis

er

IS: 7

2 Ge

15 m

M

ph

osp

hat

e b

uff

er

pH

7.4

co

nd

uct

ed

at 3

0 kV

(17µ

A)

On

line

In v

itro

[122

]

DN

A a

nd

o

ligo

nu

cleo

tid

es

Cis

pla

tin

D

NA

an

d

olig

on

ucl

eoti

de

bin

din

g

Prio

r to

sp

ecia

tio

n:

Dig

esti

on

wit

h

DN

Ase

I an

d

nu

clea

se P

1

Prio

r to

ICP-

MS:

Fr

acti

on

s d

ilute

d

An

ion

exc

han

ge

Mo

no

-Q H

R 5/

5 w

ith

ch

arg

ed

gro

up

of C

H2N

+

(CH

3) 3

1 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 6

000

Cro

ss-f

low

n

ebu

liser

Gra

die

nt

12.5

mM

Tri

s-H

Cl,

(pH

8.8

) (A

) an

d

12.5

mM

Tri

s-H

Cl,

pH

8.8

wit

h 1

M

NaC

l (B

)

Off

lin

e

In v

itro

[170

]

Pho

sph

ate

bu

ffer

ed

salin

e O

xalip

lati

n

Ho

lo-t

ran

sfer

rin

b

ind

ing

Pr

ior t

o s

pec

iati

on

: D

irec

tly

inje

cted

Si

ze e

xclu

sio

n

Bio

sep

-SEC

200

0 30

0 x

4.6

mm

0.

8 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 6

100

DRC

plu

s

Iso

crat

ic

10 m

M

amm

on

ium

b

icar

bo

naa

t (p

H

7.4)

On

line

In v

itro

[141

]

Plas

ma

Aq

ueo

us

solu

tio

n

KP10

19

Alb

um

in a

nd

tr

ansf

erri

n b

ind

ing

N

S Si

ze e

xclu

sio

n

An

ion

exc

han

ge

SEC

: Bio

Ass

ist

G3S

WxL

300

x 7

.8

mm

IEC

: SK

-DEA

E-N

PR

colu

mn

SEC

: 1 m

l/m

in

IEC

: 0.3

5 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N D

RC-I

I SE

C: 2

150

mM

N

aCl,

20 m

M T

ris

HC

l (p

H 7

.4)

IEC

Gra

die

nt

wat

er (A

), 15

0 m

M N

aCl+

20m

M

Tris

pH

10

(B),

500

+20

mM

Tri

s p

H

10, (

C)

(Max

imal

-500

m

M N

aCl )

On

line

2-D

sp

ecia

tio

n

In v

ivo

[143

]

Seru

m

KP10

19

Pro

tein

bin

din

g

NS

Size

exc

lusi

on

An

ion

exc

han

ge

NS

NS

ICP-

MS

wit

h D

RC

NS

On

line

2-D

sp

ecia

tio

n

In v

ivo

[144

]

Tabl

e 5:

Con

tinue

d

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

wra

te

Det

ecti

on

IS

+m

ob

ile p

has

e R

emar

ks

Ref

.

Pho

sph

ate

bu

ffer

ed

salin

e KP

1019

A

lhu

min

e an

d

tran

sfer

rin

bin

din

g

NS

Cap

illar

y el

ectr

op

ho

resi

s N

S N

S A

gile

nt 7

500

MS

Mic

roco

nce

ntr

ic

neb

ulis

er

IS:72

Ge

On

line

In v

itro

[111

]

G, o

ligo

nu

cleo

tid

es, a

nd

D

NA

C

isp

lati

n

Plat

inu

m b

ou

nd

to

G

, olig

on

ucl

eoti

des

, an

d D

NA

Prio

r to

sp

ecia

tio

n:

Dilu

tio

n

Reve

rsed

ph

ase

C8

250

x 2.

1 m

m, 5

µ

m

0.2

ml/

min

A

gile

nt 7

500

wit

h

a co

llisi

on

cel

l sy

stem

Mic

ron

ebu

liser

Co

nce

ntr

ic

neb

ulis

er

Iso

crat

ic

60 m

M

amm

on

ium

ac

etat

e (p

H 5

.8)

and

7.5

%

met

han

ol

On

line

In v

itro

[136

]

Plas

ma

Red

blo

od

cel

ls

Oxa

lipla

tin

H

aem

og

lob

in

bin

din

g

Prio

r to

sp

ecia

tio

n:

1:10

00 d

iluti

on

Si

ze e

xclu

sio

n

Bio

sep

-SEC

200

0 30

0 x

4.6

mm

0.

8 m

l/m

in

Perk

in-E

lmer

Sc

iex

ELA

N 6

100

DRC

plu

s

10 m

M

amm

on

ium

b

icar

bo

nat

e (p

H

7.4)

On

line

In v

ivo

[171

]

Gel

fro

m S

DS-

PAG

E N

AM

I-A

Pr

ote

in b

ind

ing

N

A

2-D

gel

el

ectr

op

ho

resi

s N

A

NA

LA

-IC

P-M

S

Perk

in-E

lmer

Sc

iex

ELA

N 6

100

DRC

ICP-

MS

NA

O

fflin

e

In v

itro

[127

]

Plas

ma

Car

bo

pla

tin

Pr

ote

in b

ind

ing

Pr

ior t

o s

pec

iati

on

: d

irec

tly

inje

cted

Si

ze e

xclu

sio

n

Silic

a b

ased

To

soH

aas

G30

00SW

x2

SEC

30

cm x

7.8

m

m, 5

µm

1 m

l/m

in

Plas

maq

uad

3 V

G

Elem

enta

l

IS:N

S

Iso

crat

ic

50 m

M T

ris-

HC

l (p

H 7

.4)

On

line

In v

ivo

In v

itro

[138

]

Pho

sph

ate

bu

ffer

ed

salin

e KP

1019

A

lbu

min

e,

glu

tath

ion

e, a

nd

tr

ansf

erri

n b

ind

ing

NS

Cap

illar

y el

ectr

op

ho

resi

s Fu

sed

-sili

ca

cap

illar

y (3

0.5

and

39

cm

, 75

µm

id

and

375

od

)

NS

Ag

ilen

t mo

del

75

00 w

ith

Mic

roco

nce

ntr

ic

neb

ulis

er

IS: 7

2 Ge

10 m

M

ph

osp

hat

e b

uff

er,

100

mM

NaC

l (p

H

7.4)

On

line

In v

itro

[172

]

Tabl

e 6:

Spe

ciat

ion

of m

etal

-bas

ed a

ntic

ance

r com

poun

ds in

env

ironm

enta

l sam

ples

(NS

= no

t spe

cifie

d)

Sam

ple

C

om

po

un

d

Ap

plic

atio

n

Sam

ple

p

rep

arat

ion

Sp

ecia

tio

n

Co

lum

n

Flo

w r

ate

Det

ecti

on

IS

+m

ob

ile

ph

ase

Rem

arks

R

ef.

Wat

er

Cis

pla

tin

Car

bo

pla

tin

Spec

iati

on

in w

aste

w

ater

N

S A

nio

n e

xch

ang

e H

yper

sil-

OD

S C

18

80 x

4.6

mm

, 3µ

m

Ch

rom

asil-

OD

S C

18

300

x 4

mm

, 5µ

m

0.8

ml/

min

Pe

rkin

-Elm

er S

ciex

EL

AN

600

0

Ult

raso

nic

neb

ulis

er

Gra

die

nt

Hex

adec

yltr

imet

hyl

amm

on

ium

b

rom

ide

+

citr

ateb

uff

er (p

H

7) in

m

eth

ano

l/w

ater

(1

0:90

to

50:

50)

On

line

Inte

rfac

e

[132

]

Wat

er

Uri

ne

Cis

pla

tin

Sp

ecia

tio

n in

was

te

wat

er a

nd

uri

ne

Prio

r to

sp

ecia

tio

n:

Wat

er: d

irec

tly

inje

cted

Uri

ne:

1:20

dilu

tio

n

Reve

rsed

ph

ase

Hyp

ersi

l-K

eyst

on

e H

yper

carb

150

x 2

.1

mm

0.25

-0.3

ml/

min

IDM

S 0.

11 m

l/m

in

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

DRC

-II

plu

s

Ther

mo

Fin

nig

an

Elem

ent 1

Gra

die

nt

1 m

M N

aOH

(A)

wat

er (B

)

On

line

IDM

S

enri

ched

196 Pt

[95]

Uri

ne

Was

te w

ater

Cis

pla

tin

Oxa

lipla

tin

Car

bo

pla

tin

Mo

nit

or h

osp

ital

w

aste

wat

er a

nd

u

rin

e

Prio

r to

sp

ecia

tio

n:

Wat

er:

dir

ectl

y in

ject

ed

Uri

ne:

1:20

dilu

tio

n

Reve

rsed

ph

ase

HS-

F5 1

50 x

2.1

mm

(p

enta

fluo

rph

enyl

pro

pyl

-bo

nd

ed s

ilica

)

0.25

ml/

min

Perk

in-E

lmer

Sci

ex

ELA

N D

RCII

Gra

die

nt

20 m

M

amm

on

ium

fo

rmat

e (in

4 v

/v

% m

eth

ano

l) p

H

3.75

(A),

wat

er

(B),

met

han

ol (

C).

Max

imal

12%

m

eth

ano

l

On

line

[133

]

Was

te w

ater

C

isp

lati

n

Car

bo

pla

tin

[PtC

l 4]2-

[PtC

l 6]2

Spec

iati

on

in w

aste

w

ater

Pr

ior t

o s

pec

iati

on

:

Dir

ectl

y in

ject

ed

Reve

rsed

ph

ase

HS-

F5 1

50 x

2.1

mm

(p

enta

fluo

rph

enyl

pro

pyl

-bo

nd

ed s

ilica

)

0.25

ml/

min

Perk

in-E

lmer

Sci

ex

ELA

N 6

000

DRC

-II

plu

s

Gra

die

nt :

20 m

M

amm

on

ium

fo

rmat

e (in

4 v

/v

% m

eth

ano

l) p

H

3.75

(A),

wat

er

(B),

met

han

ol (

C).

Max

imal

12%

m

eth

ano

l

[1

34]

ICP-MS in oncology

55

4.2 Assay development

The development of speciation methods using ICP-MS for Pt or Ru detection is generally

focused on the selection of a speciation system that is capable of separating the

compounds of interest and that is compatible with ICP-MS. This compatibility can be

improved by instrumental modifications. In addition, it is relevant that the mobile phase

is not reactive with the Pt compounds and their metabolites.

High performance liquid chromatography (HPLC) combined with ICP-MS detection is a

versatile speciation technique. The chromatographic columns used in the investigations

published in literature on the speciation of metal-based compounds in combination

with ICP-MS included reversed phase, size exclusion, ion-exchange, and ion-pair

chromatography (Table 5). The ion-pairing reagents were anionic and cationic

surfactants.

ICP-MS can be used as an off-line or on-line detector. When online detection is used, the

coupling of the chromatographic method to the ICP-MS is achieved by connecting the

outlet of the column to the liquid sample inlet of the nebuliser. The feasibility of

coupling HPLC-ICP-MS is mainly affected by the composition and flow rate of the mobile

phases used to perform chromatographic separation. The high amounts of organic

solvents frequently used in reversed or normal phase HPLC results in physical changes of

the plasma. This might cause plasma instability or even extinguishing of the plasma

[135]. Further problems are encountered with organic solvents when performing

gradient elution. As the eluent composition changes, the nature of the plasma changes,

which could lead to a variable sensitivity during the gradient elution and thus calibration

problems [118]. Furthermore, the presence of high levels of organic solvent or dissolved

solids (e.g. salts) can result in constriction of the cone orifices owing to build-up of solids.

High solid contents can also lead to clogging of the nebuliser. These issues affect the

robustness, sensitivity, and precision of the technique. Therefore, the organic solvent

load and salt content of the mobile phase should be kept to a minimum.

4.2.1 Reversed phase chromatography (RP)

RP is the most commonly used speciation mechanism in liquid chromatography and

consists of a hydrophobic stationary phase bonded to a solid support. The mobile

phases are less hydrophobic, usually water containing different amounts of organic

modifiers such as methanol or acetonitril. The sample compounds are partitioned

between the mobile phase and stationary phase. Analyte retention is determined by the

affinity for each of the two phases and can be altered by changes in e.g pH. The main

limitation of RP-HPLC-ICP-MS is that most organic modifiers are not ICP-MS compatible.

Despite this limitation, however, RP is the most frequently used technique in the

speciation of metabolites of metal-based anticancer agents. Several approaches have

Chapter 1.2

56

been used to surmount the compatibility problems. Screnci et al. avoided the use of

large amounts of organic solvent and used a mobile phase consisting of 100% water to

assess differences in hydrophobicity of several Pt compounds in order to evaluate

relationships between hydrophobicity, Pt accumulation in dorsal root ganglia, and

neurotoxicity [56]. For satraplatin, however, methanol was added. Allain et al. also used

isocratic elution with a mobile phase containing a large percentage of water (90%) for

the speciation of oxaliplatin biotransformation products [120]. Hann et al. hyphenated

RP-HPLC to ICP-MS for the analysis of cisplatin metabolites in waste water [95]. They

used gradient elution from 1mM sodium hydroxide to water, entering the ICP-MS

through a Tee. The other entrance of the Tee was used to introduce the enriched 196Pt

isotope for online IDMS. Because of low salt concentrations and the absence of organic

solvent, no compatibility issues were met. A gradient elution with up to 12% methanol,

used for the speciation of several Pt anticancer agents in hospital waste water, was

described by the same authors [133]. Smith et al. reported the speciation and

quantification of metabolites of ZD0473 using isocratic elution with a mobile phase

containing 20% of methanol [26]. The deposition of carbon on the cones was prevented

by the addition of oxygen to the nebuliser gas. Cairns et al. used an eluent containing

water and acetonitrile for the speciation of metabolites of satraplatin [117,118]. To allow

a gradient elution with acetonitrile concentrations up to 95%, an interface was

developed to desolvate the HPLC eluent prior to introduction to the plasma [118]. The

interface consisted of a common concentric nebuliser, a heated spray chamber, a

membrane desolvator, and a condenser. Galettis et al. also developed a method for the

quantification of satraplatin metabolites [119]. They, however, managed to develop and

validate a method with gradient elution using methanol concentrations from 25 to 45%,

without the necessity to desolvate the mobile phase. Methanol was chosen because of

its lower vapour pressure and carbon loading compared to acetonitrile. The gradient

from 25 to 45% of methanol suppressed Pt signals by 70%. Because not all metabolites

showed a similar signal suppression, to achieve reliable results, this approach does

require the availability of reference compounds of all the metabolites tested.

Metabolites should be quantified using calibration curves for each separate metabolite.

Another application using similar chromatographic conditions was described with

offline ICP-MS detection [82].

In addition to the speciation of metabolites of metal-based anticancer agents, RP proved

to be suitable for studying the reaction products of cisplatin with methionine [123],

using an isocratic elution with an aqueous mobile phase. The interaction of cisplatin

with DNA nucleotides was also studied [136] using isocratic elution with low amounts of

methanol. The investigators tested two columns with different characteristics and

optimal flow rates. The eluent from the C8 column (0.2 ml/min) was introduced into the

ICP-MS via a micronebuliser to facilitate the introduction of low flow rates, whereas for

ICP-MS in oncology

57

the C16 column (1 ml/min) a common concentric nebuliser was used. Both methods

were compatible with ICP-MS because low amounts of methanol and isocratic elution

were used. With the C8 column, however, a better speciation of nucleobases was

achieved.

4.2.2 Reversed phase ion-pairing chromatography (RPIP)

RPIP is used for the speciation of ionic or ionisable compounds for which an ion-pair (IP)

is formed between the solute ion and an appropriate counter ion. The resulting IP is

partitioned between the mobile and stationary phase. The mobile and stationary phases

used in RPIP are similar to those employed in RP-HPLC, although an IP reagent is added

to the mobile phase. The IP reagent contains a polar and non-polar moiety.

Because of their positive charges, Pt metabolites can be successfully separated using

RPIP. Zhao et al. applied RPIP to retain ionic and neutral cisplatin derivatives [114].

Hydrolysis products of cisplatin were separated using sodium dodecylsulfate (SDS) or

heptanosulfonate as IP reagent. It was shown that SDS had low activity with cisplatin

and its metabolites [137]. Heptanosulfonate, which resulted in a better speciation,

however, did shift the equilibrium of hydrolysis when the analytes were diluted in the

mobile phase. Because the shift was slow, this effect could be avoided by diluting the

sample immediately before injection. Heptanosulfonate was also used for the speciation

of the reaction products of cisplatin with thiol compounds. The organic content of the

mobile phases was purposely kept low (3% n-propanol for the hydrolysis products and

10% methanol for the thiol compounds) to ensure the stability of the plasma. SDS in a

3% methanol containing mobile phase was also used in another application for the

speciation of cisplatin and monohydrated cisplatin [115]. Although the mobile phase

changed the overall appearance of the argon plasma, the Pt counts were only slightly

suppressed and thus adequate results were achieved.

Although the effect of the applied IP reagents on cisplatin metabolism in the previous

publications seemed to be insignificant, the possible effect of IP reagents should be

taken into account when developing RPIP methods for the speciation of Pt compounds.

4.2.3 Size exclusion chromatography (SEC)

SEC is used to separate molecules according to their effective size in solution using a

stationary phase gel with pores of a particular dimension. Molecules that are too large to

enter pores elute first, while small molecules interact with the stationary phase and elute

depended on their size. This technique is useful for the speciation of proteins. The

mobile phases used for SEC are usually buffered solutions such as Tris-HCl or phosphate

buffers [138]. Some applications, however, use high saline solutions as a mobile phase,

which limits the hyphenation of SEC to ICP-MS. Several methods for the speciation and

Chapter 1.2

58

quantification of Pt bound to proteins have been reported. Allain et al. demonstrated

the binding of oxaliplatin to γ-globulins, albumin, and haemoglobin using an aqueous

mobile phase with 150 mM sodium hydroxide [120]. No compatibility problems were

described. The interaction of cisplatin with haemoglobin was also extensively studied

with SEC using a mobile phase of 10 mM ammonium bicarbonate [139,140]. A similar

procedure was utilised to separate oxaliplatin and oxaliplatin bound to holo-transferrin

[141]. The interaction of carboplatin with plasma proteins analysed by SEC was also

reported [138]. For this application a mobile phase of 50 mM Tris-HCl was used, because

a mobile phase containing phosphate buffer resulted in an unstable plasma. Szpunar et

al. investigated various SEC systems for the speciation of the protein-bound and

unbound fractions of Pt and Ru drugs prior to online ICP-MS detection [142]. The mobile

phase used (30 mM Tris-HCl buffer) was well tolerated by the ICP-MS. The speciation of

Pt and Ru compounds using SEC was complicated because parent compounds and

metabolites were partly retained on the stationary phase. Furthermore, the authors

demonstrated that using SEC it was not possible to separate albumin- and transferring-

bound drug because of a proximity in molar masses. Another speciation mechanism

such as anion-exchange could solve this limitation. Other authors also mentioned that

SEC alone could not adequately distinguish between metallo-proteins that show only

small differences in amino acid sequence. They, therefore, combined SEC with anion-

exchange chromatography coupled to ICP-MS to separate and quantify Ru-albumine

and Ru-transferrin adducts [143,144].

4.2.4 Ion-exchange chromatography (IEC)

IEC is based on the interactions of charged functional groups of the stationary phase

with charged analytes. Two types of IEC exist; cation-exchange, where positively charged

analytes react with anionic sites on the column and anion-exchange, where cationic sites

on the column are used to react with negatively charged analytes. Mobile phases

generally consist of an aqueous salt buffer, which might cause difficulties when

hyphenating IEC to ICP-MS. Additionally, an organic modifier is often added to the

mobile phase. The pH of the mobile phase is of great interest because it affects the

dissociation of the compounds analysed.

A cation-exchange method with ICP-MS detection was reported for the speciation of

BBR3464 and drug related products [116]. Because of the strong cationic character of the

compounds, RP-HPLC appeared to be unsuitable. Speciation of the compounds could be

achieved by the addition of an IP reagent and 35% acetonitrile to the mobile phase. This,

however, resulted in a progressive loss of sensitivity. The use of cation-exchange

chromatography, with a gradient from 20 to 200 mM pyridine, resulted in a proper

separation. Another cation-exchange method was applied to separate cisplatin, its

ICP-MS in oncology

59

metabolites, methionine, and their adducts [145]. The method employed gradient

elution fom 0.01 to 0.05 mM HCl and the mobile phase did not affect cisplatin kinetics.

No problems for the hyphenation of cation-exchange chromatography to ICP-MS for any

of the described applications were reported.

Falter et al. applied a solvent-generated anion exchanger for the separation of cisplatin

and carboplatin [132]. The column was pre-treated with hexadecyltrimethylammonium

bromide (HTAB) and speciation was carried out by the use of a gradient containing 10 to

50% methanol in water. The high amounts of organic solvent could be used because of

the application of an ultrasonic nebuliser followed by desolvation, minimising the

solvent load entering the ICP. Anion-exchange was also applied to separate the cisplatin

reaction products with DNA after enzymatic digestion. The negatively charged

nucleotides could be separated using a stationary phase with quaternary ammonium

groups [70,146]. The platinated DNA was digested and eluted off the column by a linear

salt gradient of NaCl. Fractions were collected and, after dilution, analysed using ICP-MS.

The interaction of cisplatin with guanosine monophosphate has been investigated by

Hann et al. using anion-exchange chromatography coupled to sector field ICP-MS [147].

The column was also functionalised with quaternary ammonium groups. The mobile

phase consisted of a carbonate buffer system with 5% acetonitril to reduce retention

times. To avoid matrix interferences, only 22% of the mobile phase was directed to the

ICP-MS. Before entering the introduction system the mobile phase was diluted by a

make-up flow containing 1% HNO3 and IS.

4.2.5 Speciation techniques other than liquid chromatography

In addition to methods using liquid chromatography, few other speciation methods

hyphenated to ICP-MS detection for the analysis of metal-based anticancer drugs have

been described. Gel electrophoresis (SDS-PAGE) was used for the speciation of serum

proteins after reaction with Pt [82,148]. The bands from the gel were cut out and

dissolved in 70% HNO3 and then heated at 90 °C before being diluted with water [82] or

they were extracted with aqua regia [148]. The Pt content of the bands was measured by

ICP-MS. Timeraev et al. coupled capillary electrophoresis to ICP-MS using an interface

consisting of a microconcentric nebuliser to study the interaction of cisplatin with

albumine [122]. A similar method was applied to study the interaction of KP1019 with

albumine and transferrin [111].

Chapter 1.2

60

5 Conclusions and perspectives

The successful application of ICP-MS in oncology has had an enormous impact on the

field of quantitative analysis of metal-based anticancer agents from biological and

environmental samples. ICP-MS provides enormous sensitivity, which makes the

technique applicable to study metal pharmacokinetics, metal accumulation in cells, and

DNA- and protein-binding. Furthermore, environmental monitoring to assess the

exposure of hospital personnel to metal-based anticancer agents can be studied using

this technique. Pretreatment of the samples depends on the composition of the sample

and on the available instrument and is generally focused on the reduction of matrix

effects. In the absence of a suitable matrix removal pretreatment procedure, matrix

effects can be circumvented by using appropriate calibration methods.

In addition to the analysis of total metal content of samples, the hyphenation of

speciation techniques to ICP-MS has extended the application of ICP-MS to the analysis

of parent compounds, metabolites, and adducts of metal-based anticancer agents. The

development of speciation methods using ICP-MS is commonly focused on the selection

of a speciation system that is capable of separating the compounds of interest and that

is compatible with ICP-MS.

Publications discussed in this paper appeared in the last 17 years. During this period, the

application of ICP-MS to study metal-based anticancer agents increased rapidly.

Currently available ICP-MS instruments are capable of quantifying picograms of metallic

elements, and thus can detect ambient Pt or Ru reliably in humans. With this capability,

current ICP-MS technologies offer great potential for continued investigations of metal-

based oncology treatments and the elucidation of their anticancer mechanisms.

ICP-MS in oncology

61

References 1. Rosenberg B, Vancamp L, Krigas T. Inhibition of cell division in escherichia coli by electrolysis products

from a platinum electrode. Nature 1965; 205: 698-9.

2. Kelland L. The resurgence of platinum-based cancer chemotherapy. Nat Rev Cancer 2007; ..

3. Rademaker-Lakhai JM, van den BD, Pluim D, Beijnen JH, Schellens JH. A Phase I and pharmacological study with imidazolium-trans-DMSO-imidazole-tetrachlororuthenate, a novel ruthenium anticancer agent. Clin Cancer Res 2004; 10: 3717-27.

4. Hartinger CG, Zorbas-Seifried S, Jakupec MA, Kynast B, Zorbas H, Keppler BK. From bench to bedside--preclinical and early clinical development of the anticancer agent indazolium trans-[tetrachlorobis(1H-indazole)ruthenate(III)] (KP1019 or FFC14A). J Inorg Biochem 2006; 100: 891-904.

5. Scheulen ME, Kynast B, Jaehde U, Hilger RA, Scheuch E, Arion V, Kupsch P, Henke MM, Dittrich C, Keppler BK. A phase I dose-escalation trial with the new redox activated compound sodium trans-[tetrachlorobis(1H-indazole)ruthenate(III)]/indazolhydrochloride (1:1.1) (FFC14A) in patients with solid tumors - a CESAR study. J Clin Oncol , 2004 ASCO Annual Meeting Proceedings 2004; 22: 2101.

6. Brouwers EEM, Tibben MM, Rosing H, Hillebrand MJX, Joerger M, Schellens JHM, Beijnen JH. Sensitive inductively coupled plasma mass spectrometry assay for the determination of platinum originating from cisplatin, carboplatin, and oxaliplatin in human plasma ultrafiltrate. J Mass Spectrom 2006; 41: 1186-94.

7. Crul M, van den Bongard HJ, Tibben MM, van Tellingen O, Sava G, Schellens JH, Beijnen JH. Validated method for the determination of the novel organo-ruthenium anticancer drug NAMI-A in human biological fluids by Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 2001; 369: 442-5.

8. Vermorken JB, van der Vijgh WJ, Pinedo HM. Pharmacokinetic evidence for an enterohepatic circulation in a patient treated with cis-dichlorodiammineplatinum (II). Res Commun Chem Pathol Pharmacol 1980; 28: 319-28.

9. Nygren O, Vaughan GT, Florence TM, Morrison GM, Warner IM, Dale LS. Determination of platinum in blood by adsorptive voltammetry. Anal Chem 1990; 62: 1637-40.

10. Gelevert T, Messerschmidt J, Meinardi MT, Alt F, Gietema JA, Franke JP, Sleijfer DT, Uges DR. Adsorptive voltametry to determine platinum levels in plasma from testicular cancer patients treated with cisplatin. Ther Drug Monit 2001; 23: 169-73.

11. Gerl A, Schierl R. Urinary excretion of platinum in chemotherapy-treated long-term survivors of testicular cancer. Acta Oncol 2000; 39: 519-22.

12. Gietema JA, Meinardi MT, Messerschmidt J, Gelevert T, Alt F, Uges DR, Sleijfer DT. Circulating plasma platinum more than 10 years after cisplatin treatment for testicular cancer. Lancet 2000; 355: 1075-6.

13. Shearan P, Smyth MR. Comparison of voltammetric and graphite furnace atomic absorption spectrometric methods for the direct determination of inorganic platinum in urine. Analyst 1988; 113: 609-12.

14. Vrana O, Brabec V, Kleinwachter V. Polarographic studies on the conformation of some platinum complexes: relations to anti-tumour activity. Anticancer Drug Des 1986; 1: 95-109.

15. Xilei L, Heydorn K, Rietz B. Limit of detection for the determination of Pt in biological material by RNAA using electrolytic separation of gold. Journal of Radioanalytical and Nuclear Chemistry 1992; 160: 85-99.

16. Rietz B, Heydorn K, Krarup-Hansen A. Determination of platinum by neutron activation analysis in nerve tissue from rats treated with cisplatin. Biol Trace Elem Res 1994; 43-45:343-50.: 343-50.

17. Gorodetsky R, Mou X, Vexler A, Kaufman B, Catane R, Loewenthal E. Noninvasive follow-up of platinum pharmacokinetics in the skin of patients on cisplatin chemotherapy. Cancer 1993; 72: 446-54.

18. Seifert WE, Jr., Stewart DJ, Benjamin RS, Caprioli RM. Energy-dispersive X-ray fluorescence determination of platinum in plasma, urine, and cerebrospinal fluid of patients administered cis-dichlorodiammineplatinum(II). Cancer Chemother Pharmacol 1983; 11: 120-3.

19. Jonson R, Mattsson S, Unsgaard B. A method for in vivo analysis of platinum after chemotherapy with cisplatin. Phys Med Biol 1988; 33: 847-57.

20. Stewart DJ, Benjamin RS, Luna M, Feun L, Caprioli R, Seifert W, Loo TL. Human tissue distribution of platinum after cis-diamminedichloroplatinum. Cancer Chemother Pharmacol 1982; 10: 51-4.

21. Dominici C, Alimonti A, Caroli S, Petrucci F, Castello MA. Chemotherapeutic agent cisplatin monitoring in biological fluids by means of inductively-coupled plasma emission spectrometry (ICP-AES). Clin Chim Acta 1986; 158: 207-15.

Chapter 1.2

62

22. Brandsteterova E, Kiss F, Chovancova V, Reichelova V. HPLC analysis of platinum cytostatics. Neoplasma 1991; 38: 415-24.

23. Krull IS, Ding XD, Braverman S, Selavka C, Hochberg F, Sternson LA. Trace analysis for cis-platinum anti-cancer drugs via LCEC. J Chromatogr Sci 1983; 21: 166-73.

24. Bouma M, Nuijen B, Jansen MT, Sava G, Picotti F, Flaibani A, Bult A, Beijnen JH. Development of a LC method for pharmaceutical quality control of the antimetastatic ruthenium complex NAMI-A. J Pharm Biomed Anal 2003; 31: 215-28.

25. Ehrsson H, Wallin I. Liquid chromatographic determination of oxaliplatin in blood using post-column derivatization in a microwave field followed by photometric detection. J Chromatogr B Analyt Technol Biomed Life Sci 2003; 795: 291-4.

26. Smith CJ, wilson ID, Abou-Shakra F, Payne R, Parry TC, Sinclair P, Roberts DW. A comparison of the quantitative methods for the analysis of the platinum-containing anticancer drug [cis-[amminedichloro(2-methylpyridine)]platinum(II)] (ZD0473) by HPLC coupled to either a triple quadrupole mass spectrometer or an inductively coupled plasma mass spectrometer. Anal Chem 2003; 75: 1463-9.

27. Oe T, Tian Y, O'Dwyer PJ, Roberts DW, Malone MD, Bailey CJ, Blair IA. A validated liquid chromatography/tandem mass spectrometry assay for cis-amminedichloro(2-methylpyridine) platinum(II) in human plasma ultrafiltrate. Anal Chem 2002; 74: 591-9.

28. Minakata K, Nozawa H, Okamoto N, Suzuki O. Determination of platinum derived from cisplatin in human tissues using electrospray ionisation mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2006; 832: 286-91.

29. Tothill P, Matheson LM, Smyth JF. Inductively coupled plasma mass spectrometry for the determination of platinum in animal tissue and a comparison with atomic absorption spectrometry. J Anal Atom Spectrom 1990; 5: 619-22.

30. Jarvis KE, Gray AL, Houk RS. Handbook of inductively coupled plasma mass spectrometry. Chapman & Hall. London: Blackie Academic & Professional, 1992

31. Giessmann U, Greb U. High resolution ICP-MS - A new concept for elemental mass spectrometry. Fresenius J Anal Chem 1994; 350: 186-93.

32. Bings NH. Plasma time-of-flight mass spectrometry as a detector for short transient signals in elemental analysis. Anal Bioanal Chem 2005; 382: 887-90.

33. Evans HE, Giglio JJ. Interferences in inductively coupled plasma mass spectrometry-A review. J Anal Atom Spectrom 1993; 8: 18.

34. Dams RFJ, Goossens J, Moens L. Spectral and non-spectral interferences in inductively coupled plasma mass-spectrometry. Mikrochim Acta 1995; 119: 286.

35. Vanhaecke F, Saverwyns S, de Wannemacker G, Moens L, Dams R. Comparison of the applicaiton of higher mass resolution and cool plasma conditions to avoid spectral interferences in Cr(III)/Cr(VI) speciation by means of high-performance liquid chromatography-inductively coupled plasma mass spectrometry. Anal Chim Acta 2000; 419: 55-64.

36. Feldmann I, Jakubowski N, Thomas C, Struewer D. Application of a hexapole collision and reaction cell in ICP-MS. Part II: Analytical figures of merit and first applications. Fresenius J Anal Chem 1999; 365: 422-8.

37. Nonose N, Kubota M. Non-spectral and spectral interferences in inductively coupled plasma high-resolution mass spectrometry. Part I. Optical characteristics of micro-plasmas observed just behind the sampler and the skimmer in inductively coupled plasma high resolution mass spectrometry. J Anal Atom Spectrom 2001; 16: 551-9.

38. Hsiung CS, Andrade JD, Costa R, Ash KO. Minimizing interferences in the quantitative multielement analysis of trace elements in biological fluids by inductively coupled plasma mass spectrometry. Clin Chem 1997; 43: 2303-11.

39. Brouwers EEM, Tibben MM, Rosing H, Schellens JHM, Beijnen JH. Determination of ruthenium originating from the investigational anti-cancer drug NAMI-A in human plasma ultrafiltrate, plasma, and urine by inductively coupled plasma mass spectrometry. Rapid Commun Mass Spectrom 2007; 21: 1521-30.

40. Lustig L, Zang S, Michalke B, Schramel P, Beck W. Platinum determination in nutrient plants by inductively coupled plasma mass spectrometry with special respect to the hafnium oxide interference. Fresenius J Anal Chem 1997; 357: 1157-63.

41. Rudolph E, Hann S, Stingeder G, Reiter C. Ultra-trace analysis of platinum in human tissue samples. Anal Bioanal Chem 2005; 382: 1500-6.

ICP-MS in oncology

63

42. Rodushkin I, Engstrom E, Stenberg A, Baxter DC. Determination of low-abundance elements at ultra-trace levels in urine and serum by inductively coupled plasma-sector field mass spectrometry. Anal Bioanal Chem 2004; 380: 247-57.

43. Hann S, Koellensperger G, Kanitsar K, Stingeder G, Brunner M, Erovic B, Muller M, Reiter C. Platinum determination by inductively coupled plasma-sector field mass spectrometry (ICP-SFMS) in different matrices relevant to human biomonitoring. Anal Bioanal Chem 2003; 376: 198-204.

44. Thomas R. A beginner's guide to ICP-MS: Part XII - A review of interferences. Spectroscopy 2002; 17: 24-31.

45. U.S.Food and Drug Administration, Center for Drug Evaluation and Research, Guidance for Industry, Bioanalytical Method Validation. FDA. U.S. Food and Drug Administration: Center for Drug Evaluation and Research: Guidance for Industry: Bioanalytical Method Validation. http://www fda gov/cder/ guidance/4252fnl pdf 2001.

46. Rosing H, Man WY, Doyle E, Bult A, Beijnen JH. Bioanalytical liquid chromatography method validation: a review of current practices and procedures. J Liq Chrom & Rel Technol 2000; 23: 329-54.

47. Desoize B, Madoulet C. Particular aspects of platinum compounds used at present in cancer treatment. Crit Rev Oncol Hematol 2002; 42: 317-25.

48. Alessio E, Mestroni G, Bergamo A, Sava G. Ruthenium antimetastatic agents. Curr Top Med Chem 2004; 4: 1525-35.

49. Calvert H, Judson I, van der Vijgh WJ. Platinum complexes in cancer medicine: pharmacokinetics and pharmacodynamics in relation to toxicity and therapeutic activity. Cancer Surv 1993; 17: 189-217.

50. Graham MA, Lockwood GF, Greenslade D, Brienza S, Bayssas M, Gamelin E. Clinical pharmacokinetics of oxaliplatin: a critical review. Clin Cancer Res 2000; 6: 1205-18.

51. Yang Z, Hou X, Jones BT. Determination of platinum in clinical samples. Applied Spectroscopy Reviews 2002; 37: 57-88.

52. De Waal WA, Maessen FJ, Kraak JC. Analytical methodologies for the quantitation of platinum anti-cancer drugs and related compounds in biological media. J Pharm Biomed Anal 1990; 8: 1-30.

53. Tothill P, Klys HS, Matheson LM, McKay K, Smyth JF. The long-term retention of platinum in human tissues following the administration of cisplatin or carboplatin for cancer therapy. Eur J Cancer 1992; 28: 1358-61.

54. Robbins ME, Bywaters TB, Jaenke RS, Hopewell JW, Matheson LM, Tothill P, Whitehouse E. Long-term studies of cisplatin-induced reductions in porcine renal functional reserve. Cancer Chemother Pharmacol 1992; 29: 309-15.

55. Ding H, Goldberg MM, Raymer JH, Holmes J, Stanko J, Chaney SG. Determination of platinum in rat dorsal root ganglion using ICP-MS. Biol Trace Elem Res 1999; 67: 1-11.

56. Screnci D, McKeage MJ, Galettis P, Hambley TW, Palmer BD, Baguley BC. Relationships between hydrophobicity, reactivity, accumulation and peripheral nerve toxicity of a series of platinum drugs. Br J Cancer 2000; 82: 966-72.

57. Boulikas T, Vougiouka M. Cisplatin and platinum drugs at the molecular level. (Review). Oncol Rep 2003; 10: 1663-82.

58. Bose RN. Biomolecular targets for platinum antitumor drugs. Mini Rev Med Chem 2002; 2: 103-11.

59. Eastman A. The formation, isolation and characterization of DNA adducts produced by anticancer platinum complexes. Pharmacol Ther 1987; 34: 155-66.

60. Fichtinger-Schepman AM, van der Veer JL, den Hartog JH, Lohman PH, Reedijk J. Adducts of the antitumor drug cis-diamminedichloroplatinum(II) with DNA: formation, identification, and quantitation. Biochemistry 1985; 24: 707-13.

61. Reedijk J. Why does Cisplatin reach Guanine-N7 with competing S-donor ligands available in the cell? Chem Rev 1999; 99: 2499-510.

62. Centerwall CR, Tacka KA, Kerwood DJ, Goodisman J, Toms BB, Dubowy RL, Dabrowiak JC. Modification and uptake of a cisplatin carbonato complex by Jurkat cells. Mol Pharmacol 2006; 70: 348-55.

63. Bonetti A, Apostoli P, Zaninelli M, Pavanel F, Colombatti M, Cetto GL, Franceschi T, Sperotto L, Leone R. Inductively coupled plasma mass spectroscopy quantitation of platinum-DNA adducts in peripheral blood leukocytes of patients receiving cisplatin- or carboplatin-based chemotherapy. Clin Cancer Res 1996; 2: 1829-35.

64. Liu J, Kraut E, Bender J, Brooks R, Balcerzak S, Grever M, Stanley H, D'Ambrosio S, Gibson-D'Ambrosio R, Chan KK. Pharmacokinetics of oxaliplatin (NSC 266046) alone and in combination with paclitaxel in cancer patients. Cancer Chemother Pharmacol 2002; 49: 367-74.

Chapter 1.2

64

65. Cooper BW, Veal GJ, Radivoyevitch T, Tilby MJ, Meyerson HJ, Lazarus HM, Koc ON, Creger RJ, Pearson G, Nowell GM, Gosky D, Ingalls ST, Hoppel CL, Gerson SL. A phase I and pharmacodynamic study of fludarabine, carboplatin, and topotecan in patients with relapsed, refractory, or high-risk acute leukemia. Clin Cancer Res 2004; 10: 6830-9.

66. McDonald ES, Randon KR, Knight A, Windebank AJ. Cisplatin preferentially binds to DNA in dorsal root ganglion neurons in vitro and in vivo: a potential mechanism for neurotoxicity. Neurobiol Dis 2005; 18: 305-13.

67. Rice JR, Gerberich JL, Nowotnik DP, Howell SB. Preclinical efficacy and pharmacokinetics of AP5346, a novel diaminocyclohexane-platinum tumor-targeting drug delivery system. Clin Cancer Res 2006; 12: 2248-54.

68. Walker DL, Reid JM, Svingen PA, Rios R, Covey JM, Alley MC, Hollingshead MG, Budihardjo II, Eckdahl S, Boerner SA, Kaufmann SH, Ames MM. Murine pharmacokinetics of 6-aminonicotinamide (NSC 21206), a novel biochemical modulating agent. Biochem Pharmacol 1999; 58: 1057-66.

69. Bible KC, Boerner SA, Kirkland K, Anderl KL, Bartelt D, Jr., Svingen PA, Kottke TJ, Lee YK, Eckdahl S, Stalboerger PG, Jenkins RB, Kaufmann SH. Characterization of an ovarian carcinoma cell line resistant to cisplatin and flavopiridol. Clin Cancer Res 2000; 6: 661-70.

70. Azim-Araghi A, Ottley CJ, Pearson DG, Tilby MJ. Sensitive detection of platinum-bound DNA using ICP-MS and comparison to graphite furnace atomic absorption spectroscopy (GFAAS). In: Plasma Source Mass Spectrometry. Holland JG, Tanner SD, eds. Holland JG, Tanner SD, eds.: Royal Society of Chemistry, 2001: 412-8

71. John T et al. Determination, by ICP-MS, of background and BBR 3464 induced levels of platinum bound to DNA isolated from blood cells: a comparison of the sensitivity of the perkin elmer sciex elan 6000 and the thermofinnigan neptune instruments. In: Plasma Source Spectrometry. Holland JG, Tanner SD, eds. Holland JG, Tanner SD, eds.: Royal Society of Chemistry, 2003: 82-90

72. Yamada K, Kato N, Takagi A, Koi M, Hemmi H. One-milliliter wet-digestion for inductively coupled plasma mass spectrometry (ICP-MS): determination of platinum-DNA adducts in cells treated with platinum(II) complexes. Anal Bioanal Chem 2005; 382: 1702-7.

73. Ta LE, Espeset L, Podratz J, Windebank AJ. Neurotoxicity of oxaliplatin and cisplatin for dorsal root ganglion neurons correlates with platinum-DNA binding. Neurotoxicology 2006; 27: 992-1002.

74. Zhang J, Zhao X. Synthesis, cytotoxicity and DNA-binding levels of ammine/propylamine platinum(II) complexes with carboxylates. Eur J Med Chem 2007; 42: 286-91.

75. Bjorn E, Nygren Y, Nguyen TT, Ericson C, Nojd M, Naredi P. Determination of platinum in human subcellular microsamples by inductively coupled plasma mass spectrometry. Anal Biochem 2007; 363: 135-42.

76. Lenz K, Hann S, Koellensperger G, Stefanka Z, Stingeder G, Weissenbacher N, Mahnik SN, Fuerhacker M. Presence of cancerostatic platinum compounds in hospital wastewater and possible elimination by adsorption to activated sludge. Sci Total Environ 2005; 345: 141-52.

77. Barefoot RR. Determination of platinum at trace levels in environmental and biological samples. Environmental Science & Technology 1997; 31: 309-14.

78. Morazzoni F, Canevali C, Moschetti I, Todeschini R, Caroli S, Alimonti A, Petrucci F, Ravasi G, Bedini AV, Milani F, . Determination of platinum in plasma of patients affected by inoperable lung carcinoma treated with radiotherapy and concurrent low-dose continuous infusion of cis-dichlorodiammine platinum(II). Cancer Chemother Pharmacol 1995; 35: 529-32.

79. Liu J, Kraut EH, Balcerzak S, Grever M, D'Ambrosio S, Chan KK. Dosing sequence-dependent pharmacokinetic interaction of oxaliplatin with paclitaxel in the rat. Cancer Chemother Pharmacol 2002; 50: 445-53.

80. Sessa C, Minoia C, Ronchi A, Zucchetti M, Bauer J, Borner M, de Jong J, Pagani O, Renard J, Weil C, D'Incalci M. Phase I clinical and pharmacokinetic study of the oral platinum analogue JM216 given daily for 14 days. Ann Oncol 1998; 9: 1315-22.

81. Sessa C, Capri G, Gianni L, Peccatori F, Grasselli G, Bauer J, Zucchetti M, Vigano L, Gatti A, Minoia C, Liati P, Van den BS, Bernareggi A, Camboni G, Marsoni S. Clinical and pharmacological phase I study with accelerated titration design of a daily times five schedule of BBR3464, a novel cationic triplatinum complex. Ann Oncol 2000; 11: 977-83.

82. Carr JL, Tingle MD, McKeage MJ. Rapid biotransformation of satraplatin by human red blood cells in vitro. Cancer Chemother Pharmacol 2002; 50: 9-15.

83. Casetta B, Roncadin M, Montanari G, Fulanut M. Determination of platinum in biological fluids by ICP-mass spectrometry. Atomic Spectroscopy 1991; 12: 81-6.

ICP-MS in oncology

65

84. Allain P, Berre S, Mauras Y, Le Bouil A. Evaluation of inductively coupled mass spectrometry for the determination of platinum in plasma. Biol Mass Spectrom 1992; 21: 141-3.

85. Gamelin E, Allain P, Maillart P, Turcant A, Delva R, Lortholary A, Larra F. Long-term pharmacokinetic behavior of platinum after cisplatin administration. Cancer Chemother Pharmacol 1995; 37: 97-102.

86. Gamelin E, Bouil AL, Boisdron-Celle M, Turcant A, Delva R, Cailleux A, Krikorian A, Brienza S, Cvitkovic E, Robert J, Larra F, Allain P. Cumulative pharmacokinetic study of oxaliplatin, administered every three weeks, combined with 5-fluorouracil in colorectal cancer patients. Clin Cancer Res 1997; 3: 891-9.

87. Morrison JG, White P, McDougall S, Firth JW, Woolfrey SG, Graham MA, Greenslade D. Validation of a highly sensitive ICP-MS method for the determination of platinum in biofluids: application to clinical pharmacokinetic studies with oxaliplatin. J Pharm Biomed Anal 2000; 24: 1-10.

88. Ziegler E, Mason HJ, Baxter PJ. Occupational exposure to cytotoxic drugs in two UK oncology wards. Occup Environ Med 2002; 59: 608-12.

89. Turci R, Sottani C, Spagnoli G, Minoia C. Biological and environmental monitoring of hospital personnel exposed to antineoplastic agents: a review of analytical methods. J Chromatogr B Analyt Technol Biomed Life Sci 2003; 789: 169-209.

90. Bettinelli M, Spezia S, Ronchi A, Minoia C. Determination of Pt in biological fluids with ICP-MS: Evaluation of analytical uncertainty. Atom Spectrosc 2004; 25: 103-11.

91. Mason HJ, Blair S, Sams C, Jones K, Garfitt SJ, Cuschieri MJ, Baxter PJ. Exposure to antineoplastic drugs in two UK hospital pharmacy units. Ann Occup Hyg 2005; 49: 603-10.

92. Siim BG, Lee AE, Shalal-Zwain S, Pruijn FB, McKeage MJ, Wilson WR. Marked potentiation of the antitumour activity of chemotherapeutic drugs by the antivascular agent 5,6-dimethylxanthenone-4-acetic acid (DMXAA). Cancer Chemother Pharmacol 2003; 51: 43-52.

93. Johnsson A, Bjork H, Schutz A, Skarby T. Sample handling for determination of free platinum in blood after cisplatin exposure. Cancer Chemother Pharmacol 1998; 41: 248-51.

94. McCurdy E, Potter D. Optimising ICP-MS for the determination of trace metals in high matrix samples. Spectroscopy Europe 2001; 13: 10-3.

95. Hann S, Koellensperger G, Stefanka Z, Stingeder G, Furhacker M, Buchberger W, Mader RM. Application of HPLC-ICP-MS to speciation of cisplatin and its degradation products in water containing different chloride concentrations and in human urine. J Anal Atom Spectrom 2003; 18: 1391-5.

96. Tothill P, Matheson LM, Robbins ME, Hopewell JW, McKay K, Smyth JF. Long-term retention and distribution of platinum in animals after cisplatin administration. Cancer Therapy and Control 1992; 2: 275-87.

97. Minami T, Hashii K, Tateyama I, Kadota E, Tohno Y, Tohno S, Utsumi M, Yamada MO, Ichii M, Namikawa K, . Accumulation of platinum in the intervertebral discs and vertebrae of ovarian tumor-bearing patients treated with cisplatin. Biol Trace Elem Res 1994; 42: 253-7.

98. Minami T, Ichii M, Okazaki Y. Comparison of three different methods for measurement of tissue platinum level. Biol Trace Elem Res 1995; 48: 37-44.

99. Liang X, Huang Y. Physical state changes of membrane lipids in human lung adenocarcinoma A(549) cells and their resistance to cisplatin. Int J Biochem Cell Biol 2002; 34: 1248-55.

100. Ghezzi A, Aceto M, Cassino C, Gabano E, Osella D. Uptake of antitumor platinum(II)-complexes by cancer cells, assayed by inductively coupled plasma mass spectrometry (ICP-MS). J Inorg Biochem 2004; 98: 73-8.

101. Samimi G, Kishimoto S, Manorek G, Breaux JK, Howell SB. Novel mechanisms of platinum drug resistance identified in cells selected for resistance to JM118 the active metabolite of satraplatin. Cancer Chemother Pharmacol 2007; 59: 301-12.

102. Perry BJ, Balazs RE. ICP-MS method for the determination of platinum in suspensions of cells exposed to cisplatin. Analytical Proceedings including Analytical Communications 1994; 31: 269-71.

103. Hanada T, Isobe H, Saito T, Ogura S, Takekawa H, Yamazaki K, Tokuchi Y, Kawakami Y. Intracellular accumulation of thallium as a marker of cisplatin cytotoxicity in nonsmall cell lung carcinoma: an application of inductively coupled plasma mass spectrometry. Cancer 1998; 83: 930-5.

104. Hanada T, Isobe H, Saitoh T, Ogura S, Saito K, Kawakami Y. Inductively coupled plasma mass spectrometry for the determination of platinum accumulation in human non-small cell lung cancer cell lines. Int J Clin Oncol 1998; 3: 98-101.

105. Zoriy M, Matusch A, Spruss T, Becker S. Laser ablation inductively coupled plasma mass spectrometry for imaging of copper, zinc, and platinum in this sections of a kidney from a mouse treated with cisplatin. International Journal of Mass Spectrometry 2007; 260: 102-6.

Chapter 1.2

66

106. Mason HJ, Morton J, Garfitt SJ, Iqbal S, Jones K. Cytotoxic drug contamination on the outside of vials delivered to a hospital pharmacy. Ann Occup Hyg 2003; 47: 681-5.

107. Brouwers EEM, Huitema ADR, Bakker EN, Douma JW, Schimmel KJ, van Weringh G, de Wolf PJ, Schellens JHM, Beijnen JH. Monitoring of platinum surface contamination in seven Dutch hospital pharmacies using inductively coupled plasma mass spectrometry. Int Arch Occup Environ Health 2007; 80:689-99.

108. Wittgen BP, Kunst PW, Perkins WR, Lee JK, Postmus PE. Assessing a system to capture stray aerosol during inhalation of nebulized liposomal cisplatin. J Aerosol Med 2006; 19: 385-91.

109. De Boer JL, Ritsema R, Piso S, Van Staden H, Van Den BW. Practical and quality-control aspects of multi-element analysis with quadrupole ICP-MS with special attention to urine and whole blood. Anal Bioanal Chem 2004; 379: 872-80.

110. Thompson JJ, Houk RS. A study of internal standardisation in inductively coupled plasma-mass spectrometry. Applied Spectroscopy 1987; 41: 801806.

111. Polec-Pawlak K, Abramski JK, Semenova O, Hartinger CG, Timerbaev AR, Keppler BK, Jarosz M. Platinum group metallodrug-protein binding studies by capillary electrophoresis - inductively coupled plasma-mass spectrometry: a further insight into the reactivity of a novel antitumor ruthenium(III) complex toward human serum proteins. Electrophoresis 2006; 27: 1128-35.

112. Gamelin E, Boisdron-Celle M, Lebouil A, Turcant A, Cailleux A, Krikorian A, Brienza S, Cvitkovic E, Larra F, Robert J, Allain P. Determination of unbound platinum after oxaliplatin administration: comparison of currently available methods and influence of various parameters. Anticancer Drugs 1998; 9: 223-8.

113. Todoli JL, Mermet JM. Sample introduction systems for the analysis of liquid microsamples by ICP-AES and ICP-MS. Spectrochimica Acta Part B 2006; 61: 239-83.

114. Zhao Z, Tepperman K, Dorsey JG, Elder RC. Determination of cisplatin and some possible metabolites by ion-pairing chromatography with inductively coupled plasma mass spectrometric detection. J Chromatogr 1993; 615: 83-9.

115. Bell DN, Liu JJ, Tingle MD, McKeage MJ. Specific determination of intact cisplatin and monohydrated cisplatin in human plasma and culture medium ultrafiltrates using HPLC on-line with inductively coupled plasma mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2006; 837: 29-34.

116. Vacchina V, Torti L, Allievi C, Lobinski R. Sensitive species-specific monitoring of a new triplatinum anti-cancer drug and its potential related compounds in spiked human plasma by cation-exchange HPLC-ICP-MS. J Anal Atom Spectrom 2003; 18: 884-90.

117. Cairns WRL, Ebdon L, Hill SJ. Development of an HPLC-ICP-MS method for the determination of platinum species from new anti-tumour drugs. Analytical Proceedings including Analytical Communications 1994; 31: 295-7.

118. Cairns WR, Ebdon L, Hill SJ. A high performance liquid chromatography - inductively coupled plasma-mass spectrometry interface employing desolvation for speciation studies of platinum in chemotherapy drugs. Anal Bioanal Chem 1996; 355: 202-8.

119. Galettis P, Carr L, Paxton JW, McKeage MJ. Quantitative determination of platinum complexes in human plasma generated from the oral antitumour drug JM216 using directly coupled high-performance liquid chromatography-inductively coupled plasma mass spectrometry without desolvation. J Anal Atom Spectrom 1999; 14: 953-6.

120. Allain P, Heudi O, Cailleux A, Le Bouil A, Larra F, Boisdron-Celle M, Gamelin E. Early biotransformations of oxaliplatin after its intravenous administration to cancer patients. Drug Metab Dispos 2000; 28: 1379-84.

121. Wang K, Lu J, Li R. The events that occur when cisplatin encounters cells. Coordination Chemistry Reviews 1996; 151: 53-88.

122. Timerbaev AR, Aleksenko SS, Polec-Pawlak K, Ruzik R, Semenova O, Hartinger CG, Oszwaldowski S, Galanski M, Jarosz M, Keppler BK. Platinum metallodrug-protein binding studies by capillary electrophoresis-inductively coupled plasma-mass spectrometry: characterization of interactions between Pt(II) complexes and human serum albumin. Electrophoresis 2004; 25: 1988-95.

123. Heudi O, Cailleux A, Allain P. Kinetic studies of the reactivity between cisplatin and its monoaquo species with L-methionine. J Inorg Biochem 1998; 71: 61-9.

124. Lecomte T, Landi B, Beaune P, Laurent-Puig P, Loriot MA. Glutathione S-transferase P1 polymorphism (Ile105Val) predicts cumulative neuropathy in patients receiving oxaliplatin-based chemotherapy. Clin Cancer Res 2006; 12: 3050-6.

ICP-MS in oncology

67

125. Stoehlmacher J, Park DJ, Zhang W, Groshen S, Tsao-Wei DD, Yu MC, Lenz HJ. Association between glutathione S-transferase P1, T1, and M1 genetic polymorphism and survival of patients with metastatic colorectal cancer. J Natl Cancer Inst 2002; 19;94: 936-42.

126. Oldenburg J, Kraggerud SM, Cvancarova M, Lothe RA, Fossa SD. Cisplatin-induced long-term hearing impairment is associated with specific glutathione s-transferase genotypes in testicular cancer survivors. J Clin Oncol 2007; 25: 708-14.

127. Khalaila I, Bergamo A, Bussy F, Sava G, Dyson PJ. The role of cisplatin and NAMI-A plasma-protein interactions in relation to combination therapy. Int J Oncol 2006; 29: 261-8.

128. Pluim D, van Waardenburg RC, Beijnen JH, Schellens JH. Cytotoxicity of the organic ruthenium anticancer drug Nami-A is correlated with DNA binding in four different human tumor cell lines. Cancer Chemother Pharmacol 2004; 54: 71-8.

129. Kratz F, Hartmann M, Keppler B, Messori L. The binding properties of two antitumor ruthenium(III) complexes to apotransferrin. J Biol Chem 1994; 269: 2581-8.

130. Messori L, Orioli P, Vullo D, Alessio E, Iengo E. A spectroscopic study of the reaction of NAMI, a novel ruthenium(III)anti-neoplastic complex, with bovine serum albumin. Eur J Biochem 2000; 267: 1206-13.

131. Pongratz M, Schluga P, Jakupec MA, Arion VB, Hartinger CG, Allmaier G, Keppler BK. Transferrin binding and transferrin-mediated cellular uptake of the ruthenium coordination compound KP1019, studied by means of AAS, ESI-MS and CD spectroscopy. J Anal Atom Spectrom 2004; 19: 46-51.

132. Falter R, Wilken RD. Determination of carboplatinum and cisplatinum by interfacing HPLC with ICP-MS using ultrasonic nebulisation. Sci Total Environ 1999; 225: 167-76.

133. Hann S, Stefanka Z, Lenz K, Stingeder G. Novel separation method for highly sensitive speciation of cancerostatic platinum compounds by HPLC-ICP-MS. Anal Bioanal Chem 2005; 381: 405-12.

134. Lenz K, Koellensperger G, Hann S, Weissenbacher N, Mahnik SN, Fuerhacker M. Fate of cancerostatic platinum compounds in biological wastewater treatment of hospital effluents. Chemosphere 2007; epub ahead of print.

135. Zoorob GK, McKiernan JW, Caruso JA. ICP-MS for elemental speciation studies. Mikrochim Acta 1998; 128: 145-68.

136. Garcia Sar D, Montes-Bayon M, Blanco Gonzalez E, Sanz-Medel A. Speciation studies of cisplatin adducts with DNA nucleotides via elemental specific detection (P and Pt) using liquid chromatography-inductively coupled plasma-mass spectrometry and structural characterization by electrospray mass spectrometry. J Anal Atom Spectrom 2006; 21: 861-8.

137. El Khateeb M, Appleton TG, Charles BG, Gahan LR. Development of HPLC conditions for valid determination of hydrolysis products of cisplatin. J Pharm Sci 1999; 88: 319-26.

138. Xie R, Johnson W, Rodriguez L, Gounder M, Hall GS, Buckley B. A study of the interactions between carboplatin and blood plasma proteins using size exclusion chromatography coupled to inductively coupled plasma mass spectrometry. Anal Bioanal Chem 2007; 387: 2815-22.

139. Mandal R, Teixeira C, Li XF. Studies of cisplatin and hemoglobin interactions using nanospray mass spectrometry and liquid chromatography with inductively-coupled plasma mass spectrometry. Analyst 2003; 128: 629-34.

140. Mandal R, Kalke R, Li XF. Interaction of oxaliplatin, cisplatin, and carboplatin with hemoglobin and the resulting release of a heme group. Chem Res Toxicol 2004; 17: 1391-7.

141. Zhao YY, Mandal R, Li XF. Intact human holo-transferrin interaction with oxaliplatin. Rapid Commun Mass Spectrom 2005; 19: 1956-62.

142. Szpunar J, Makarov A, Pieper T, Keppler BK, Lobinski R. Investigation of metallodrug-protein interactions by size-exclusion chromatography coupled with inductively coupled plasma mass spectrometry (ICP-MS). Anal Chim Acta 1999; 387: 135-44.

143. Sulyok M, Hann S, Hartinger CG, Keppler BK, Stingeder G, Koellensperger G. Two dimensional separation schemes for investigation of the interaction of an anticancer ruthenium(III) compound with plasma proteins. J Anal Atom Spectrom 2005; 20: 856-63.

144. Hartinger CG, Hann S, Koellensperger G, Sulyok M, Groessl M, Timerbaev AR, Rudnev AV, Stingeder G, Keppler BK. Interactions of a novel ruthenium-based anticancer drug (KP1019 or FFC14a) with serum proteins--significance for the patient. Int J Clin Pharmacol Ther 2005; 43: 583-5.

145. Stefanka Z, Hann S, Koellensperger G, Stingeder G. Investigation of the reaction of cisplatin with methionine in aqueous media using HPLC-ICP-DRCMS. J Anal Atom Spectrom 2004; 19: 894-8.

146. Morrison JG, Bissett D, Stephens IFD, McKay K, Brown R, Graham MA, Fichtinger-Schepman AM, Kerr DJ. The isolation and identification of cis-diamminedichloroplatinum(II)-DNA adducts by anion exchange HPLC and inductively coupled plasma mass spectrometry. Int J Oncol 1993; 2: 33-7.

Chapter 1.2

68

147. Hann S, Zenker A, Galanski M, Bereuter TL, Stingeder G, Keppler BK. HPIC-UV-ICP-SFMS study of the interaction of cisplatin with guanosine monophosphate. Fresenius J Anal Chem 2001; 370: 581-6.

148. Lustig S, de Kimpe J, Cornelis R, Schramel P. Development of native two-dimensional electrophoresis methods for the separation and detection of platinum carrying serum proteins: initial steps. Fresenius J Anal Chem 1999; 363: 484-7.

149. Morazzoni F, Canevali C, Zucchetti M, Caroli S, Alimonti A, Petrucci F, Giudice G, Masoni E, Bedini AV. cis-Diamminedichloroplatinum(II) given in low-dose continuous infusion with concurrent radiotherapy to patients affected by inoperable lung carcinoma: a pharmacokinetic approach. J Cancer Res Clin Oncol 1998; 124: 37-43.

150. Holmes J, Stanko J, Varchenko M, Ding H, Madden VJ, Bagnell CR, Wyrick SD, Chaney SG. Comparative neurotoxicity of oxaliplatin, cisplatin, and ormaplatin in a Wistar rat model. Toxicol Sci 1998; 46: 342-51.

151. Screnci D, Galettis P, Baguley BC, McKeage MJ. Optimization of an ICP-MS assay for the detection of trace levels of platinum in peripheral nerves. Atom Spectrosc 1998; 19: 172-5.

152. Filippich LJ, Bucher AM, Charles BG. Platinum pharmacokinetics in sulphur-crested cockatoos (Cacatua galerita) following single-dose cisplatin infusion. Aust Vet J 2000; 78: 406-11.

153. Jamieson SM, Liu J, Hsu T, Baguley BC, McKeage MJ. Paclitaxel induces nucleolar enlargement in dorsal root ganglion neurons in vivo reducing oxaliplatin toxicity. Br J Cancer 2003; 88: 1942-7.

154. Kolb RJ, Ghazi AM, Barfuss DW. Inhibition of basolateral transport and cellular accumulation of cDDP and N-acetyl- L-cysteine-cDDP by TEA and PAH in the renal proximal tubule. Cancer Chemother Pharmacol 2003; 51: 132-8.

155. Charlier C, Kintz P, Dubois N, Plomteux G. Fatal overdosage with cisplatin. J Anal Toxicol 2004; 28: 138-40.

156. Filippich LJ, Charles BG, Sutton RH, Bucher AM. Carboplatin pharmacokinetics following a single-dose infusion in sulphur-crested cockatoos (Cacatua galerita). Aust Vet J 2004; 82: 366-9.

157. Ekborn A, Hansson J, Ehrsson H, Eksborg S, Wallin I, Wagenius G, Laurell G. High-dose Cisplatin with amifostine: ototoxicity and pharmacokinetics. Laryngoscope 2004; 114: 1660-7.

158. Smit T, Snyman JR, Neuse EW, Bohm L, van Rensburg CE. Evaluation of cisplatin and a novel platinum polymer conjugate for drug toxicity and drug distribution in mice. Anticancer Drugs 2005; 16: 501-6.

159. Royer B, Guardiola E, Polycarpe E, Hoizey G, Delroeux D, Combe M, Chaigneau L, Samain E, Chauffert B, Heyd B, Kantelip JP, Pivot X. Serum and intraperitoneal pharmacokinetics of cisplatin within intraoperative intraperitoneal chemotherapy: influence of protein binding. Anticancer Drugs 2005; 16: 1009-16.

160. Lardinois D, Jung FJ, Opitz I, Rentsch K, Latkoczy C, Vuong V, Varga Z, Rousson V, Gunther D, Bodis S, Stahel R, Weder W. Intrapleural topical application of cisplatin with the surgical carrier Vivostat increases the local drug concentration in an immune-competent rat model with malignant pleuromesothelioma. J Thorac Cardiovasc Surg 2006; 131: 697-703.

161. Pasqua AJ, Goodisman J, Kerwood DJ, Toms BB, Dubowy RL, Dabrowiak JC. Role of carbonate in the cytotoxicity of Carboplatin. Chem Res Toxicol 2007; 20: 896-904.

162. Pohlen U, Rieger H, Meyer BT, Loddenkemper C, Buhr HJ, Heitland P, Koester HD, Schneider P. Chemoembolization of lung metastases--pharmacokinetic behaviour of carboplatin in a rat model. Anticancer Res 2007; 27: 809-15.

163. Turci R, Sottani C, Ronchi A, Minoia C. Biological monitoring of hospital personnel occupationally exposed to antineoplastic agents. Toxicol Lett 2002; 134: 57-64.

164. Guerbet M, Goulle JP, Lubrano J. Evaluation of the risk of contamination of surgical personnel by vaporisation of oxaliplatin during the intraoperative hyperthermic intraperitoneal chemotherapy (HIPEC). Eur J Surg Oncol 2007; 33:623-6.

165. Smith CJ, wilson ID, Abou-Shakra F, Payne R, Grisedale H, Long A, Roberts D, Malone M. Analysis of a [14C]-labelled platinum anticancer compound in dosing formulations and urine using a combination of HPLC-ICPMS and flow scintillation counting. Chromatographia 2002; 55: S-151-S-155.

166. Heudi O, Brisset H, Cailleux A, Allain P. Chemical instability and methods for measurement of cisplatin adducts formed by interactions with cysteine and glutathione. Int J Clin Pharmacol Ther 2001; 39: 344-9.

167. Azim-Araghi A, Ottley CJ, Pearson DG, Tilby MJ. Combination of ICP-MS and ion exchange chromatography permits analyses of specific cisplatin DNA modifications formed in cancer cells. In: Plasma Source Mass Spectrometry. Holland JG, Tanner SD, eds. Holland JG, Tanner SD, eds.: Royal Society of Chemistry, 2003: 66-73

ICP-MS in oncology

69

168. Mandal R, Kalke R, Li XF. Mass spectrometric studies of cisplatin-induced changes of hemoglobin. Rapid Commun Mass Spectrom 2003; 17: 2748-54.

169. Mandal R, Jiang G, Li XF. Direct evidence for co-binding of cisplatin and cadmium to a native zinc- and cadmium-containing metallothionein. Appl Organometal Chem 2003; 17: 675-81.

170. Meczes EL, Azim-Araghi A, Ottley CJ, Pearson DG, Tilby MJ. Specific adducts recognised by a monoclonal antibody against cisplatin-modified DNA. Biochem Pharmacol 2005; 70: 1717-25.

171. Mandal R, Sawyer MB, Li XF. Mass spectrometry study of hemoglobin-oxaliplatin complexes in colorectal cancer patients and potential association with chemotherapeutic responses. Rapid Commun Mass Spectrom 2006; 20: 2533-8.

172. Timerbaev AR, Foteeva LS, Rudnev AV, Abramski JK, Polec-Pawlak K, Hartinger CG, Jarosz M, Keppler BK. Probing the stability of serum protein-ruthenium(III) drug adducts in the presence of extracellular reductants using CE. Electrophoresis 2007; 28: 2235-40.

173. Crul M, van Waardenburg RC, Beijnen JH, Schellens JH. DNA-based drug interactions of cisplatin. Cancer Treat Rev 2002; 28: 291-303.

Chapter 2

Determination of platinum and ruthenium in biological fluids

Chapter 2.1

Determination of oxaliplatin in human plasma and plasma ultrafiltrate by

graphite-furnace atomic-absorption-spectrometry

Elke E.M. Brouwers Matthijs M. Tibben

Markus Joerger Olaf van Tellingen

Hilde Rosing Jan H.M. Schellens

Jos H. Beijnen

Analytical and Bioanalytical Chemistry 2005; 382; 1484-1490

Chapter 1.2

74

Abstract

A method for the sensitive determination of the anticancer agent oxaliplatin in human

plasma and human plasma ultrafiltrate is presented. The method is based on the

quantification of platinum (Pt) by graphite-furnace atomic-absorption-spectrometry,

with Zeeman correction and an atomisation temperature of 2700 °C. Sample

pretreatment involves dilution of the samples with a solution containing 0.15 M NaCl

and 0.20 M HCl in water. Validation was performed in accordance with the most recent

FDA guidelines for bioanalytical method validation. All results were within requirements.

The validated ranges of quantification were 19.5-1.95x104 µg/L Pt for human plasma

ultrafiltrate and 97.5-1.95x104 µg/L Pt for plasma. The assay is now successfully used to

support pharmacokinetic studies in cancer patients treated with oxaliplatin.

Determination of platinum by GF-AAS

75

Introduction

Effective platinum (Pt) anticancer agents, for example cisplatin, were discovered in the

1960s [1]. Clinical use of cisplatin is, however, often accompanied by severe side effects,

such as nephrotoxicity and neurotoxicity, and some cancers have intrinsic resistance to

cisplatin or develop resistance during cisplatin treatment [2]. The search for Pt

anticancer agents with less severe side effects and increased efficacy has led to the

development of several Pt-based compounds, including oxaliplatin [(1R,-2R)-1,2-

cyclohexanediamine-N,N´] [oxalato (2-)-O,O´]platinum; Figure 1), which differs from

cisplatin by the presence of a diaminocyclohexane (DACH) ligand and an oxalato group

in its chemical structure. The drug was first introduced into clinical trials in 1986 [3] and

has preclinical and clinical activity against a wide variety of tumour types, including

colorectal cancer [4-6].

Figure 1. Chemical structure of oxaliplatin

Currently, oxaliplatin (Eloxatin®) is part of the standard first-line treatment in patients

with colorectal cancer. Quantitative determination of the drug in clinical samples is a

prerequisite in determining the pharmacokinetics of this Pt-based agent. Atomic

absorption spectrometry (AAS) is one of the most widely used spectrometric techniques

for specific determination of elements, including Pt, in biological materials. Graphite-

furnace atomic-absorption-spectrometry (GF-AAS), has good sensitivity and needs a

sample volume of only 20 µL. GF-AAS methods have been developed and validated for

the determination of Pt in biological fluids for patients receiving carboplatin [7,8],

cisplatin [9], cisplatin in a liposomal source, SPI-77 [10], polymer bound Pt AP5280 [11]

and the orally available Pt derivative, JM216 [12]. Although oxaliplatin has been used

since 1986 and has previously been analysed by use of GF-AAS [13-16], to the best of our

knowledge no GF-AAS method validation has been described for this drug. In this and

previous GF-AAS method development we observed that the ligands around the Pt

atom can affect quantification of the Pt compound using atomic absorption

spectrometry. Hence, a GF-AAS method should be validated for each Pt compound

individually, including oxaliplatin.

For analysis of Pt in biological matrices it is necessary to discriminate between bound

and free Pt in blood. Similar to other Pt compounds, oxaliplatin rapidly becomes

partitioned into protein-bound plasma Pt, tissue Pt, erythrocyte-sequestered Pt, and free

plasma Pt [17,18]. The free, ultrafilterable Pt is the active compound, whereas Pt bound

Pt

O

O

O

O

NH

2

NH2

Chapter 1.2

76

to plasma proteins and erythrocytes is usually regarded as pharmacologically inactive

[18].

To analyse the total amount of Pt in the plasma and the unbound active Pt we

developed a simple and sensitive GF-AAS assay for determination of Pt derived from

oxaliplatin in human plasma and human plasma ultrafiltrate (pUF).

Plasma and pUF samples were diluted with an appropriate diluent to reduce

contamination of the GF. As a compromise between good detection limits and good

robustness of the method, we chose to dilute tenfold as standard. We also validated the

possibility of diluting pUF samples twofold to maximise the quantification range of Pt in

pUF. In contrast with GF-AAS methods for other Pt compounds it was necessary to

include extra blank solutions in the sequences to prevent effects resulting from carry-

over of oxaliplatin. The method described here was validated in accordance with the

most recent FDA guidelines on bioanalytical method validation [19]: its implementation

into clinical pharmacokinetic studies is also described.

Experimental

Chemicals

Oxaliplatin reference standard used for the preparation of calibration and quality-control

(QC) samples was generously provided by Sanofi-Synthelabo, Malvern, PA, USA.

Chloroplatinic acid, containing 1,000 mg/L Pt in 3.3% hydrochloric acid (HCl), used for

the comparative analyses with oxaliplatin, was obtained from Inorganic Ventures/IV

Labs, Lakewood, NJ, USA. HCl 37% and sodium chloride (NaCl) were purchased from

Merck (Darmstadt, Germany) and 5% glucose solution was obtained from B. Braun.

Triton X-100 originated from BDH (Poole, UK). Deionised water, produced by use of a

Milli-Q plus system (Millipore, Milford, MA, USA) was used throughout the analyses.

Drug-free human plasma was obtained from the Central Laboratory for Blood

Transfusion (Sanquin Amsterdam, The Netherlands). Hoek Loos (Schiedam, The

Netherlands) provided argon gas.

Preparation of reagents

A modifier solution was prepared by diluting Triton X-100 (1:199 v/v) with deionised

water. A solution, for primary sample dilutions, consisting of 0.15 M NaCl and 0.20 M HCl

was also prepared (NaCl/HCl solution). Both solutions were stored at room temperature

and were stable for at least six months.

Determination of platinum by GF-AAS

77

Preparation of stock solutions, calibration standards, and quality control samples

Two stock solutions of 390 mg/L Pt (corresponding to 795 mg/L oxaliplatin) were

prepared by independent weighing in 5% glucose solution. One solution was further

diluted to a concentration of 19.9 mg/L and served as a working solution for the

calibration standards. The other solution was further diluted to a concentration of 117

mg/L to obtain a working solution for the QC samples. Working solutions were stored at

2-8 °C.

The 19.9 mg/L working solution was further diluted with 1:9 (v/v) Pt-free plasma or pUF

and NaCl/HCl solution to obtain calibration standards at seven concentrations ranging

from 9.75-390 µg/L Pt, corresponding to 97.5-3,90x103 µg/L Pt in the undiluted matrix.

From the calibration standards, the 390 µg/L standard was divided into 1,200 µL

portions and the remaining standards into 200 µL portions. The standards were stored at

–20 °C. The 117 mg/L working solution was used to spike Pt-free plasma and pUF to

obtain QC samples at four concentrations (97.5, 293, 975, and 3,51x103 µg/L Pt). Two

additional levels were also prepared. The first, used for pUF, was prepared to validate the

ability to quantify samples below the lower limit of quantitation (<LLOQ; 19.5 µg/L in

pUF). The second, used for pUF and plasma, was prepared to validate the ability to

quantify samples originally exceeding the upper limit of quantification (>ULOQ; 1.95x104

µg/L). Samples (120 µL) of the QC samples were stored at –20 °C.

Before analysis, calibration standards and QC samples were thawed and processed at

room temperature. Calibration standards were analysed without further dilution. QC

samples were diluted with NaCl/HCl solution (19.5 µg/L; twofold diluted, 97.5-3.51x103

µg/L; tenfold diluted, 1.95x104 µg/L; tenfold diluted with NaCl/HCl solution and then

diluted twentyfold with pUF:NaCl/HCl solution or plasma:NaCl/HCl solution (1:9 v/v)).

Sample processing

Whole blood samples were collected in 10 mL heparin-containing tubes. Plasma was

obtained by centrifuging the whole blood samples for 5 min (1,000 g, 4 °C). PUF was

obtained by centrifuging the plasma through an ultrafiltrate filter (Centriplus YM-30

Millipore, Catnr. 4422) for 15 min (1,000 g, 20 °C). Preparation of pUF from plasma was

performed immediately after blood collection, to prevent any decrease of free Pt levels

as a result of progressive ex vivo binding of Pt to plasma proteins and erythrocytes. All

samples were stored at –20 °C until analysis. Before analysis, pUF and plasma samples

were thawed and vortex mixed and subsequently diluted tenfold with NaCl/HCl

solution. When the Pt concentration in a pUF sample was known or expected to be

below the LLOQ (97.5 µg/L), however, the sample was diluted two-fold with NaCl/HCl

solution. When the Pt concentration in a pUF or plasma sample was above the ULOQ,

Chapter 1.2

78

successive dilution with pUF:NaCl/HCl solution or plasma:NaCl/HCl solution (1:9) was

performed. After dilution of the samples, aliquots of at least 120 µL of diluted samples

were placed into the autosampler vials.

Instrumentation

A Thermo Electron Solaar MQZ graphite-furnace spectrophotometer with Zeeman

correction (Thermo Electron, Cambridge, England) equipped with a FS95 sample

dispenser (Thermo Electron) and a graphite tube atomiser (Thermo Electron) were used.

Small adjustments to the autosampler were made for sampling from vials sealed with

Parafilm (American National Can, Greenwich, CT, USA) to prevent evaporation of the

solutions upon standing in the autosampler tray. The standard Teflon sampler tip was

replaced by a sharp stainless steel needle (25 gauge), capable of penetrating the sealed

vials.

The Pt hollow-cathode lamp was operated with a current of 12 mA with a

monochromator slit width of 0.2 nm. Absorbances were measured at a wavelength of

265.9 nm. Wall atomisation was used. The temperature program of the instrument

(Table 1) comprised a drying stage (steps 1 and 4), an ashing stage (5 and 6), an

atomisation stage (7), and a stage of burning-clean with cooling (8 and 9). Cooling of the

graphite tube atomiser was performed using a M33 recirculation cooling device (Thermo

Neslab, Portsmouth, NH, USA). During the atomisation stage absorbance was monitored

using Zeeman correction. The inert carrier gas argon was used to purge the graphite

tube at a flow rate of 0.3 L/min. The gas flow was turned off during the atomisation

stage.

A total of 30 µL of fluid was introduced into the graphite tube. First, 5 µL modifier

solution (0.5% (v/v) Triton X-100 in water) was pre-injected to reduce the surface tension

of the sample, followed by 20 µL of sample and 5 µL of diluent. Each sample was

analysed in duplicate and the absorbance readings were averaged. The total

measurement time for one duplicate sample was 11 min. After every ten samples the

blank and highest calibration standard were re-assayed to re-calculate the slope of the

calibration plot. This was done to correct for the decrease in atomisation efficiency of the

graphite furnace during an analytical run. This correction was only allowed when the

response decreased less than 10% compared to the previous measurement of this

standard. Otherwise, the graphite tube was regarded as defective and was replaced by a

new one. The pyrolytically-coated partitioned graphite tubes (Extended life cuvettes,

Thermo Electron) were routinely replaced after 750 firings.

Data were acquired using the Spectrometer Software version 1.21 (SOLAAR House,

Cambridge, UK) and processed (integrated) using SOLAAR Data Station version 9.12

Determination of platinum by GF-AAS

79

software (SOLAAR House). Further data handling was performed using Excel 2000

(Microsoft, Redmond, WA, USA).

Table 1. Temperature program GF-AAS method for oxaliplatin

Step Temp (°C) Time (s)* Ramp (°C/s)** Gas Flow (L/min)

1 50 1.0 0 0.3

2 85 5.0 17 0.3

3 95 30.0 1 0.3

4 120 20.0 2 0.3

5 250 30.0 5 0.3

6 1400 40.0 30 0.3

7 2700 3.0 0 0

8 2800 4.0 0 0.3

9 50 10.0 0 0.3 *

Time (s) is the time the temperature remains constant ,**

Ramp (°C/s) is the velocity at which the temperature is reached

The use of Pt standard instead of oxaliplatin standard

We tested the possibility of using of chloroplatinic acid instead of oxaliplatin as a

standard for the preparation of calibration standards and quality control samples. QC

samples were therefore prepared at three concentrations (58.5, 3.51x103, and 1.95x104

µg/L Pt) using certified chloroplatinic acid and oxaliplatin reference standards. These

samples were processed and Pt levels were quantified using calibration standards spiked

with chloroplatinic acid. Calibration standards were processed and analysed in singly.

QC samples were diluted and analysed in fivefold.

Validation procedures

Full validation in accordance with the FDA guidelines [19] was performed for the assay in

human pUF. For the assay in plasma a partial validation was performed. According to

the FDA guidelines a partial validation is sufficient to test the method when only a

change in matrix, with the same species is concerned.

Re-calibrations were performed after every ten samples throughout sample analysis and

duplicate analyses were performed on each individual sample.

Chapter 1.2

80

Limit of quantification

The limit of detection (LOD) was determined by use of a signal-to-noise ratio of 3. The

analyte response at the LLOQ should be at least 5 times the response for a blank sample.

The LLOQ should be determined with a precision better than 20% and the mean value

should deviate from the actual value by no more than 20% [19].

Linearity

Seven non-zero calibration standards were processed and analysed singly for plasma

and pUF in one and three separate analytical runs, respectively. Concentrations were

back-calculated from the corresponding calibration plot. Deviations from the nominal

concentrations should be within ±20% for the LLOQ and within ±15% for other

concentrations [19].

Accuracy and precision

The accuracy and within-run and between-run precision of the method were

determined by assaying QC samples at different concentrations and with different

dilution factors, for both matrices. Five replicates of each sample were analysed in one

and three analytical runs, for plasma and pUF respectively. QC samples were analysed

together with independently prepared calibration standards. The accuracy was

determined as the percentage of the nominal concentration. The accuracy should be

within 80-120% of the nominal concentration for the LLOQ and within 85-115% of the

nominal concentrations for the other concentrations. Within-run and between-run

precision were calculated by analysis of variances (ANOVA) for each test concentration

using the analytical run as the grouping variable. The precision should not exceed ±20%

for the LLOQ and ±15% for the other concentrations [19].

Specificity

The specificity of the method was assessed by analysis of six individual batches of

control drug-free human pUF and plasma, both analysed blank and spiked at LLOQ level.

Samples were processed according to the procedures described above and analysed in

one run. The accuracy for samples spiked at the LLOQ should be within 80-120% of the

nominal value [19]. GF-AAS absorbance peak heights of the blanks and LLOQ samples

were monitored and compared for spectrometric integrity and potential interferences.

The peak heights for blank matrix samples should not exceed 20% of peak heights at the

LLOQ level.

Determination of platinum by GF-AAS

81

Stability

For evaluation of the stability of Pt concentrations during storage and sample

processing in the (un)diluted matrix, two QC and calibration concentrations were

sampled for each matrix. Long-term storage stability was assessed by determining the Pt

concentration in the two QC and calibration samples before and after 6 months storage

at -20 °C.

The stability in human plasma and pUF after three freeze (-20 °C)/thaw cycles was

investigated by comparing results from QC and calibration samples that have been

frozen and thawed three times with results from freshly prepared samples.

Furthermore, short-term stability of the analyte in human plasma and pUF was

evaluated by comparing results from QC and calibration samples that had been stored

for 24 h at ambient temperatures with results from freshly prepared samples. This was

done to determine the stability of the analyte in the samples during sample preparation

and analysis. The analyte is considered stable in the (un)diluted biological matrix when

85-115% of the initial concentration is recovered.

Finally, the stability was evaluated in stock and working solutions under storage

conditions. The analyte is considered stable in the stock and working solutions when 95-

105% of the original concentration is recovered.

Carry-over

To establish the effect of carry-over, we observed the signals of blank readings after the

most concentrated QC sample (3.51x103 µg/L Pt) and the most concentrated calibration

standard (3.90x103 µg/L Pt). The response of the blank should be less than 20% of the

response of the LLOQ standard.

Application of the GF-AAS assay

The analytical method described in this paper is used to support clinical

pharmacokinetic studies. An example of the analysis of plasma and pUF samples of a

patient treated with oxaliplatin 130 mg/m2 as a 2 h infusion, is given here. Plasma

samples were taken up to 3 h after administration of oxaliplatin and pUF was obtained

as described above. Samples were diluted with NaCl/HCl solution and, if necessary,

diluted further to fit them within the calibration range.

Chapter 1.2

82

Results and discussion

The use of Pt standard instead of oxaliplatin standard

Mean concentrations in QC samples prepared from chloroplatinic acid and from

oxaliplatin are listed in Table 2. Use of the Student t-test to compare concentrations in

QC samples prepared from the oxaliplatin standard and the Pt standard yielded P-values

below the critical value of 0.05 for all three QC concentrations. The significant difference

may have been because of the different molecular structures of the compounds. Despite

atomisation, the coordination of the ligands around the central Pt atom may have

affected the absorbance of the analyte. We therefore conclude that it was not possible

to replace the oxaliplatin standard by the Pt standard used here.

Table 2. Assay results of QC samples spiked with chloroplatinic acid and oxaliplatin (n=5)

Nominal concentration (µg/L)

Mean Pt concentration from chloroplatinic acid QC sample (µg/L)

DEV (%) from nominal concentration

Mean Pt concentration from oxaliplatin QC sample (µg/L)

DEV (%) from nominal concentration

58.5 61.6 5.3 78.0 33.3

3.51x103 3.52x103 0.3 3.84x103 9.3

1.95x104 1.97x104 1.2 2.23x104 14.3

Validation

Limit of quantification

The LODs were 9.75 µg/L Pt in pUF (when diluted twofold with NaCl/HCl solution) and

41.0 µg/L Pt in plasma (when diluted tenfold with NaCl/HCl solution). The LLOQ of the

assay were set at a Pt concentration of 19.5 µg/L for pUF and 97.5 µg/L for plasma,

which, after dilution, correspond to the lowest Pt standard concentration in pUF and

plasma. The acceptance criteria for the LLOQ were easily met.

Linearity

Calibration standards were analysed in the dynamic range 9.75-390 µg/L Pt

(corresponding with 97.5-3.9x103 µg/L oxaliplatin in undiluted matrix), singly in one and

three analytical runs for plasma and pUF respectively. A representative calibration plot

for oxaliplatin in pUF is depicted in Figure 2. Typically, the concentration-response curve

Determination of platinum by GF-AAS

83

obtained by GF-AAS was hyperbolic in shape. The calibration curve was therefore best

described by quadratic regression.

0.00.10.20.30.40.50.60.7

0 1000 2000 3000 4000 5000

Platinum concentration (ug/L)

Sign

al p

eak

heig

ht

Figure 2. Representative calibration curve for oxaliplatin in pUF

The calibration concentrations were back-calculated from the responses. Deviations and

relative standard deviations are listed Table 3. Deviations from the nominal

concentration were between –4.3 and 4.1% for all concentrations in pUF, and between –

8.6 and 2.0% in plasma. Relative standard deviations for the calibration samples in pUF

were up to 1.8%. No relative standard deviations were determined for plasma, because

only one analytical run was performed.

Table 3. Mean deviation from theoretical concentration (DEV %) for oxaliplatin standards in human

plasma and pUF and the relative standard deviation (RSD %) (n=3) for oxaliplatin standards in pUF.

Nominal Pt concentration (µg/L)

Mean Pt concentration in plasma (µg/L)

DEV (%) from nominal concentration

Mean Pt concentration in pUF (µg/L)

DEV (%) from nominal concentration

RSD (%)

98.1 89.7 -8.6 93.8 -4.3 1.2

196 186 -5.2 204 4.1 1.8

490 486 -0.9 489 -0.2 1.3

981 932 -5.0 973 -0.8 1.8

1.96x103 2.00x103 2.0 1.98x103 1.2 1.3

2.94x103 2.90x103 -1.4 2.92x103 -0.8 0.9

3.92x103 3.91x103 -0.4 3.93x103 0.2 0.3

Chapter 1.2

84

Accuracy and precision

Within-run and between-run precision data are summarised in Table 4. For plasma, only

within-run precision data were determined, because partial validation was performed.

Accuracy was within 85.2 and 105.8% and within 93.0 and 111.0%, for pUF and plasma

respectively. This showed that for all QC concentration levels, data were within generally

accepted criteria for bioanalytical method validation [19] and that the different dilution

of the QC samples did not affect the performance of the method.

Table 4. Within-run and between-run precision data in human plasma and pUF

Initial concentration (µg/L)

Final concentration (µg/L)

Matrix dilution

Within-run precision plasma

Within-run precision pUF

Between-run precision pUF

19.5 9.75 1:2 * 3.9 7.7

97.5 9.75 1:10 1.1 6.4 10.1

293 29.3 1:10 0.4 1.5 1.6

975 97.5 1:10 4.9 1.1 1.1

3.51x103 351 1:10 4.0 1.3 **

1.95x104 97.5 1:200 4.3 5.5 **

*0.099 was not used as a QC sample in plasma validation

**no statistically significant variation was observed as a result of performing the assay in different runs

Specificity

Analysis of blank samples from six individual batches of pUF and plasma did not show

interferences from endogenous material at the absorbance wavelength of Pt with peaks

>20% of the LLOQ heights. Deviations from the nominal concentrations at the LLOQ

level were between –6.3 and 3.9% for pUF and between 2.8 and 19.8% for plasma.

Stability

When preparing the concentrated stock and working solutions it was important to avoid

chloride-containing diluents, because these can result in chemical modification into

chloride-containing transformation products [20-22] and the formation of a precipitate

[23]. We decided to use 5% glucose solution as the diluent for the concentrated solution,

which is also used as the oxaliplatin diluent in intravenous infusions. No precipitate was

formed in solutions in 5% glucose with Pt concentrations up to 390 mg/L. Pt levels in the

stock and working solutions in 5% glucose solution remained constant under storage

conditions for at least 6 months.

Determination of platinum by GF-AAS

85

For the less concentrated calibration standards and diluted QC samples, the presence of

chloride in NaCl/HCl solution did not affect the Pt concentration. Solutions were

physically stable and no precipitates were formed. The Pt levels in (un)diluted pUF and

plasma remained constant under all the conditions tested. No significant changes in

concentrations were observed after three freeze/thaw cycles or after 24 h at ambient

temperatures. Stability in (un)diluted pUF and plasma during storage was established for

at least 6 months and further testing is still ongoing.

Carry-over

After the most concentrated QC sample or calibration standard the blank response was

occasionally observed to be higher than 20% of the response of the LLOQ standard. This

phenomenon has not been reported for any of other GF-AAS method developed for Pt

anticancer agents (carboplatin, cisplatin, SPI-77, AP5280 and JM216). Although

speculative, a possible explanation for this carry-over could be the different ligands

around the Pt atom in oxaliplatin, which have different affinities for the sample

introduction system. Further research is needed to confirm this, however.

To prevent the carry-over from affecting sample readings, additional blank readings

were necessary after every high concentration QC sample and calibration standard. This

procedure resulted in any residual Pt being washed out of the system. Human plasma

and pUF samples were analysed in order of increasing concentration, to limit the effects

of carry-over.

Application of the GF-AAS assay

The clinical applicability of the assay was demonstrated by analysis of human pUF

samples from a patient who received 130 mg/m2 oxaliplatin administered as a 2 h

infusion. Plasma samples were collected before administration of oxaliplatin and up to 3

h after infusion. Pt concentration versus time profiles for oxaliplatin in human plasma

and pUF are presented in Figure 3. Up to 3 h after infusion the Pt concentrations were

higher than the LLOQ of 19.5 and of 97.5 µg/L for pUF and plasma respectively.

Conclusion

In conclusion, a GF-AAS assay has been developed for reliable, quantitative

determination of Pt originating from oxaliplatin in human plasma and pUF. The method

was subsequently validated in accordance with current FDA guidelines. The validated

ranges of determination were 97.5-1.95x104 µg/L for plasma and 19.5-1.95x104 µg/L for

Chapter 1.2

86

pUF. The assay is now successfully applied in pharmacokinetic studies of patients being

treated with oxaliplatin.

0

1000

2000

3000

4000

0 1 2 3 4 5Time (h)

Plat

inum

con

cent

ratio

n (µ

g/L) Plasma

PUF

Figure 3. Pt concentration versus time profiles in human plasma and pUF after administration of

oxaliplatin administered as a 2 h infusion at a dose level of 130 mg/m2

References 1. Rosenberg B, Vancamp L, Krigas T. Inhibition of cell division in escherichia coli by electrolysis products

from a platinum electrode. Nature 1965; 205: 698-9.

2. Kelland LR, Farrell NP. Platinum-Based Drugs in Cancer Therapy.: Humana Press, Totowa, NJ, 2000

3. Mathé G, Kidani Y, Triana K, Brienza S, Ribaud P, Goldschmidt E, Ecstein E, Despax R, Musset M, Misset JL. A phase I trial of trans-1-diaminocyclohexane oxalato-platinum (l-OHP). Biomed Pharmacother 1986; 40: 372-6.

4. Raymond E, Chaney SG, Taamma A, Cvitkovic E. Oxaliplatin: a review of preclinical and clinical studies. Ann Oncol 1998; 9: 1053-71.

5. Pendyala L, Creaven PJ. In vitro cytotoxicity, protein binding, red blood cell partitioning, and biotransformation of oxaliplatin. Cancer Res 1993; 53: 5970-6.

6. Schmidt W, Chaney SG. Role of carrier ligand in platinum resistance of human carcinoma cell lines. Cancer Res 1993; 53: 799-805.

7. van Warmerdam LJC, van Tellingen O, Maes RAA, Beijnen JH. Validated method for the determination of carboplatin in biological fluids by Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 1995; 351: 777-81.

8. Kloft C, Appelius H, Siegert W, Schunack W, Jaehde U. Determination of platinum complexes in clinical samples by a rapid flameless atomic absorption spectrometry assay. Ther Drug Monit 1999; 21: 631-7.

9. LeRoy AF, Wehling ML, Sponseller HL, Friauf WS, Solomon RE, Dedrick RL, Litterst CL, Gram TE, Guarino AM, Becker DA. Analysis of platinum in biological materials by flameless atomic absorption spectrophotometry. Biochem Med 1977; 18: 184-91.

10. Meerum Terwogt JM, Tibben MM, Welbank H, Schellens JHM, Beijnen JH. Validated method for the determination of platinum from a liposomal source (SPI-77) in human plasma using graphite furnace Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 2000; 366: 298-302.

11. Tibben MM, Rademaker-Lakhai JM, Rice JR, Stewart DR, Schellens JHM, Beijnen JH. Determination of total platinum in plasma and plasma ultrafiltrate, from subjects dosed with the platinum-containing N-(2-hydroxypropyl)methacrylamide copolymer AP5280, by use of graphite-furnace Zeeman atomic-absorption spectrometry. Anal Bioanal Chem 2002; 373: 233-6.

Determination of platinum by GF-AAS

87

12. Vouillamoz-Lorenz S, Bauer J, Lejeune F, Decosterd LA. Validation of an AAS method for the determination of platinum in biological fluids from patients receiving the oral platinum derivative JM216. J Pharm Biomed Anal 2001; 25: 465-75.

13. Merkel U, Wedding U, Roskos M, Hoffken K, Hoffmann A. Pharmacokinetics of oxaliplatin during chronomodulated infusion in metastatic gastrointestinal cancer patients: a pilot investigation with preliminary results. Exp Toxicol Pathol 2003; 54: 475-9.

14. Bastian G, Barrail A, Urien S. Population pharmacokinetics of oxaliplatin in patients with metastatic cancer. Anticancer Drugs 2003; 14: 817-24.

15. Delord JP, Umlil A, Guimbaud R, Gregoire N, Lafont T, Canal P, Bugat R, Chatelut E. Population pharmacokinetics of oxaliplatin. Cancer Chemother Pharmacol 2003; 51: 127-31.

16. Massari C, Brienza S, Rotarski M, Gastiaburu J, Misset JL, Cupissol D, Alafaci E, Dutertre-Catella H, Bastian G. Pharmacokinetics of oxaliplatin in patients with normal versus impaired renal function. Cancer Chemother Pharmacol 2000; 45: 157-64.

17. Extra JM, Marty M, Brienza S, Misset JL. Pharmacokinetics and safety profile of oxaliplatin. Semin Oncol 1998; 25: 13-22.

18. Graham MA, Lockwood GF, Greenslade D, Brienza S, Bayssas M, Gamelin E. Clinical pharmacokinetics of oxaliplatin: a critical review. Clin Cancer Res 2000; 6: 1205-18.

19. U.S.Food and Drug Administration, Center for Drug Evaluation and Research, Guidance for Industry, Bioanalytical Method Validation. FDA. U.S. Food and Drug Administration: Center for Drug Evaluation and Research: Guidance for Industry: Bioanalytical Method Validation. http://www fda gov/cder/guidance/ 4252fnl pdf 2001.

20. Verstraete S, Heudi O, Cailleux A, Allain P. Comparison of the reactivity of oxaliplatin, pt(diaminocyclohexane)Cl2 and pt(diaminocyclohexane1)(OH2)2(2+) with guanosine and L-methionine. J Inorg Biochem 2001; 84: 129-35.

21. Jerremalm E, Hedeland M, Wallin I, Bondesson U, Ehrsson H. Oxaliplatin degradation in the presence of chloride: identification and cytotoxicity of the monochloro monooxalato complex. Pharm Res 2004; 21: 891-4.

22. Hann S, Stefanka Z, Lenz K, Stingeder G. Novel separation method for highly sensitive speciation of cancerostatic platinum compounds by HPLC-ICP-MS. Anal Bioanal Chem 2005; 381: 405-12.

23. Levi F, Metzger G, Massari C, Milano G. Oxaliplatin: pharmacokinetics and chronopharmacological aspects. Clin Pharmacokinet 2000; 38: 1-21.

Chapter 2.2

Sensitive inductively coupled plasma mass spectrometry assay for the determination

of platinum originating from cisplatin, carboplatin, and oxaliplatin in human

plasma ultrafiltrate

Elke E.M. Brouwers Matthijs M. Tibben

Hilde Rosing Michel J.X. Hillebrand

Markus Joerger Jan H.M. Schellens

Jos H. Beijnen

Journal of Mass Spectrometry 2006; 41; 1186-1194

Chapter 2.2

90

Abstract

We present a highly sensitive, rapid method for the determination of platinum (Pt)

originating from the anticancer agents cisplatin, carboplatin, and oxaliplatin in human

plasma ultrafiltrate. The method is based on the quantification of Pt by inductively

coupled plasma mass spectrometry and allows quantification of 7.50 ng/L Pt in only 150

µL of matrix. Sample pretreatment involves dilution of samples with 1% HNO3. Validation

fulfilled the most recent FDA guidelines for bioanalytical method validation. Validated

ranges of quantification were 7.50 to 1.00x105 ng/L in plasma ultrafiltrate for all three Pt

compounds. The assay is now successfully used to support pharmacokinetic studies in

cancer patients treated with cisplatin, carboplatin, or oxaliplatin.

Determination of platinum by ICP-MS

91

Introduction

Platinum (Pt) anticancer drugs are an important class of chemotherapeutics. The three Pt

compounds currently used on a large scale in the treatment of cancer patients are

cisplatin (cis-diaminedichloridoplatinum(II)), carboplatin (cis-diammine(1,1-cyclobutane

dicarboxylato)platinum(II)) and oxaliplatin ([(1R,-2R)-1,2-cyclohexanediamine-N,N´]

[oxalato(2-)-O,O´]platinum).

The ability to measure Pt exposure in biological matrices is a prerequisite in under-

standing the pharmacokinetics of Pt anticancer agents. Literature describing the early

pharmacokinetics, distribution, and elimination of Pt after treatment with a Pt anticancer

agent is extensive. However, much less has been reported on the long-term retention of

these drugs. This is because of the fact that many studies rely on atomic absorption

spectrometry (AAS) for the analysis of Pt, which is not sensitive enough for evaluating

long-term Pt retention. Because a drug like cisplatin has a high curative potential and its

long-term side effects could possibly be associated with prolonged retention of Pt in the

body [1-4], it is important to be able to describe these long-term pharmacokinetics.

Inductively coupled plasma mass spectrometry (ICP-MS) greatly increases the

pharmacokinetic timescale that can be studied owing to its extremely high sensitivity

and specificity. Since the development of the first ICP-MS in 1975 [5], considerable

changes have been made in the design and performance of the ICP-MS, and the

instrument has proven its applicability in the determination of Pt originating from

anticancer agents in biological fluids of patients treated with these drugs. Methods have

been described for the analysis of Pt in plasma [6-12], plasma ultrafiltrate (pUF) [7-12],

serum [13], blood [10,14], urine [6,9,11,12], and cerebrospinal fluid [13].

In order to assess the long-term exposure to ultrafilterable circulating Pt, we developed

a highly sensitive, rapid ICP-MS assay for the determination of Pt originating from

cisplatin, carboplatin, and oxaliplatin in pUF. PUF contains the pharmacologically active

Pt fraction, and is, therefore, essential in clinical pharmacologic studies with Pt anti-

tumour agents [15,16]. The origin of Pt in pUF, however remains to be further

investigated [8]. To the best of our knowledge, the method described here is one of the

most sensitive methods on the analysis of Pt in pUF using a conventional ICP-

quadrupole-MS [7-12].

For this method, pUF samples were diluted with an appropriate diluent to reduce matrix

effects and contamination of the sample introduction system. As a compromise

between a low detection limit of the method and minimal contamination of the sample

introduction system, we diluted the samples 100-fold. However, to maximise the limit of

quantification of the method, we also validated a tenfold dilution of pUF samples,

enabling the quantification of samples with Pt concentrations below the concentration

of the lower limit of quantification (LLOQ) after a 100-fold dilution. The relatively high

sample volume needed for a single ICP-MS measurement (1.5 mL) and the limited

Chapter 2.2

92

availability of pUF from one patient sample (typically 150 µL) hampered the validation

of even lower dilution factors.

In this article, we describe the validation of the method, according to most recent FDA

guidelines on bioanalytical method validation [17], and its implementation into clinical

pharmacokinetic studies.

Experimental

Chemicals

Cisplatin and carboplatin reference standards were purchased from Calbiochem (San

Diego, CA, USA). Oxaliplatin reference standard was generously provided by Sanofi-

Synthelabo (Malvern, PA, USA). Chloroplatinic acid, containing 1,000 mg/L Pt in 3.3%

hydrochloric acid (HCl), used for preparation of calibration solutions, was obtained from

Inorganic Ventures/IV Labs (Lakewood, NJ, USA). Iridium chloride, containing 1,000 mg/L

iridium (Ir) in 3.3% HCl, used for internal standardisation, was also purchased from

Inorganic Ventures/IV Labs. Nitric acid (HNO3) 70% Ultrex II ultrapure reagent was

obtained from Mallinckrodt Baker (Philipsburg, NJ, USA). Water used for the ICP-MS

analysis was sterile water for irrigation (Aqua B. Braun Medical, Melsungen, Germany). A

multi-element solution containing 10 mg/L of Ba, Be, Ce, Co, In, Mg, Pb, Th, and Tl (VAR-

TS-MS) was purchased from Inorganic Ventures/IV Labs. Drug-free human plasma was

obtained from the Central Laboratory for Blood Transfusion (Sanquin, Amsterdam, The

Netherlands).

Instrumentation

Analyses were performed on an ICP-quadrupole-MS (Varian 810-MS) equipped with a

90° reflecting ion mirror (Varian, Mulgrave, Victoria, Australia). The sample introduction

system consisted of a Micromist glass low-flow nebuliser (sample uptake 0.4 ml/min), a

peltier-cooled (4 °C) double pass glass spray chamber and a quartz torch. The spray

chamber was cooled to reduce the vapour loading on the plasma, increasing the

available energy for atomisation and ionisation of the elements of interest and to reduce

the formation of solvent based interferences. Sample transport from the SPS-3

autosampler (Varian) to the nebuliser was perfomed using a peristaltic pump. The

instrument was cooled by using a Kühlmobil 142 VD (Van der Heijden, Dörentrup,

Germany). Hoek Loos (Schiedam, The Netherlands) provided argon gas (4.6) with a

99.996% purity. Data were acquired and processed using the ICP-MS Expert Software

version 1.1 b49 (Varian). Further data handling was performed using Excel 2000

(Microsoft, Redmond, WA, USA). All measurements were carried out in a dedicated

Determination of platinum by ICP-MS

93

temperature-controlled, positively pressurised environment in order to maintain

optimum instrument performance and minimise contamination. All solutions were

prepared using plastic pipettes (Falcon, Becton Dickinson Labware, Franklin Lakes, NJ,

USA) and polypropylene tubes of 10 mL (Plastiques-Gosselin, Hazebrouck Cedex, France)

and 30 mL (Sarstedt AG&Co, Nümbrecht, Germany). Prior to method development, all

sample pretreatment devices were checked thoroughly for Pt, Ir, and hafnium

contamination and appeared to be suitable for Pt analyses.

Determination of Pt by ICP-MS

To optimise the ICP-MS signal for the high masses and to reduce the formation of oxides,

a solution containing 1,000 ng/L of Th, In, Ce, Ba, and Pt was used. Typically this 1,000

ng/L solution gave readings of 115In: 7x105 c/s; 232Th: 1x106 c/s, and 194Pt: 2x105 c/s. The

production of [CeO]+ was less than 1.0% of the total [Ce]+ counts. The formation of

doubly charged [Ba]2+ was less than 3%. The instrument settings are summarised in Table

1. Performance was checked daily. Other than a daily torch alignment, there was no

need to tune any of the other instrumental parameters. Conditions as depicted in Table

1 were kept constant and only replacement of consumables such as the torch, nebuliser,

and cones required additional tuning of the instrument settings. In this way, signals for

In, Th, and Pt deviated no more than 15% from the values mentioned above.

Table 1. ICP-MS instrument settings

Flow parameters (L/min) Ion optics (volts)

Plasma flow 18.0 First extraction lens -12

Auxiliary flow 1.80 Second extraction lens -240

Sheath gas 0.14 Third extraction lens -200

Nebuliser flow 1.00 Corner lens -230

Mirror lens left 29

Torch alignment (mm) Mirror lens right 25

Sampling depth 5.0 Mirror lens bottom 33

Entrance lens 6

Other Fringe bias -3.5

RF power (kW) 1.30 Entrance plate -60

Pump rate (mL/min) 0.4 Detector focus -500

Stabilisation delay 40 Pole bias 0.0

Chapter 2.2

94

For the detection of Pt, three isotopes 194Pt (abundance 33.0%), 195Pt (33.8%), and 196Pt

(25.2%) were monitored [18]. All three monitored Pt isotopes can be subject to the

interference of hafnium(Hf)-oxides [19]. But, because of low oxide formation and low

observed Hf counts in all of the analysed samples, these oxides were insignificant and no

corrections were necessary. The interference from Hg on 196Pt was corrected on-line by

monitoring 202Hg. In order to monitor unanticipated isobaric interferences, the 194Pt/195Pt

and 196Pt/195Pt ratios were measured for all samples. When ratios were similar to those

reported for natural Pt, it proved that the isotopic signals reflected the Pt content of the

sample with no other spectral interference.

The Pt isotope used for calculation of the validation parameters was 194Pt. The detection

mode for all isotopes was based on peak jumping, with peak dwell times of 50 ms, 25

scans per replicate, and three replicates per sample. The total measurement time for one

sample was 2.5 min. Ir was used as internal standard. Ir is expected to respond to matrix

effects and possible plasma fluctuations in the same way as Pt because of its similar

mass and ionisation potential. Internal standardisation was performed on each replicate

using 191Ir. Quantitation was based on the mean concentration of three replicates

analysed against a calibration curve using weighted linear regression analysis.

Preparation of reagents

A 1% v/v HNO3 solution, used for primary dilutions, was prepared by diluting ultrapure

concentrated HNO3 with B. Braun water. A solution containing 1% v/v drug-free pUF in

1% HNO3 was also prepared. Fresh solutions were prepared daily.

Assay development

Use of chloroplatinic acid as calibration standard

We tested the possibility of the use of chloroplatinic acid instead of cisplatin, carboplatin

and oxaliplatin as a standard for the preparation of calibration standards. For that

purpose, drug-free pUF samples were spiked at six concentration levels (7.50; 75.0; 225;

1.00x103; 7.00x103 and 1.00x105 ng/L Pt) using certified chloroplatinic acid, cisplatin,

carboplatin and oxaliplatin reference standards. These samples were diluted and

analysed in five-fold. Pt levels were quantified using calibration standards spiked with

chloroplatinic acid.

Determination of platinum by ICP-MS

95

Matrix effect and effectiveness of internal standardisation

To test the effect of pUF constituents on the detector response and to validate the use of

Ir as internal standard, the following matrices: 10%, 5%, 2%, 1.5%, 1%, 0.5% and 0% pUF

in 1% HNO3 were spiked with 50.0 ng/L Pt. Signals at a mass to charge (m/z) of 194 were

monitored and Pt concentrations were calculated against a calibration curve in 1% pUF

using the 191Ir signal as internal standard.

Preparation of stock solutions, calibration standards and, quality control samples

For this method, pUF samples were diluted with 1% HNO3 to reduce matrix effects and

contamination of the sample introduction system. As a compromise between a low

detection limit of the method and a minimal contamination of the sample introduction

system, we chose to dilute the samples 100-fold.

The chloroplatinic acid reference solution containing 1,000 mg/L Pt was diluted with 1%

HNO3 to obtain working solutions with concentrations ranging from 50.0 to 5.00x103

ng/L Pt. Working solutions were diluted with drug-free pUF:1% HNO3 (1:100) to obtain

calibration standards, ranging from 75.0 to 1.00x104 ng/L Pt in pUF (corresponding to

0.750 - 100 ng/L Pt in 1:100 diluted pUF). Calibration standards were analysed without

further dilution.

Stock solutions of cisplatin, carboplatin, and oxaliplatin in water, each containing a

concentration of drug equivalent to 400 mg/L Pt, were prepared. These stock solutions

were further diluted to obtain working solutions with concentrations ranging from 500

to 5.00x106 ng/L. Drug-free pUF was spiked with these working solutions to obtain

quality control (QC) samples at four concentration levels (75.0, 225, 1.00x103 and

7.00x103 ng/L Pt in pUF). Prior to analysis, these samples were diluted 100-fold with 1%

HNO3.

Furthermore, two additional concentration levels were prepared. The first was prepared

to validate the ability to quantify samples under the lower limit of quantitation (<LLOQ;

7.50 ng/L in pUF). This sample was diluted tenfold prior to analysis. The second was

prepared to validate the ability to quantify samples originally exceeding the upper limit

of quantification (>ULOQ; 1.00x105 ng/L). Prior to analysis, this sample was diluted 100-

fold with 1% HNO3 solution and then diluted 100-fold with pUF:1% HNO3 (1:100).

An internal standard solution of 1.00x104 ng/L Ir was prepared from an Ir chloride

reference solution containing 1,000 mg/L Ir. To 1.5 mL calibration standard or diluted QC

sample, 15 µL of internal standard solution was added (final internal standard

concentration 100 ng/L).

Chapter 2.2

96

Sample pretreatment

Whole blood samples were collected in 10 mL heparin-containing tubes (Becton

Dickinson Vacutainer Systems, Plymouth, UK). Plasma was obtained by centrifuging the

whole blood samples for 5 min (1,000 g, 4 °C). PUF was obtained by centrifuging the

plasma fraction through a 30 kDa cut-off ultrafiltrate filter (Centriplus YM-30, Millipore

Corporation, Bedford, MA, USA) for 15 min (1,000 g, 20 °C). The preparation of pUF from

plasma was performed immediately after blood collection, to prevent the decrease of

free Pt levels due to progressive ex vivo binding of Pt to plasma proteins and

erythrocytes [20]. All samples were stored at –20 °C until analysis. Prior to analysis, pUF

samples were thawed, vortex mixed, and subsequently diluted 100-fold with 1% HNO3

solution. However, when a pUF sample was expected to have a Pt concentration under

the LLOQ (75.0 ng/L), the sample was diluted tenfold with 1% HNO3 solution. When a

pUF sample contained a Pt concentration above the ULOQ (1.00x104 ng/L), successive

dilutions in pUF:1% HNO3 solution (1:100) were performed. To each 1.5 mL sample, 15

µL of internal standard solution was added. Subsequently, diluted samples were

transferred to autosampler tubes.

Validation procedures

Full validation according to the FDA guidelines [17] was, as far as applicable for ICP-MS,

performed for the assay.

Application of ICP-MS assay

The analytical method described in this paper is used to support clinical

pharmacokinetic studies of ultrafilterable Pt. An example of the analysis of pUF samples

of a patient treated with oxaliplatin 130 mg/m2 as a 2 h infusion is given here. Plasma

samples were taken up to 3 weeks after the administration of oxaliplatin, and pUF

samples were obtained and processed as described above.

Determination of platinum by ICP-MS

97

Results

Assay development

Use of chloroplatinic acid as calibration standard

The Student t-test performed for the comparison of the test samples prepared from the

four reference solutions yielded no significant difference between the four compounds

at any of the tested concentration levels.

Matrix effect and effectiveness of internal standardisation

As shown in Figure 1, significant signal suppression occurs with increasing pUF

concentration. Compared to 1% HNO3, a 10% pUF solution caused an ion suppression of

25%. However, as can also be seen in Figure 1, the internal standard corrected well for

this matrix effect.

Figure 1. Matrix effect and

effectiveness of internal stand-

ard correction. On the left y-axis,

the signal of 50 ng/L Pt in the

matrix is plotted relative to the

signal of 50 ng/L Pt in 0% pUF

(▲). On the right y-axis, the

concentration of Pt in ng/L is

plotted ( ), which is calculated

using the internal standard signal

Validation

Limit of quantification

The LLOQ of the assay was set at a Pt concentration of 75.0 ng/L in pUF when using a

standard 100-fold dilution. However, a tenfold dilution of pUF was used to enlarge the

sensitivity of the method to a LLOQ of 7.50 ng/L. Signal to noise ratios at the LLOQ level

ranged between 5 and 10 dependent on the drug-free pUF batch used, which was in

accordance with the requirement that the analyte response at the LLOQ should be at

least 5 times the response in a blank sample [17]. The acceptance criteria, that the LLOQ

60

70

80

90

100

110

0 5 10% pUF

Rela

tive

sign

al o

f 50

ng/L

pl

atin

um (

%)

45

47

49

51

53

55Pl

atin

um c

once

ntra

tion

(in n

g/L)

Chapter 2.2

98

was determined with a precision less than 20% and that the mean value deviated no

more than 20% from the actual value, were easily met [17].

Carry-over

To evaluate and minimise the effect of carry-over, we studied the signals of blank

samples following the ULOQ calibration sample (1.00x104 ng/L Pt diluted 100-fold), and

optimised the rinse time. A 35-second rinse time with 1% HNO3 between two samples

was required to avoid a memory effect from the preceding high concentration sample

and to achieve a blank signal <20% of the LLOQ standard signal. Shorter rinse times

resulted in carry-over of Pt from the preceding high concentration sample.

Linearity

Seven nonzero calibration standards in a dynamic range of 75.0-1.00x104 ng/L of Pt in

pUF were processed and analysed in singular in three separate analytical runs.

The calibration curve was best described by linear regression, using 1/(RSD% of triplicate

sample reading) as the weight-factor, to avoid bias in favour of samples with high

standard deviations. The calibration concentrations were back-calculated from the

responses. For deviations and relative standard deviations see Table 2. Deviations from

the nominal concentration were between –4.05 and 1.22% for all concentration levels,

which were all within the requirements of within ±20% for the LLOQ and ±15% for other

concentrations [17]. Relative standard deviations for the calibration samples were up to

5.32%. Correlation coefficients were between 0.999 and 1.00.

Table 2. Mean deviation from theoretical concentration (DEV %) and the relative standard deviation

(RSD %) (n=3) for chloroplatinic acid standards in pUF:1% HNO3 (1:100).

Concentration in pUF (ng/L)

Concentration in final matrix after 1:100 dilution (ng/L)

Mean concentration back calculated (ng/L)

DEV (%) from nominal concentration

RSD (%)

75.0 0.750 0.733 -2.31 1.30

150 1.50 1.44 -4.05 4.22

500 5.00 5.06 1.22 1.90

1.00x103 10.0 10.0 0.076 5.32

2.50x103 25.0 25.0 0.179 1.94

7.50x103 75.0 75.3 0.381 3.01

1.00x104 100 100 0.182 0.751

Determination of platinum by ICP-MS

99

Accuracy and precision

Accuracy, within-run, and between-run precisions of the method were determined by

assaying QC samples at six concentration levels, with different dilution factors. Five

replicates of each sample were analysed in three analytical runs. The accuracy was

expressed as a percentage of the nominal concentration and had to be within 80-120%

for the LLOQ and within 85-115% for the other concentrations. The within-run and

between-run precision were calculated by analysis of variances (ANOVA) for each test

concentration using the analytical run as the grouping variable. Precision should not

exceed ±20% for the LLOQ and ±15% for the other concentrations [17].

The within-run and between-run precision data are summarised in Table 3. The

accuracies for cisplatin, carboplatin, and oxaliplatin were between 85.4 and 119% for the

LLOQ concentration level and between 88.7 and 109% for the other concentration

levels. This showed that for all QC concentration levels, data were within generally

accepted limits for bioanalytical method validation. The different dilution factors of the

QC samples did not influence the performance of the method.

Table 3. Within-run and between-run precision data in pUF

Cisplatin Carboplatin Oxaliplatin Concentration in pUF (ng/L)

Concentration in final matrix after dilution (ng/L)

Within-run

Between-run

Within- run

Between-run

Within-run

Between-run

7.50 0.750 7.85 13.7 3.84 11.9 7.47 12.7

75.0 0.750 6.28 12.0 3.77 14.1 8.53 11.7

225 2.25 2.04 7.36 3.85 7.18 3.40 8.07

1.00x103 10.0 1.54 6.92 2.35 4.11 4.80 5.19

7.00x103 70.0 1.01 4.17 1.38 4.83 1.26 4.24

1.00x105 10.0 1.78 6.52 3.14 4.05 2.22 6.08

Specificity

From six individual batches of drug-free pUF, samples containing neither analyte nor

internal standard (blank) and samples containing 7.50 ng/L Pt (chloroplatinic acid) and

internal standard were prepared and diluted tenfold. These samples were prepared in

order to determine whether endogenous compounds interfered at the masses selected

for Pt or internal standard in the most concentrated matrix used (pUF:1% HNO3 1:10). All

samples were analysed in one analytical run. The signal of any interfering peak at m/z

194 in the blank solutions was not allowed to exceed 20% of the response of the LLOQ

standard. The response of any interfering peak at m/z 191 in the blank solution should

not exceed 5% of the response of 100 ng/L internal standard. Accuracies of the samples

Chapter 2.2

100

spiked with Pt at the LLOQ standard level had to be within 80-120% of the nominal value

[17].

Blank samples from six individual batches did not show interferences from endogenous

material at the m/z selected for Pt with a response >20% of the LLOQ standard signal.

The response of interfering peaks at m/z 191 did not exceed 5% of the response of 100

ng/L internal standard. Deviations from the nominal concentrations at the LLOQ level

were between –1.85 and 2.27%.

Internal standard interference test

Interference of the internal standard solution on the masses selected for Pt and

interference of Pt on the masses selected for the internal standard had to be assessed.

Drug-free pUF was spiked with Pt at ULOQ standard level and after a 100-fold dilution,

the Ir signal at m/z 191 was monitored. The response of the interfering peak at m/z 191

was less than the maximum allowed 5% of the response of 100 ng/L internal standard.

Drug-free pUF was diluted 1:100 and spiked with 100 ng/L Ir. The response of the

interfering peak at m/z 194 was less than the maximum allowed 20% of the response of

the LLOQ standard.

Stability

Stability was evaluated in Pt and Ir stock and working solutions under both processing

(24 h at ambient temperatures) and storage (2-8 °C) conditions. The analytes were

considered stable in the stock and working solutions when 95-105% of the original

concentration was recovered.

For evaluation of stability of Pt concentrations in pUF during storage and sample

processing, two QC solutions of each Pt agent (cisplatin, carboplatin, and oxaliplatin)

were sampled. Furthermore, two calibration samples in 1:100 diluted pUF containing

chloroplatinic acid were sampled to evaluate the stability of calibration standards.

Short-term stability of the analytes under processing conditions was evaluated by

comparing QC and calibration samples (stored for 24 h at ambient temperatures) with

freshly prepared samples. To determine the stability of diluted QC samples in the

autosampler, QC samples were analysed for one day and concentrations were compared

to concentrations measured at the start of the run.

Furthermore, the stability after three freeze-thaw cycles was investigated by comparing

QC and calibration samples that had been frozen (-20 °C) and thawed three times with

freshly prepared samples. Finally, long-term storage stability was assessed by

determining the Pt concentration in QC and calibration samples before and after three

months storage at -20 °C.

Determination of platinum by ICP-MS

101

The analytes were considered stable in the (un)diluted biological matrix when 85-115%

of the initial concentration was recovered.

The stability experiments showed that Pt levels in cisplatin, carboplatin, and oxaliplatin

stock and working solutions remained constant under processing conditions and during

storage for at least 6 months. However, chloroplatinic acid working solutions used for

preparation of calibration samples were not stable. Pt concentrations decreased 6.91%

after 24 h at ambient temperatures and 11.9% after three months of storage at 2-5 °C.

Pt levels in cisplatin, carboplatin, and oxaliplatin QC samples remained constant under

all tested conditions. No significant changes in concentration were observed after three

freeze-thaw cycles, nor after 24 h at ambient temperatures, after one day in the

autosampler, or after three months of storage. Calibration samples however, showed a

decrease in Pt concentration in all of the tested conditions. Although the decrease did

not exceed 15% in all of the tested conditions, the calibration samples were not

considered stable for more than one day. Furthermore, internal standard solutions were

stable for 24 h at ambient temperatures and for three months of storage.

Long-term stability has now been established for three months, but further testing is still

ongoing.

Application of the ICP-MS assay

The clinical applicability of the assay was demonstrated by analysis of human pUF

samples from a patient who received 130 mg/m2 oxaliplatin administered as a 2 h

infusion. Plasma samples were collected prior to administration and up to three weeks

after oxaliplatin infusion. A Pt concentration versus time profile for oxaliplatin in pUF is

presented in Figure 2. Up to three weeks after infusion, the Pt concentrations were still

higher than the LLOQ. For a typical mass spectrum of Pt and the internal standard Ir in a

100-fold diluted pUF sample, see Figure 3.

Chapter 2.2

102

1

10

100

1000

10000

0 100 200 300 400 500 600

Time (h)

Plat

inum

con

cent

ratio

n(µ

g/L)

Figure 2. A Pt concentration versus time profile in human pUF after intravenous infusion of oxaliplatin

administered as a 3 h infusion at a dose level of 130 mg/m2. Figure a shows the complete profile (up

to 3 weeks), while Figure b zooms in on the first 5 h after start of the infusion

Figure 3. Typical mass spectrum of Pt and Ir in a 1:100 diluted pUF sample

1

10

100

1000

10000

0 1 2 3 4 5Time (h)

a

b

Determination of platinum by ICP-MS

103

Discussion

We describe an ICP-MS assay for the quantification of Pt originating from cisplatin,

carboplatin, and oxaliplatin in human pUF. Sample pretreatment is very user friendly and

involves a simple dilution step of pUF with 1% HNO3.

Although the Varian ICP-MS has a linear performance over nine orders of magnitude, the

assay described here did not allow us to fully utilise this range. The concentration of the

highest calibration sample was limited by the memory effect it produced on a

successively analysed blank solution and by the required limit of quantification. As the

response of the blank solution following the ULOQ calibration standard should be less

than 20% of the LLOQ standard signal, the highest possible calibration standard and

rinse time had to be assessed. We found that rinsing 35 seconds with 1% HNO3 after the

analysis of the ULOQ calibration standard of 1.00x104 ng/L Pt (diluted 100-fold) allowed

the determination of the LLOQ standard of 75.0 ng/L (diluted 100-fold). Higher

concentrations caused the LLOQ to rise substantially, even with longer rinse times. This

was mainly caused by Pt adsorbing to the tubings and the spray chamber of the sample

introduction system. Replacing the tubings or cleaning the spray chamber immediately

resulted in a reduction of Pt signals. However, to reach Pt levels below 20% of LLOQ

level, it was necessary to both replace tubings and clean the spray chamber. Because the

memory effect is confined to the sample introduction system, a PFA spray chamber

could possibly reduce the memory effect. However, the PFA spray chamber cannot be

combined with a standard torch of the Varian 810-MS. It has to be combined with a

demountable torch with a Pt tipped injector. Obviously, this torch is not suitable for the

assay described here, because it raises the Pt background signal. Therefore, we solved

the memory problem by making demands on the ULOQ and by optimising the rinse

time. The concentration range of the assay was not confined by the memory effect

because we substantially enlarged this range by successfully validating up to 104-fold

dilution of high concentration samples to within the calibration range.

We also tested the matrix effect of pUF by spiking different concentrations of pUF with

Pt and analysing the samples against a calibration curve prepared in pUF:1% HNO3

(1:100). As was shown, a significant signal reduction with increasing pUF concentration

was observed, which could be due to dissolved salts in pUF samples [21]. Because matrix

effects were corrected for sufficiently by using Ir as internal standard, we could validate

the analysis of samples which were only diluted tenfold against a calibration curve

prepared in pUF:1% HNO3 (1:100). This was done to lower the LLOQ to 7.50 ng/L. There

was no additional value in validating even lower dilution factors because of relatively

high sample volumes needed for a single ICP-MS measurement (1.5 mL) and the limited

availability of pUF from one patient sample (typically 150 µL). Although the lowest

dilution factor validated was ten, our standard procedure involved a 100-fold dilution to

reduce contamination of the sample introduction system and cones.

Chapter 2.2

104

The limit of detection of the assay was not affected by the instrumental noise, but was

determined by the reagents and materials used. Effective control of pre-analytical factors

proved to be indispensable in order to ensure low background levels. Furthermore, it is

important to consider that Pt can also be present in control drug-free human plasma.

Several investigators have detected Pt in human body fluids like serum [22-24] and

blood [22-26] in humans who did not receive chemotherapy. Catalytic converters in cars

and dental alloys may cause elevated Pt levels in the human body. For this assay, we

thoroughly screened pUF for the presence of Pt before preparation of calibration

standards and QC samples.

In this assay, we used chloroplatinic acid for preparation of calibration standards. It was

shown that the different ligands around the central Pt atom (see Figure 4) did not affect

the signal generated by ICP-MS. Therefore, we could analyse samples with Pt originating

from the different anticancer agents in one analytical run. Moreover, chloroplatinic acid

is an easily available and well-certified reference compound. Because this compound is

available in solution, it is also relatively harmless compared to the solid Pt anticancer

agents.

Pt

Cl

Cl

NH3

NH3

O

O

O

O

Pt

NH3

NH3

Cisplatin Carboplatin

Pt

O

O

O

O

NH

2

NH2

Pt

Cl

Cl

Cl

Cl

Cl

Cl

H3O+

2-

2

Oxaliplatin Platinum chloride

Figure 4. Structural formula of Pt compounds

Results of the stability experiments showed us, however, that fresh solutions containing

chloroplatinic acid have to be prepared daily. The decrease of the amount of Pt analysed

in these solutions can probably be explained by the precipitation of the hydrolysed

chloroplatinic acid [27,28] or by adsorption of the chloroplatinic acid to the surface of

the plastic tubes.

Determination of platinum by ICP-MS

105

Performance of the assay was assessed by validating it according to most recent FDA

guidelines [17]. All validation results were within the requirements. The LLOQ of 7.50

ng/L in pUF was lower (1.3-7x) than reported in other papers which were published since

the year 2000 and describe detection of Pt originating from Pt anticancer agents in pUF

using ICP-quadrupole-MS [9-12]. Comparison of the method described here to our

former used GF-AAS methods [29,30] showed a 2600-fold gain in LLOQ using the ICP-

quadrupole-MS, which dramatically increased our time-window for the evaluation of

long-term pharmacokinetics of Pt agents. Using ICP-sector-field-MS, however, 2.5-30-

fold lower limits of detection have been reported in biological fluids, compared to our

method [25,31-35]. Yet, its additional value for the use in pharmacokinetic studies is

limited because the LLOQ was not determined by the sensitivity of the ICP-MS, but by

the background Pt levels of control drug-free human plasma and reagents.

Conclusion

A highly sensitive assay for the reliable and fast quantitative determination of Pt

originating from cisplatin, carboplatin, and oxaliplatin in human pUF using ICP-MS was

developed and subsequently validated according to current FDA guidelines. The

validated range of determination was 75.0 ng/L to 1.00x104 ng/L Pt in human pUF. The

LLOQ was lowered to 7.50 ng/L by including a tenfold dilution into the validation

procedures. Besides, by validating the possibility of diluting samples 104-fold, the

dynamic range was raised to 1.00x105 ng/L. The assay is now successfully applied in

long-term pharmacokinetic studies of patients being treated with cisplatin, carboplatin,

and oxaliplatin.

Chapter 2.2

106

References 1. Meinardi MT, Gietema JA, van der Graaf WT, van Veldhuisen DJ, Runne MA, Sluiter WJ, de Vries EG,

Willemse PB, Mulder NH, van den Berg MP, Koops HS, Sleijfer DT. Cardiovascular morbidity in long-term survivors of metastatic testicular cancer. J Clin Oncol 2000; 18: 1725-32.

2. Meinardi MT, Gietema JA, van Veldhuisen DJ, van der Graaf WT, de Vries EG, Sleijfer DT. Long-term chemotherapy-related cardiovascular morbidity. Cancer Treat Rev 2000; 26: 429-47.

3. Chaudhary UB, Haldas JR. Long-term complications of chemotherapy for germ cell tumours. Drugs 2003; 63: 1565-77.

4. Hartmann JT, Kollmannsberger C, Kanz L, Bokemeyer C. Platinum organ toxicity and possible prevention in patients with testicular cancer. Int J Cancer 1999; 83: 866-9.

5. Gray AL. Mass-spectrometric analysis of solutions using an atmosperic pressure ion source. The Analyst 1975; 100: 289-99.

6. Tothill P, Matheson LM, Smyth JF. Inductively coupled plasma mass spectrometry for the determination of platinum in animal tissue and a comparison with atomic absorption spectrometry. J Anal Atom Spectrom 1990; 5: 619-22.

7. Allain P, Berre S, Mauras Y, Le Bouil A. Evaluation of inductively coupled mass spectrometry for the determination of platinum in plasma. Biol Mass Spectrom 1992; 21: 141-3.

8. Gamelin E, Allain P, Maillart P, Turcant A, Delva R, Lortholary A, Larra F. Long-term pharmacokinetic behavior of platinum after cisplatin administration. Cancer Chemother Pharmacol 1995; 37: 97-102.

9. Sessa C, Capri G, Gianni L, Peccatori F, Grasselli G, Bauer J, Zucchetti M, Vigano L, Gatti A, Minoia C, Liati P, Van den BS, Bernareggi A, Camboni G, Marsoni S. Clinical and pharmacological phase I study with accelerated titration design of a daily times five schedule of BBR3464, a novel cationic triplatinum complex. Ann Oncol 2000; 11: 977-83.

10. Morrison JG, White P, McDougall S, Firth JW, Woolfrey SG, Graham MA, Greenslade D. Validation of a highly sensitive ICP-MS method for the determination of platinum in biofluids: application to clinical pharmacokinetic studies with oxaliplatin. J Pharm Biomed Anal 2000; 24: 1-10.

11. Liu J, Kraut E, Bender J, Brooks R, Balcerzak S, Grever M, Stanley H, D'Ambrosio S, Gibson-D'Ambrosio R, Chan KK. Pharmacokinetics of oxaliplatin (NSC 266046) alone and in combination with paclitaxel in cancer patients. Cancer Chemother Pharmacol 2002; 49: 367-74.

12. Bettinelli M, Spezia S, Ronchi A, Minoia C. Determination of Pt in biological fluids with ICP-MS: Evaluation of analytical uncertainty. Atom Spectrosc 2004; 25: 103-11.

13. Casetta B, Roncadin M, Montanari G, Fulanut M. Determination of platinum in biological fluids by ICP-mass spectrometry. Atomic Spectroscopy 1991; 12: 81-6.

14. Nygren O, Vaughan GT, Florence TM, Morrison GM, Warner IM, Dale LS. Determination of platinum in blood by adsorptive voltammetry. Anal Chem 1990; 62: 1637-40.

15. Calvert H, Judson I, van der Vijgh WJ. Platinum complexes in cancer medicine: pharmacokinetics and pharmacodynamics in relation to toxicity and therapeutic activity. Cancer Surv 1993; 17: 189-217.

16. Graham MA, Lockwood GF, Greenslade D, Brienza S, Bayssas M, Gamelin E. Clinical pharmacokinetics of oxaliplatin: a critical review. Clin Cancer Res 2000; 6: 1205-18.

17. U.S.Food and Drug Administration, Center for Drug Evaluation and Research, Guidance for Industry, Bioanalytical Method Validation. FDA. U.S. Food and Drug Administration: Center for Drug Evaluation and Research: Guidance for Industry: Bioanalytical Method Validation. http://www fda gov/cder/guidance/ 4252fnl pdf 2001.

18. Rosman KJR, Taylor PDP. Isotopic compositions of the elements 1997; International Union of Pure and Applied Chemistry. http://www physics curtin edu au/iupac/docs/Final97 pdf 1997.

19. Lustig L, Zang S, Michalke B, Schramel P, Beck W. Platinum determination in nutrient plants by inductively coupled plasma mass spectrometry with special respect to the hafnium oxide interference. Fresenius J Anal Chem 1997; 357: 1157-63.

20. Johnsson A, Bjork H, Schutz A, Skarby T. Sample handling for determination of free platinum in blood after cisplatin exposure. Cancer Chemother Pharmacol 1998; 41: 248-51.

21. McCurdy E, Potter D. Optimising ICP-MS for the determination of trace metals in high matrix samples. Spectroscopy Europe 2001; 13: 10-3.

22. Barany E, Bergdahl IA, Bratteby LE, Lundh T, Samuelson G, Schutz A, Skerfving S, Oskarsson A. Relationships between trace element concentrations in human blood and serum. Toxicol Lett 2002; 134: 177-84.

Determination of platinum by ICP-MS

107

23. Barany E, Bergdahl IA, Bratteby LE, Lundh T, Samuelson G, Schutz A, Skerfving S, Oskarsson A. Trace element levels in whole blood and serum from Swedish adolescents. Sci Total Environ 2002; 286: 129-41.

24. Goulle JP, Mahieu L, Castermant J, Neveu N, Bonneau L, Laine G, Bouige D, Lacroix C. Metal and metalloid multi-elementary ICP-MS validation in whole blood, plasma, urine and hair Reference values. Forensic Sci Int 2005; 153: 39-44.

25. Begerow J, Turfeld M, Dunemann L. Determination of physiological palladium, platinum, iridium and gold levels in human blood using double focusing magnetic sector field inductively coupled plasma mass spectrometry. Journal of Analytical Atomic Spectrometry 1997; 12: 1095-8.

26. Messerschmidt J, Alt F, Tolg G, Angerer J, Schaller KH. Adsorptive voltammetric procedure for the determination of platinum baseline levels in human body fluids. Fresenius J Anal Chem 1992; 343: 391-4.

27. Spieker WA, Liu J, Miller JT, Kropf AJ, Regalbuto JR. An EXAFS study of the co-ordination chemistry of hydrogen hexachloroplatinate(IV) 1. Speciation in aqueous solution. Applied Catalysis A: General 2002; 232: 219-35.

28. Nischwitz V, Michalke B, Kettrup A. Speciation of Pt(II) and Pt(IV) in spiked extracts from road dust using on-line liquid chromatography-inductively coupled plasma mass spectrometry. J Chromatogr A 2003; 1016: 223-34.

29. van Warmerdam LJC, van Tellingen O, Maes RAA, Beijnen JH. Validated method for the determination of carboplatin in biological fluids by Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 1995; 351: 777-81.

30. Brouwers EEM, Tibben MM, Joerger M, van Tellingen O, Rosing H, Schellens JHM, Beijnen JH. Determination of oxaliplatin in human plasma and plasma ultrafiltrate by graphite-furnace atomic-absorption spectrometry. Anal Bioanal Chem 2005; 382: 1484-90.

31. Hann S, Koellensperger G, Kanitsar K, Stingeder G, Brunner M, Erovic B, Muller M, Reiter C. Platinum determination by inductively coupled plasma-sector field mass spectrometry (ICP-SFMS) in different matrices relevant to human biomonitoring. Anal Bioanal Chem 2003; 376: 198-204.

32. Krachler M, Alimonti A, Petrucci F, Irgolic KJ, Forastiere F, Caroli S. Analytical problems in the determination of platinum-group metals in urine by quadrupole and magnetic sector field inductively coupled plasma mass spectrometry. Analytica Chimica Acta 1998; 363: 1-10.

33. Rodushkin I, Engstrom E, Stenberg A, Baxter DC. Determination of low-abundance elements at ultra-trace levels in urine and serum by inductively coupled plasma-sector field mass spectrometry. Anal Bioanal Chem 2004; 380: 247-57.

34. Begerow J, Turfeld M, Dunemann L. Determination of physiological palladium and platinum levels in urine using double focusing magnetic sector field ICP-MS. Fresenius J Anal Chem 1997; 359: 427-9.

35. Spezia S, Bocca B, Forte G, Gatti A, Mincione G, Ronchi A, Bavazzano P, Alimonti A, Minoia C. Comparison of inductively coupled plasma mass spectrometry techniques in the determination of platinum in urine: quadrupole vs. sector field. Rapid Commun Mass Spectrom 2005; 19: 1551-6.

Chapter 2.3

Determination of ruthenium originating from the investigational anticancer drug NAMI-A in human plasma ultrafiltrate,

plasma, and urine by inductively coupled plasma mass spectrometry

Elke E.M. Brouwers Matthijs M. Tibben

Hilde Rosing Jan H.M. Schellens

Jos H. Beijnen

Rapid Communications in Mass Spectrometry 2007; 21; 1521-1530

Chapter 2.3

110

Abstract

We present a highly sensitive, rapid method for the determination of ruthenium (Ru)

originating from the investigational anticancer drug NAMI-A in human plasma

ultrafiltrate, plasma, and urine. The method is based on the quantification of Ru by

inductively coupled plasma mass spectrometry and allows quantification of 30 ng/L Ru

in plasma ultrafiltrate and urine, and 75 ng/L Ru in human plasma in 150 µL of matrix.

The sample pretreatment procedure is straight forward and only involves dilution with

appropriate diluents. The performance of the method, in terms of accuracy and

precision, fulfilled the most recent FDA guidelines for bioanalytical method validation.

Validated ranges of quantification were 30.0 to 1x105 ng/L for Ru in plasma ultrafiltrate

and urine and 75.0 to 1x105 ng/L for Ru in plasma. The applicability of the method and its

superiority to atomic absorption spectrometry were demonstrated in two patients who

were treated with intravenous NAMI-A in a phase I trial. The assay is now successfully

used to support pharmacokinetic studies in cancer patients treated with NAMI-A.

Determination of ruthenium by ICP-MS

111

Introduction

Platinum (Pt) coordination compounds such as cisplatin, carboplatin, and oxaliplatin

play a major role in the treatment of cancer. Their clinical utility is, however, hampered

by severe side effects such as nephro-, oto-, and neurotoxicity [1]. Besides, intrinsic and

acquired resistance of several tumour types limit their optimal therapeutic use [2,3].

These limitations have encouraged the search for other cytotoxic coordination

compounds with better safety profiles and enhanced antitumour characteristics. In

addition to a wide range of Pt-containing compounds, also ruthenium (Ru) compounds

have also been synthesised and tested for their therapeutic potentials. Ru complexes are

regarded as promising alternatives for Pt complexes. NAMI-A [Imidazolium-trans

(imidazole)(dimethylsulfoxide)tetrachlororuthenate(III)] (Figure 1) [4], and KP1019 or

FFC14A (Indazolium-trans-[tetrachlorobis(1H-indazole)ruthenate(III)] [5,6] are the first Ru

complexes that have finished phase I studies. In preclinical studies, NAMI-A appeared to

be mainly effective against lung metastases [7-9], whereas KP1019 showed activity

against colon carcinomas and their metastases [10].

Figure 1. Molecular formula of NAMI-A (Mw 458.18 g/mol)

The ability to measure Ru exposure in biological matrices is a prerequisite in

understanding the pharmacokinetics of Ru anticancer agents. Graphite-furnace atomic-

absorption-spectrometry (GF-AAS), which is generally used to evaluate

pharmacokinetics of metal-based anticancer agents [11-18], lacked the required

sensitivity to quantify Ru in samples of all dose levels in a NAMI-A phase I trial which was

performed at our institute [4]. Hence, a more advanced and sensitive technique such as

inductively coupled plasma mass spectrometry (ICP-MS) is needed. Due to its extremely

high sensitivity and specificity ICP-MS greatly increases the pharmacokinetic

concentration window which can be studied and the technique is becoming the method

of choice in anticancer metallodrug research. Since the development of the first ICP-MS

assay in 1975 [19], considerable changes have been made in the design and

performance of the ICP-MS and the instrument has proven its applicability in oncology

by the determination of Pt-containing anticancer agents in biological fluids of patients

[20-27]. For Ru, ICP-MS was already used in combination with size exclusion

N

NH

-

NH+

NH

RuClCl

Cl Cl

SO(CH3)2

Chapter 2.3

112

chromatography (SEC) to study the interaction of Ru compounds with serum proteins

[28-30]. However, until now, no ICP-MS method has been described for the

determination of Ru originating from anticancer agents in biological fluids.

In order to assess Ru levels originating from NAMI-A in ultrafiltrated plasma (pUF),

plasma, and urine, we developed a highly sensitive, rapid ICP-MS assay. For this method,

samples were diluted with appropriate diluents to prevent contamination of the sample

introduction system. Although the presence of polyatomic interferences from matrix

components confined the quantification limits, the method was still 740- (pUF), 1500-

(plasma), and 3700- (urine) fold more sensitive than the GF-AAS method published by

Crul et al. [11].

In this article we describe the development and validation of the method, according to

most recent FDA guidelines on bioanalytical method validation [31]. Validated ranges of

quantification were 30.0 to 1x104 ng/L for Ru in plasma ultrafiltrate and urine and 75.0 to

1x104 ng/L for Ru in plasma. The applicability of the method and its superiority to GF-

AAS were demonstrated in two patients who had been treated intravenously with NAMI-

A in a phase I trial.

Experimental

Chemicals

NAMI-A reference standard, used for preparation of calibration solution and quality

control (QC) samples, was obtained from the Department of Chemical Sciences,

University of Trieste, Italy. Yttrium ICP standard, containing 1,000 mg/L yttrium (Y), used

for internal standardisation, was purchased from Merck (Darmstadt, Germany). Nitric

acid (HNO3) 70% Ultrex II ultrapure reagent was obtained from Mallinckrodt Baker

(Philipsburg, NJ, USA). Edta diammonium salt and Triton X-100 (4-oxtylphenol

polyethoxylate) were from Sigma-Aldrich (St. Louis, MO, USA). Water used for the ICP-MS

analysis was sterile water for irrigation (Aqua B. Braun Medical, Melsungen, Germany). A

multi-element solution containing 10 mg/L of Ba, Be, Ce, Co, In, Mg, Pb, Th, and Tl (VAR-

TS-MS) was purchased from Inorganic Ventures/IV Labs (Lakewood, NJ, USA). Drug-free

human heparinised plasma was obtained from the Central Laboratory for Blood

Transfusion (Sanquin, Amsterdam, The Netherlands). Drug-free urine from healthy

volunteers was used. Van Linde Gas Benelux (Schiedam, the Netherlands) provided

argon gas (4.6) of 99.996% purity.

Determination of ruthenium by ICP-MS

113

Instrumentation

Analyses were performed on an ICP-quadrupole-mass spectrometer (Varian 810-MS)

equipped with a 90° reflecting ion mirror (Varian, Mulgrave, Victoria, Australia). The

sample introduction system consisted of a Micromist glass low flow nebuliser (sample

uptake 0.28 mL/min), a peltier-cooled (4 °C) double pass glass spray chamber, a quartz

torch, and a nickel sampler and skimmer cone (Varian). The spray chamber was cooled to

reduce the vapour loading on the plasma, increasing the available energy for

atomisation and ionisation of the elements of interest and to reduce the formation of

solvent based interferences. Sample transport from the SPS-3 autosampler (Varian) to

the nebuliser was performed using a peristaltic pump. The instrument was cooled by

using a Kühlmobil 142 VD (Van der Heijden, Dörentrup, Germany). Data were acquired

and processed using the ICP-MS Expert Software version 1.1 b49 (Varian). Further data

handling was performed using Excel 2003 (Microsoft, Redmond, WA, USA) and SPSS 11

(SPSS, Inc. Chicago, IL, USA). All measurements were carried out in a dedicated

temperature-controlled, positively pressurised environment in order to maintain

optimum instrument performance and minimise contamination. All solutions were

prepared using plastic pipettes (Falcon, Becton Dickinson Labware, Franklin Lakes, NJ,

USA) and 10 mL (Plastiques-Gosselin, Hazebrouck Cedex, France) and 30 mL (Sarstedt

AG&Co, Nümbrecht, Germany) polypropylene tubes. Prior to method development, all

sample pretreatment devices were checked thoroughly for Ru contamination and

appeared to be suitable for Ru analyses.

Determination of Ru by ICP-MS

To optimise the ICP-MS signal for the mid range masses and to reduce the formation of

oxides and doubly charged ions, a solution containing 1,000 ng/L of Th, In, Ce, Ba, and

Ru was used. Typically this 1,000 ng/L solution gave readings of 115In: 1.4 x 106 counts per

second (c/s); 232Th: 7.0 x 105 c/s, 101Ru: 1.6 x 105 c/s, and 102Ru: 2.9 x 105 c/s. The production

of [CeO]+ was less than 2.0% of the total [Ce]+ counts. The formation of doubly charged

[Ba]2+ was less than 0.2%. Instrument settings are summarised in Table 1. The

performance was checked daily. Other than a daily torch alignment, there was no need

to tune any of the other instrumental parameters. The conditions as depicted in Table 1

were kept constant and only replacement of consumable parts such as torch, nebuliser,

and cones required additional tuning of the instrument settings. Thus, the signals never

deviated more than 15% of the values for In, Th, Ru, doubly charged, and oxides as

mentioned above.

The Ru isotope used for calculation of the validation parameters was 101Ru, because this

isotope gave the most stable results. The detection mode for all isotopes was based on

peak jumping with peak dwell times of 50 ms, 25 scans per replicate, and three

Chapter 2.3

114

replicates per sample. The total measurement time for one sample during validation

procedures was three minutes. However, this analysis time included an additional

minute to monitor signals of isotopes which could possibly interfere with the Ru signal

by formation of polyatomic interferences (a.o. Zn, Cu, Ni, Sr, Rb). This was done to

elucidate the origin of potential interferences. During routine measurements, analysis

times could be reduced to 2 minutes.

Y was used as internal standard. It is expected that, because of its similar mass and

ionisation potential, the behaviour of Y will accurately reflect that of Ru in a way that it

will respond similar to matrix effects and possible plasma fluctuations. Internal

standardisation was performed on each replicate using 89Y. Quantification was based on

the mean concentration of three replicates analysed against a calibration curve using

weighted linear regression analysis.

Table 1. ICP-MS instrument settings

Flow parameters (L/min) Ion optics (volts)

Plasma flow 15.0 First extraction lens -12

Auxiliary flow 1.65 Second extraction lens -220

Sheath gas 0.21 Third extraction lens -230

Nebuliser flow 1.02 Corner lens -240

Mirror lens left 29

Torch alignment (mm) Mirror lens right 23

Sampling depth 5.0 Mirror lens bottom 23

Entrance lens 4

Other Fringe bias -3.0

RF power (kW) 1.40 Entrance plate -30

Pump rate (mL/min) 0.28 Detector focus -500

Stabilisation delay 30 Pole bias 0.0

Preparation of reagents

A 1% (v/v) HNO3 solution in water, used for primary pUF and urine dilutions, was

prepared. A 0.01% (g/v) edta diammonium salt and Triton X-100 mixture in water (0.01%

EDTA-Triton) was prepared for primary dilutions of plasma samples. Solutions containing

1% (v/v) drug-free pUF in 1% HNO3, 1% (v/v) drug-free urine in 1% HNO3, and 1% (v/v)

drug-free plasma in 0.01% EDTA-Triton were prepared for preparation of calibration

standards and for dilution of samples exceeding the upper limit of quantification.

Reagents were prepared freshly before use.

Determination of ruthenium by ICP-MS

115

Preparation of stock solutions, calibration standards, and quality control samples

For this method, samples were diluted to reduce contamination of the sample

introduction system. As a compromise between the low detection limit of the method

and a minimal contamination of the sample introduction system, we chose to dilute the

samples 100-fold.

A NAMI-A stock solution containing 400 mg/L Ru in water was prepared to obtain

working solutions with concentrations ranging from 500 to 5.00x103 ng/L Ru. For pUF

and urine, working solutions were diluted with drug-free pUF:1% HNO3 (1:100, v/v) and

drug-free urine:1% HNO3 (1:100, v/v) to obtain calibration standards ranging from 300 to

1.00x104 ng/L Ru in pUF and urine (corresponding to 3.00 to 100 ng/L Ru in 1:100 v/v

diluted matrix). For plasma, 1% HNO3 was not chosen as a diluent, because it can

precipitate proteins, and thereby block the nebuliser. Therefore, working solutions were

diluted with drug-free plasma:0.01% EDTA-Triton (1:100, v/v) to obtain calibration

standards ranging from 750 to 1.00x104 ng/L Ru in plasma (corresponding to 7.50 to 100

ng/L Ru in 1:100 v/v diluted matrix). Calibration standards were analysed without further

dilution.

An NAMI-A stock solution, prepared from a separate weighing, was diluted with water in

order to obtain working solutions with Ru concentrations ranging from 2.50x103 to

5.00x106 ng/L. Drug-free pUF and urine were spiked with these working solutions to

obtain quality control (QC) samples at four concentration levels (300, 750, 2.50x103, and

7.50x103 ng/L Ru). Prior to analysis, these samples were diluted 100-fold with 1% HNO3.

Drug-free plasma was also spiked at four concentration levels (750, 1.5 x103, 2.50x103,

and 7.50x103 ng/L Ru). These samples were diluted 100-fold with 0.01% EDTA-Triton

prior to analysis.

For all matrices, two additional concentration levels were prepared. The first (30.0 ng/L

in pUF and urine and 75.0 ng/L in plasma) was prepared to validate the ability to

quantify samples below the lower limit of quantification (<LLOQ). These samples were

diluted tenfold prior to analysis. The second (1.00x105 ng/L for all matrices) was prepared

to validate the ability to quantify samples originally exceeding the upper limit of

quantification (>ULOQ). Prior to analysis, for pUF and urine, these samples were diluted

100-fold with 1% HNO3 and then 100-fold with pUF:1% HNO3 (1:100, v/v) or urine:1%

HNO3 (1:100, v/v). For plasma these samples were diluted 100-fold with 0.01% EDTA-

Triton and successively 100-fold with plasma:0.01% EDTA-Triton (1:100, v/v).

An internal standard solution of 2.00x104 ng/L Y was prepared from an Y reference

solution of 1,000 mg/L. To 1.5 mL calibration standard or diluted QC sample, 15 µL of

internal standard solution was added (final internal standard concentration 200 ng/L).

Chapter 2.3

116

Sample pretreatment

Whole blood samples were collected in 10 mL heparin-containing tubes (Becton

Dickinson Vacutainer Systems, Plymouth, UK). Plasma was obtained by centrifuging the

whole blood samples for 5 min (1,000 g, 4 °C). PUF was obtained by centrifuging the

plasma fraction through a 30 kDa cut-off ultrafiltrate filter (Centriplus YM-30, Millipore

Corporation, Bedford, MA, USA) for 10 min (1,000 g, 20 °C). The preparation of pUF from

plasma was performed immediately after blood collection, to prevent the decrease of

free Ru levels due to progressive ex vivo binding of Ru to plasma proteins. Urine samples

were collected in 30 mL polypropylene tubes (Sarstedt AG&Co, Nümbrecht, Germany).

All samples were stored at –20 °C until analysis. Prior to analysis, pUF and urine samples

were thawed, vortex mixed and subsequently diluted 100-fold with 1% HNO3 solution.

However, when a pUF or urine sample was expected to have a Ru concentration under

the LLOQ (300 ng/L), the sample was diluted tenfold with a 1% HNO3 solution. When a

pUF or urine sample contained a Ru concentration above the ULOQ (1.00x104 ng/L),

successive dilutions in pUF:1% HNO3 (1:100, v/v) or urine:1% HNO3 (1:100, v/v) were

performed. A similar procedure was followed for plasma samples. However, these were

diluted with a 0.01% EDTA-Triton solution and if necessary successively with a

plasma:0.01% EDTA-Triton solution (1:100, v/v). To each 1.5 mL diluted sample, 15 µL of

internal standard solution was added. Subsequently, diluted samples were transferred to

autosampler tubes.

Validation procedures

For the pUF assay, full validation according to the FDA guidelines [31] was, as far as

applicable for ICP-MS, performed. For urine and plasma, a partial validation was carried

out. According to the FDA guidelines a partial validation is sufficient to test the method,

when a change in matrix with the same analyte is concerned. A full validation required

the assessment of linearity, accuracy, and precision in three analytical runs, whereas a

partial validation only required the assessment of these parameters in one analytical run.

Application of ICP-MS assay

The analytical method described in this paper is used to support clinical pharmaco-

kinetic studies of NAMI-A. Phase I study samples from a patient who was treated with 2.4

mg/m2 NAMI-A, which could not be analysed by GF-AAS [4] due to its lack of sensitivity,

were analysed using the method described here. Additionally, samples of a patient who

was treated with 78 mg/m2 NAMI-A, which could be partly analysed by GF-AAS, were re-

analysed to demonstrate the reliability of both techniques. Plasma samples were taken

Determination of ruthenium by ICP-MS

117

up to 24 h after the intravenous administration of NAMI-A and pUF samples were

obtained and processed as described above.

Non-compartmental pharmacokinetics parameters (terminal half-life (t1/2) and area under

the concentration-time curve (AUC) from time point 0 to 24 h) were estimated by the

computer program WinNonlinTM (version 5.0, Pharsight Corporation, Mountain View,

California, USA). The maximal drug concentration (Cmax) was derived directly from the

experimental data.

Results and discussion

Validation

Interferences

The determination of Ru by ICP-MS might be hampered by the presence of spectral

interferences originating from the biological matrices. Unfortunately, ICP-quadrupole-

MS lacks the resolution power to separate interfering mass to charge (m/z) ratios from

analyte m/z signals, because the quadrupole mass analyser limits the resolution to

approximately one unit mass. ICP-sector-field-MS could reduce this problem; however,

Rodushkin et al. showed that, even when using this technique, not all interferences

could be separated [32]. Despite the presence of interferences, the limit of detection of

ICP-quadrupole-MS was judged sufficient for the investigation of Ru pharmacokinetics

after a NAMI-A infusion.

To prevent spectral interferences, a careful selection of the analyte isotope was of prime

importance. In order to select the most suitable isotope and to evaluate unresolved

spectral interferences, scans of 1% HNO3, 0.01% EDTA-Triton, and drug-free pUF, urine,

and plasma were made covering the Ru masses of interest (99Ru (natural abundance

12.7%), 101Ru (17.0%), 102Ru (31.6%), and 104Ru (18.7%) [33]). Of each human matrix, six

batches were selected, which were all diluted tenfold prior to analysis. The signal

intensities of the ions are shown in Table 2. To get an impression of the height of the

signals, Table 2 also shows the signal intensities of each isotope of 1 ng/L Ru in 1% HNO3.

As expected, the diluents 1% HNO3 and 0.01% EDTA-Triton did not show interfering

peaks at the m/z ratios of Ru, indicating that the noise of the instrument was negligible.

Data from Table 2 show that, compared to 1% HNO3 and 0.01% EDTA-Triton, increased

signals were observed for all isotopes in the biological matrices. Because the ratios of the

intensities of the Ru isotopes did not approach the natural ratios, these raised signals

reflect, at least partly, substantial interferences. [64Zn35Cl]+, [59Co40Ar]+, and [63Cu36Ar]+ ions

may interfere with 99Ru analysis in plasma and urine. Although 101Ru suffers from only

minor interferences, increased signals for plasma samples, probably due to [65Cu36Ar]+,

Chapter 2.3

118

were observed. 102Ru showed interfering signals in all matrices, but signals were highest

for plasma. Interfering peaks at m/z 102 could be caused by the formation of [62Ni40Ar]+,

[66Zn36Ar]+, [67Zn35Cl]+, and [65Cu37Cl]+ ions. The mass spectra for 104Ru in pUF, plasma, and

urine were also dominated by interferences, of which urine samples showed the highest

signals. These interferences are possibly caused by amongst others [64Ni40Ar]+, [64Zn40Ar]+,

and [67Zn37Cl]+ ions. In addition to polyatomic argides and chlorides, also strontium- and

rubidium-containing metal oxide ions could interfere with Ru analysis. Due to relatively

high Sr (up to 415 µg/L) and Rb (up to 2700 µg/L) concentrations in plasma and urine

[34], oxides could well affect sub-µg/L Ru analysis, even though oxide formation was

optimised to be less than 2%. Doubly charged interferences were not expected to make

a substantial contribution to any of the Ru isotope signals, because these interferences

were minimised to less than 0.2% by the optimisation of the method.

Table 2. Intensities (c/s) at m/z of Ru isotopes in diluents and six batches drug-free human pUF, urine,

and plasma

Isotope 99Ru 101Ru 102Ru 104Ru

Intensity of 1 ng/L Ru 118 165 295 156

1% HNO3 10 3 23 6

0.01% EDTA-Triton 13 9 8 22

pUF 1 32 37 35 164

pUF 2 26 18 33 187

pUF 3 28 14 92 203

pUF 4 25 17 42 267

pUF 5 21 17 51 212

pUF 6 25 20 44 207

Plasma 1 432 188 147 281

Plasma 2 351 128 139 154

Plasma 3 261 138 128 138

Plasma 4 300 98 165 211

Plasma 5 331 119 105 32

Plasma 6 309 79 124 222

Urine 1 173 42 90 877

Urine 2 89 20 76 646

Urine 3 73 27 65 454

Urine 4 57 55 161 315

Urine 5 46 22 96 280

Urine 6 187 38 115 1179

Determination of ruthenium by ICP-MS

119

In conclusion, pUF samples showed, as expected, the least interfering peaks, because

this is the cleanest matrix. Because the signals at the m/z values for zinc and copper were

much lower in pUF samples than in human plasma (data not shown), it is thought that

protein binding prevents high levels of these elements to be present in the ultrafiltrate.

Compared with urine, plasma samples generally showed higher interfering signals. This

could be because copper is generally present at 100-fold higher concentrations in

plasma than in urine [32,34,35]. Thus the contribution from copper interferences was

expected to be lower in the latter matrix. The advantage of 101Ru and 102Ru over 99Ru and 104Ru is that the 101Ru and 102Ru isotopes suffer less from interferences. We decided to use 101Ru in future work because this isotope gave the most stable results and because

interfering signals were relatively low. Therefore, the LLOQ could be minimised without

losing precision. In addition to the detection of 101Ru, for all samples, the 101Ru/102Ru ratio

was measured in order to monitor for unanticipated spectral interferences. When ratios

were similar to those reported for natural Ru (0.538), it proved that the isotopic signals

reflected the Ru content of the sample without significant spectral interferences. Isobaric

interference of 102Pd isotope on 102Ru were corrected automatically online.

Non-spectral interferences, most probably caused by large amounts of organic

compounds and inorganic salts [36], were corrected by external calibration using matrix

matched calibration samples and internal standardisation. The effectiveness of internal

standardisation was determined by spiking solutions containing increasing

concentrations of matrices (0, 1, 5, and 10%) with 50 ng/L Ru. Signals at m/z 101 were

monitored and Ru concentrations were calculated using 89Y as internal standard. A

significant signal suppression occurred with increasing matrix concentrations (Figure 2).

However, Ru concentrations were within 85-115% of the actual concentration in all

samples, showing that the internal standard corrected for the non-spectral matrix

effects.

Because matrix effects were corrected sufficiently by using Y as internal standard, we

could validate the analysis of samples which were only diluted tenfold against a

calibration curve which was diluted 100-fold. This was done to lower the LLOQ.

Although the lowest dilution factor validated was ten, our standard procedure involved

a 100-fold dilution to reduce contamination of the sample introduction system and

cones.

Chapter 2.3

120

50

75

100

125

0 5 10

Matrix (%)

Acc

urac

y (%

)

Accuracy in pUF Accuracy in plasmaAccuracy in urine

Figure 2. Matrix effect of pUF, plasma, and urine. The left graph depicts the signal of 50 ng/L Ru in the matrix relative to the signal of 50 ng/L Ru in 0% matrix, whereas the right graph shows the accuracy after calculation of the concentrations using correction with 89Y.

Limit of quantification

For pUF and urine the LLOQ of the assay was set at a Ru concentration of 300 ng/L when

using a standard 100-fold dilution. However, a tenfold dilution could be used to lower

the LLOQ of the method to 30.0 ng/L. For plasma, the LLOQ of the assay was set at a Ru

concentration of 750 ng/L when using a standard 100-fold dilution. Using a tenfold

dilution, the LLOQ was lowered to 75 ng/L. Signal to noise (S/N) ratios for 101Ru at the

LLOQ level when using a tenfold dilution ranged between 10 and 30, 5 and 12, and 7

and 18, for pUF, plasma, and urine respectively. This was in accordance with the

requirement that the analyte response at the LLOQ should be at least 5 times the

response in a blank sample [31]. S/N ratios were dependent on the batches of the drug-

free matrices used. The acceptance criteria, which required that the LLOQ was

determined with a precision less than 20% and that the mean value deviated no more

than 20% from the actual value, were easily met (Table 3) [31]. Limits of detection

(LODs), defined as S/N ratios of three in the final dilution (the analyte response was at

least 3 times the response of a blank sample), were 5.00 ng/L, 29.0 ng/L, and 9.00 ng/L in

pUF, plasma, and urine respectively.

The LLOQs were 740- (pUF), 1,500- (plasma), and 3,700- (urine) fold lower than the GF-

AAS method reported by Crul et al. [11], which dramatically increased the

pharmacokinetic window for evaluation of Ru concentrations after treatment with

NAMI-A.

50

75

100

125

0 5 10

Matrix (%)

Rela

tive

sign

al o

f 50

ng/L

ru

then

ium

(%)

Ru signal in pUFRu signal in plasmaRu signal in urine

Determination of ruthenium by ICP-MS

121

Carry-over

Although ICP-MS has a linear performance over nine orders of magnitude, the assay

described here did not allow us to fully utilise this range. The concentration of the

highest calibration sample was limited by the memory effect it produced and by the

desired LLOQ. The response of a blank solution following the ULOQ calibration standard

should be less than 20% of the LLOQ standard signal. Therefore, the highest possible

calibration standard and optimal rinse time had to be assessed.

To evaluate and minimise the effect of carry-over, we studied the signals of blank

readings following the ULOQ calibration sample (1.00x104 ng/L Ru diluted 100-fold) and

we optimised the rinse time. A 20 s rinse time with 1% HNO3 between two samples was

required to avoid a memory effect from the preceding high concentration sample and to

allow the determination of 300 ng/L (diluted 100-fold) in pUF and urine and 750 ng/L

(diluted 100-fold) in plasma.

Samples above the ULOQ should be diluted to fit them within the calibration range and

re-assayed. Dilutions up to 104 were validated (see Table 3).

Linearity

Seven non-zero calibration standards in a dynamic range of 300-1.00x104 ng/L of Ru in

pUF and urine and of 750-1.00x104 ng/L of Ru in plasma were processed and single

measurements were performed. For pUF, calibration samples were analysed in three

separate analytical runs. As a partial validation was carried out for urine and plasma,

calibration samples for these matrices were analysed in one analytical run.

The calibration curves were best described by linear regression, using 1/(relative

standard deviation (RSD) of triplicate sample reading) as weight-factor, to avoid bias in

favour of samples with high standard deviations or high concentrations. The calibration

concentrations were back-calculated from the responses. In Table 4, deviations for pUF,

plasma, and urine and relative standard deviations (RSD) for pUF are presented.

Deviations from the nominal concentration were between –2.09 and 3.24% for all

concentration levels, which all met the requirements of within ±20% for the LLOQ and

±15% for other concentrations [31]. RSDs for the calibration samples in pUF were less

than 5.17%. Correlation coefficients were better than > 0.9999.

Accuracy and precision

The accuracy and precision of the method were determined by assaying QC samples at

six concentration levels, with different dilution factors. For pUF, the accuracy, and

within- and between-run precision data were assessed by analysing five replicates of

each sample in three analytical runs. For plasma and urine, the accuracy and within-run

Chapter 2.3

122

precision data were determined by analysing five replicates of each sample in one

analytical run. The accuracy was expressed as a percentage of the nominal concentration

and it had to be within 80-120% for the LLOQ and within 85-115% for the other

concentrations. The within- and between-run precision were calculated by analysis of

variances (ANOVA) for each test concentration using the analytical run as the grouping

variable. The precision should not exceed ±20% for the LLOQ and ±15% for the other

concentrations [31].

The accuracy and precision data are summarised in Table 4. It can be seen from Table 4,

that at all QC concentration levels, the data were within the limits for bioanalytical

method validation. The different dilution factors of the QC samples did not affect the

performance of the method.

Table 3. Accuracy and precision for quality control samples in each biological matrix (n=5)

Matrix

Concentration in matrix (ng/L)

Dilution factor

Within-run precision (%)

Between-run precision (%)

Accuracy (%)

Plasma ultrafiltrate 30.0 10 1.93 3.79 101

300 100 3.70 1.14 105

750 100 2.01 0.97 105

2.50x103 100 1.08 1.12 104

7.50x103 100 0.663 0.634 104

1.00x105 1.00x104 1.27 -0.496 103

Plasma 75.0 10 3.76 * 111

750 100 0.921 * 107

1.50x103 100 2.14 * 105

2.50x103 100 1.07 * 108

7.50x103 100 2.10 * 107

1.00x105 1.00x104 3.68 * 106

Urine 30.0 10 1.99 * 108

300 100 2.86 * 108

750 100 1.74 * 107

2.50x103 100 0.958 * 106

7.50x103 100 0.897 * 104

1.00x105 1.00x104 1.72 * 109 *

A partial validation was performed for plasma and urine. Therefore no between-run precision was calculated.

Determination of ruthenium by ICP-MS

123

Table 4. Deviation from theoretical concentration (DEV %) for NAMI-A standards in pUF, plasma, and

urine and relative standard deviations (RSD %) (n=3) for NAMI-A standards in pUF

Matrix Concentration (ng/L)

Concentration in final matrix after 1:100 dilution (ng/L)

DEV (%) from nominal concentration

RSD (%)

Plasma ultrafiltrate 300 3.00 -1.57 1.96

600 6.00 3.24 1.58

1.00x103 10.0 -2.61 5.17

2.50x103 25.0 2.50 1.06

5.00x103 50.0 -0.160 0.768

7.50x103 75.0 -0.795 1.51

1.00x104 100 -1.09 1.11

Plasma 750 7.50 -0.660 *

1.00x103 10.0 -0.079 *

1.50x103 15.0 0.205 *

2.50x103 25.0 1.66 *

5.00x103 50.0 0.675 *

7.50x103 75.0 2.03 *

1.00x104 100 2.10 *

Urine 300 3.00 2.40 *

600 6.00 0.974 *

1.00x103 10.0 1.75 *

2.50x103 25.0 -1.85 *

5.00x103 50.0 -1.48 *

7.50x103 75.0 -1.06 *

1.00x104 100 -2.09 * *

A partial validation was performed for plasma and urine. Therefore no relative standard deviations for NAMI-A standards were

calculated.

Specificity

From six individual batches of drug-free pUF and urine, samples containing neither

analyte nor internal standard (blank), and samples containing 30.0 ng/L Ru and 200 ng/L

internal standard were prepared and diluted tenfold. These samples were prepared in

order to determine whether endogenous compounds interfered at the masses selected

Chapter 2.3

124

for Ru or internal standard in the most concentrated matrix used (tenfold dilution).

Similar solutions were prepared using six individual batches of drug-free plasma,

however, the samples containing Ru and internal standard were spiked with 75.0 ng/L

Ru. All samples were analysed in one analytical run. The signal of any interfering peak at

m/z 101 in blank solutions was not allowed to exceed 20% of the response of the LLOQ

standard. The response of any interfering peak at m/z 89 in the blank solution should not

exceed 5% of the response of 200 ng/L internal standard. The accuracies of the samples

spiked with Ru at the LLOQ standard level had to be within 80-120% of the nominal

value [31].

Blank samples from six individual batches did not show interferences from endogenous

material at the m/z selected for Ru with a response >20% of the LLOQ standard signal.

The response of interfering peaks at m/z 89 did not exceed 5% of the response of 200

ng/L internal standard. Deviations from the nominal concentrations at the LLOQ level

were between –4.20 and 10.3% (data not shown).

Cross analyte/internal standard interference test

Interference of the internal standard solution on the m/z 101 and interference of Ru on

m/z 89 had to be assessed. Drug-free pUF, plasma, and urine were spiked with NAMI-A

at ULOQ standard level and after a 100-fold dilution, the Y signal at m/z 89 was

monitored. The response of the interfering peak at m/z 89 was less than the maximum

allowed 5% of the response of 200 ng/L internal standard.

Drug-free pUF, plasma, and urine were diluted 1:100 and spiked with 200 ng/L Y. The

response of the interfering peak at m/z 101 was less than the maximum allowed 20% of

the response of the LLOQ standard.

Stability

The stability was evaluated in NAMI-A and Y stock and working solutions under both

processing (24 h at ambient temperatures) and storage (2-8 °C) conditions. The analytes

were considered stable in the stock and working solutions when 95-105% of the original

metal concentration was recovered.

For evaluation of stability of Ru concentrations in pUF, plasma, and urine during storage

and sample processing, two QC solutions (low and high) of each matrix were sampled.

Short-term stability of the analyte under processing conditions was evaluated by

comparing QC samples (stored for 24 h at ambient temperatures) with the

concentrations of these samples at time zero. To determine the stability of diluted QC

samples in the autosampler, samples at two concentration levels (low and high) were

analysed during one day and concentrations were compared with concentrations

measured at the start of the run.

Determination of ruthenium by ICP-MS

125

The stability after three freeze-thaw cycles was investigated by comparing QC samples

that had been frozen (-20 °C) and thawed three times with the concentrations of these

samples at t=0. Finally, long-term storage stability was assessed by determining the Ru

concentration in QC samples before and after four months storage at -20 °C.

The analytes were considered stable in the (un)diluted biological matrix when 85-115%

of the initial concentration was recovered.

The stability experiments showed that Ru and Y levels in stock and working solutions

remained constant under processing conditions and during storage for at least four

months. The colour of the NAMI-A stock solutions, however, changed from deep yellow-

orange to brown, which was probably caused by hydrolysis accompanied by the

dissociation of the dimethylsulfoxide (DMSO) ligand [37]. The observation that Ru

concentrations measured in freshly prepared stock solutions and stock solutions under

processing and storage conditions were similar indicated that the change in metal

speciation did not alter the ICP-MS signal.

Ru levels in QC samples remained constant under all tested conditions. No significant

changes in concentration were observed after three freeze/thaw cycles, nor after 24 h at

ambient temperatures, after one day in the autosampler, or after four months of storage.

Long-term stability has now been established for four months, but further testing is still

ongoing.

Application of the ICP-MS assay

The clinical applicability of the assay was demonstrated by analysis of human pUF and

plasma samples from patients who received 2.4 mg/m2 and 78 mg/m2 NAMI-A,

administered as a 3 h intravenous infusion. Plasma samples were collected prior to

administration and up to 24 h after start of the infusion. Ru concentration versus time

profiles for NAMI-A in human plasma and pUF are presented in Figure 3 and 4

respectively. Non-compartmental pharmacokinetic parameters are listed in Table 5. In

the Figures, ICP-MS data are compared to GF-AAS data. For the 78 mg/m2 administration

level, all plasma samples and a number of the pUF samples could be analysed by the two

methods. When ICP-MS data were plotted versus GF-AAS data, data were symmetrically

located around the line of unity, and the correlation coefficient was 0.999. This

demonstrates excellent uniformity of the techniques. For a number of the pUF samples

of the 78 mg/m2 administration level and for all plasma and pUF samples of the 2.4

mg/m2 administration level, no GF-AAS data were available because plasma and pUF

concentrations were lower than the LLOQ of the GF-AAS method [4]. By ICP-MS,

however, it was possible to analyse samples of both dose levels at all time points. This

shows the advantage of ICP-MS over GF-AAS.

Chapter 2.3

126

0.00

0.01

0.10

1.00

10.00

100.00

1000.00

10000.00

0 5 10 15 20 25Time (h)

Ruth

eniu

m c

once

ntra

tion

(µg/

L)

Plasma GF-AAS 78 mg/m2Plasma ICP-MS 78 mg/m2Plasma ICP-MS 2.4 mg/m2

Figure 3. Ru concentration versus time profiles in human plasma after intravenous infusion of NAMI-A.

Plasma samples of 78 mg/m2 were analysed by GF-AAS and ICP-MS. Plasma samples of 2.4 mg/m2

were analysed by ICP-MS.

0.00

0.01

0.10

1.00

10.00

100.00

1000.00

0 5 10 15 20 25Time (h)

Ruth

eniu

m c

once

ntra

tion

(µg/

L)

PUF GF-AAS 78 mg/m2PUF ICP-MS 78 mg/m2PUF ICP-MS 2.4 mg/m2

Figure 4. Ru concentration versus time profiles in human pUF after intravenous infusion of NAMI-A.

Eight of the pUF samples of the 78 mg/m2 level were analysed by GF-AAS and all pUF samples were

analysed by ICP-MS. PUF samples of 2.4 mg/m2 were analysed by ICP-MS.

Determination of ruthenium by ICP-MS

127

Table 5. Summary of non-compartmental pharmacokinetic parameters for Ru in plasma ultrafiltrate

and plasma

Matrix Dose level

(mg/m2/day)

Cmax

(mg/L)

AUC 0-24

(mgxh/L)

t1/2

(h)

Plasma ultrafiltrate 2.4 0.002 0.013 7.21

78 0.145 0.693 8.90

Plasma 2.4 0.116 1.91 33.8

78 2.97 53.4 49.9

Conclusion

A highly sensitive ICP-MS assay for the reliable and fast quantitative determination of Ru

originating from NAMI-A in human pUF, plasma, and urine, was developed.

Subsequently, the assay was validated according to current FDA guidelines. The

validated range were from 300 to 1x105 ng/L Ru in human pUF and urine, and from 750

to 1x105 ng/L Ru in human plasma. The LLOQ was lowered from 300 to 30.0 ng/L for Ru

in pUF and urine and from 750 to 75.0 ng/L for Ru in plasma by including a tenfold

dilution into the validation procedures. A dilution factor of 104 was validated, resulting in

excellent accuracy and precision data. The assay is now successfully applied to support

pharmacokinetic studies of patients treated with NAMI-A.

Chapter 2.3

128

References 1. Hartmann JT, Lipp HP. Toxicity of platinum compounds. Expert Opin Pharmacother 2003; 4: 889-901.

2. Rabik CA, Dolan ME. Molecular mechanisms of resistance and toxicity associated with platinating agents. Cancer Treat Rev 2007; 33: 9-23.

3. Kartalou M, Essigmann JM. Mechanisms of resistance to cisplatin. Mutat Res 2001; 478: 23-43.

4. Rademaker-Lakhai JM, van den BD, Pluim D, Beijnen JH, Schellens JH. A Phase I and pharmacological study with imidazolium-trans-DMSO-imidazole-tetrachlororuthenate, a novel ruthenium anticancer agent. Clin Cancer Res 2004; 10: 3717-27.

5. Hartinger CG, Zorbas-Seifried S, Jakupec MA, Kynast B, Zorbas H, Keppler BK. From bench to bedside--preclinical and early clinical development of the anticancer agent indazolium trans-[tetrachlorobis(1H-indazole)ruthenate(III)] (KP1019 or FFC14A). J Inorg Biochem 2006; 100: 891-904.

6. Scheulen ME, Kynast B, Jaehde U, Hilger RA, Scheuch E, Arion V, Kupsch P, Henke MM, Dittrich C, Keppler BK. A phase I dose-escalation trial with the new redox activated compound sodium trans-[tetrachlorobis(1H-indazole)ruthenate(III)]/indazolhydrochloride (1:1.1) (FFC14A) in patients with solid tumours - a CESAR study. J Clin Oncol , 2004 ASCO Annual Meeting Proceedings 2004; 22: 2101.

7. Sava G, Capozzi I, Clerici K, Gagliardi G, Alessio E, Mestroni G. Pharmacological control of lung metastases of solid tumours by a novel ruthenium complex. Clin Exp Metastasis 1998; 16: 371-9.

8. Sava G, Gagliardi R, Bergamo A, Alessio E, Mestroni G. Treatment of metastases of solid mouse tumours by NAMI-A: comparison with cisplatin, cyclophosphamide and dacarbazine. Anticancer Res 1999; 19: 969-72.

9. Sava G, Clerici K, Capozzi I, Cocchietto M, Gagliardi R, Alessio E, Mestroni G, Perbellini A. Reduction of lung metastasis by ImH[trans-RuCl4(DMSO)Im]: mechanism of the selective action investigated on mouse tumours. Anticancer Drugs 1999; 10: 129-38.

10. Galanski M, Arion VB, Jakupec MA, Keppler BK. Recent developments in the field of tumour-inhibiting metal complexes. Curr Pharm Des 2003; 9: 2078-89.

11. Crul M, van den Bongard HJ, Tibben MM, van Tellingen O, Sava G, Schellens JH, Beijnen JH. Validated method for the determination of the novel organo-ruthenium anticancer drug NAMI-A in human biological fluids by Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 2001; 369: 442-5.

12. Vouillamoz-Lorenz S, Bauer J, Lejeune F, Decosterd LA. Validation of an AAS method for the determination of platinum in biological fluids from patients receiving the oral platinum derivative JM216. J Pharm Biomed Anal 2001; 25: 465-75.

13. Brouwers EEM, Tibben MM, Joerger M, van Tellingen O, Rosing H, Schellens JHM, Beijnen JH. Determination of oxaliplatin in human plasma and plasma ultrafiltrate by graphite-furnace atomic-absorption spectrometry. Anal Bioanal Chem 2005; 382: 1484-90.

14. van Warmerdam LJC, van Tellingen O, Maes RAA, Beijnen JH. Validated method for the determination of carboplatin in biological fluids by Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 1995; 351: 777-81.

15. Kloft C, Appelius H, Siegert W, Schunack W, Jaehde U. Determination of platinum complexes in clinical samples by a rapid flameless atomic absorption spectrometry assay. Ther Drug Monit 1999; 21: 631-7.

16. LeRoy AF, Wehling ML, Sponseller HL, Friauf WS, Solomon RE, Dedrick RL, Litterst CL, Gram TE, Guarino AM, Becker DA. Analysis of platinum in biological materials by flameless atomic absorption spectrophotometry. Biochem Med 1977; 18: 184-91.

17. Meerum Terwogt JM, Tibben MM, Welbank H, Schellens JHM, Beijnen JH. Validated method for the determination of platinum from a liposomal source (SPI-77) in human plasma using graphite furnace Zeeman atomic absorption spectrometry. Fresenius J Anal Chem 2000; 366: 298-302.

18. Tibben MM, Rademaker-Lakhai JM, Rice JR, Stewart DR, Schellens JHM, Beijnen JH. Determination of total platinum in plasma and plasma ultrafiltrate, from subjects dosed with the platinum-containing N-(2-hydroxypropyl)methacrylamide copolymer AP5280, by use of graphite-furnace Zeeman atomic-absorption spectrometry. Anal Bioanal Chem 2002; 373: 233-6.

19. Gray AL. Mass-spectrometric analysis of solutions using an atmosperic pressure ion source. The Analyst 1975; 100: 289-99.

20. Tothill P, Matheson LM, Smyth JF. Inductively coupled plasma mass spectrometry for the determination of platinum in animal tissue and a comparison with atomic absorption spectrometry. J Anal Atom Spectrom 1990; 5: 619-22.

21. Allain P, Berre S, Mauras Y, Le Bouil A. Evaluation of inductively coupled mass spectrometry for the determination of platinum in plasma. Biol Mass Spectrom 1992; 21: 141-3.

Determination of ruthenium by ICP-MS

129

22. Gamelin E, Allain P, Maillart P, Turcant A, Delva R, Lortholary A, Larra F. Long-term pharmacokinetic behavior of platinum after cisplatin administration. Cancer Chemother Pharmacol 1995; 37: 97-102.

23. Sessa C, Capri G, Gianni L, Peccatori F, Grasselli G, Bauer J, Zucchetti M, Vigano L, Gatti A, Minoia C, Liati P, Van den BS, Bernareggi A, Camboni G, Marsoni S. Clinical and pharmacological phase I study with accelerated titration design of a daily times five schedule of BBR3464, a novel cationic triplatinum complex. Ann Oncol 2000; 11: 977-83.

24. Morrison JG, White P, McDougall S, Firth JW, Woolfrey SG, Graham MA, Greenslade D. Validation of a highly sensitive ICP-MS method for the determination of platinum in biofluids: application to clinical pharmacokinetic studies with oxaliplatin. J Pharm Biomed Anal 2000; 24: 1-10.

25. Liu J, Kraut E, Bender J, Brooks R, Balcerzak S, Grever M, Stanley H, D'Ambrosio S, Gibson-D'Ambrosio R, Chan KK. Pharmacokinetics of oxaliplatin (NSC 266046) alone and in combination with paclitaxel in cancer patients. Cancer Chemother Pharmacol 2002; 49: 367-74.

26. Bettinelli M, Spezia S, Ronchi A, Minoia C. Determination of Pt in biological fluids with ICP-MS: Evaluation of analytical uncertainty. Atom Spectrosc 2004; 25: 103-11.

27. Brouwers EEM, Tibben MM, Rosing H, Hillebrand MJX, Joerger M, Schellens JHM, Beijnen JH. Sensitive inductively coupled plasma mass spectrometry assay for the determination of platinum originating from cisplatin, carboplatin, and oxaliplatin in human plasma ultrafiltrate. J Mass Spectrom 2006; 41: 1186-94.

28. Szpunar J, Makarov A, Pieper T, Keppler BK, Lobinski R. Investigation of metallodrug-protein interactions by size-exclusion chromatography coupled with inductively coupled plasma mass spectrometry (ICP-MS). Anal Chim Acta 1999; 387: 135-44.

29. Hartinger CG, Hann S, Koellensperger G, Sulyok M, Groessl M, Timerbaev AR, Rudnev AV, Stingeder G, Keppler BK. Interactions of a novel ruthenium-based anticancer drug (KP1019 or FFC14a) with serum proteins--significance for the patient. Int J Clin Pharmacol Ther 2005; 43: 583-5.

30. Sulyok M, Hann S, Hartinger CG, Keppler BK, Stingeder G, Koellensperger G. Two dimensional separation schemes for investigation of the interaction of an anticancer ruthenium(III) compound with plasma proteins. J Anal Atom Spectrom 2005; 20: 856-63.

31. U.S.Food and Drug Administration, Center for Drug Evaluation and Research, Guidance for Industry, Bioanalytical Method Validation. FDA. U.S. Food and Drug Administration: Center for Drug Evaluation and Research: Guidance for Industry: Bioanalytical Method Validation. http://www fda gov/cder/guidance/ 4252fnl pdf 2001.

32. Rodushkin I, Engstrom E, Stenberg A, Baxter DC. Determination of low-abundance elements at ultra-trace levels in urine and serum by inductively coupled plasma-sector field mass spectrometry. Anal Bioanal Chem 2004; 380: 247-57.

33. Rosman KJR, Taylor PDP. Isotopic compositions of the elements 1997; International Union of Pure and Applied Chemistry. http://www physics curtin edu au/iupac/docs/Final97 pdf 1997.

34. Goulle JP, Mahieu L, Castermant J, Neveu N, Bonneau L, Laine G, Bouige D, Lacroix C. Metal and metalloid multi-elementary ICP-MS validation in whole blood, plasma, urine and hair Reference values. Forensic Sci Int 2005; 153: 39-44.

35. Forte G, Bocca B, Senofonte O, Petrucci F, Brusa L, Stanzione P, Zannino S, Violante N, Alimonti A, Sancesario G. Trace and major elements in whole blood, serum, cerebrospinal fluid and urine of patients with Parkinson's disease. J Neural Transm 2004; 111: 1031-40.

36. McCurdy E, Potter D. Optimising ICP-MS for the determination of trace metals in high matrix samples. Spectroscopy Europe 2001; 13: 10-3.

37. Sava G, Bergamo A, Zorzet S, Gava B, Casarsa C, Cocchietto M, Furlani A, Scarcia V, Serli B, Iengo E, Alessio E, Mestroni G. Influence of chemical stability on the activity of the antimetastasis ruthenium compound NAMI-A. Eur J Cancer 2002; 38: 427-35.

Chapter 3

Determination of platinum-DNA adducts

Chapter 3.1

Inductively coupled plasma mass spectrometric analysis of the total amount

of platinum-DNA adducts in peripheral blood mononuclear cells and tissue from

patients treated with cisplatin

Elke E.M. Brouwers Matthijs M. Tibben

Dick Pluim Hilde Rosing

Henk Boot Annemiek Cats

Jan H.M. Schellens Jos H. Beijnen

Submitted for publication

Chapter 3.1

134

Abstract

We present a highly sensitive method for the determination of platinum-DNA (Pt-DNA)

adducts in peripheral blood mononuclear cells and tissue samples from patients treated

with the anticancer agent cisplatin. The method is based on the measurement of

platinum by inductively coupled plasma mass spectrometry (ICP-MS) and allows

quantification of Pt-DNA adducts in PBMCs isolated from 10 mL of blood and in 1 mg of

tissue. The lower limit of quantification is 0.75 pg Pt or 7.5 fg Pt/µg DNA when using 100

µg DNA.

The method proved to be accurate and precise. The results obtained using the ICP-MS

method were in good agreement with results from the alternative 32P-postlabeling assay.

The ICP-MS method was, however, more sensitive than the 32P-postlabeling assay. In

addition, the ICP-MS method proved to be less laborious.

The advantages of the presented ICP-MS technique were demonstrated by the analysis

of PBMCs, normal gastric tissue, and gastric tumour tissue of patients treated with

cisplatin. This method for the analysis of Pt-DNA adducts in tissue samples allows us to

study adduct levels in biopsy samples e.g. from fine needle aspirates and to investigate

the distribution of adducts across a tumour sample.

Analysis of Pt-DNA adducts

135

Introduction

After the discovery of the antiproliferative effects of cisplatin (cis-

diaminedichloridoplatinum(II)) in the 1960s [1], the drug has developed successfully into

one of the most commonly used anticancer agents. The mechanism of action of cisplatin

is still not completely understood. It is, however, generally accepted that DNA

platination is the ultimate event in the cytotoxic activity of cisplatin. The hydrolysed

products of cisplatin attack the nucleophilic N7 positions from guanine (G) and adenine

(A) leading to the formation of various platinum-DNA (Pt-DNA) adducts, which affect the

DNA replication and transcription and thereby inhibit the tumour growth.

The interest in the formation of Pt-DNA adducts has increased the demand for analytical

methods to quantify these adducts in biological matrices. Evidently, tumour tissue is the

most relevant matrix to study. This matrix, however, is rather difficult to obtain.

Therefore, Pt-DNA adduct levels in peripheral blood mononuclear cells (PBMCs) are

often used as a surrogate marker for adduct levels in tumour tissue. Previous reports link

the adduct levels in PBMCs to clinical activity [2,3] and it was postulated that adduct

levels in PBMCs could predict chemosensitivity. Other investigations, comparing adduct

levels in PBMCs to levels in tumour tissue for head and neck carcinoma [4] and testicular

cancer [5], showed the lack of a relationship between Pt adduct levels in PBMCs and

tumour tissue. Hoebers et al. reported that adduct levels in PBMCs were 4-5 fold lower

than in tumour tissue and that adduct levels in tumour tissue were predictive for

treatment outcome [4]. These results indicate that in addition to the analysis for Pt-DNA

adducts in PBMCs, levels in tumour tissue should be investigated.

The limited availability of tissue and the fact that only 1% of the cisplatin molecules that

enter the cell actually bind to the DNA [6,7] illustrate the need for highly sensitive

techniques with the potential to detect low levels of Pt in limited amounts of DNA. The 32P-postlabeling method shows this high sensitivity [8], but it is less suitable for routine

clinical diagnostics due to its high radioactivity and its complex, labour intensive sample

pretreatment procedure. A less complex technique is inductively coupled plasma mass

spectrometry (ICP-MS), which, currently, is the most sensitive technique for the

determination of Pt originating from anticancer agents and is frequently used to study

Pt levels in various biological matrices. ICP-MS was applied previously for the

determination of Pt-DNA adducts in PBMCs [9,10] and tissue from patients [11] and

rodents [12,13]. The technique was also used to measure the amount of Pt-DNA adducts

in various cell types after incubation with Pt compounds [14,15]. For only a few assays,

however, the validation has been described [9,14,15]. The most sensitive assay was

described by Yamada et al. [14]. They were able to determine an absolute amount of 2

pg Pt.

The aim of the current work was to develop and validate a sensitive and reliable method

for the determination of Pt bound to DNA in PBMCs and tissue of patients after

Chapter 3.1

136

administration of cisplatin. Hence, we developed a sensitive ICP-MS assay requiring only

10 mL of blood or 1 mg of tissue. The method proved to be accurate and precise. The

usefulness of the technique was demonstrated in PBMCs, normal gastric tissue, and

gastric tumour tissue of patients treated with cisplatin. We compared the results of ICP-

MS with the widely applied 32P-postlabeling assay [8].

Experimental

Chemicals

Cisplatin reference standard was purchased from Calbiochem (San Diego, CA, USA).

Nitric acid (HNO3) 70% Ultrex II ultrapure reagent was obtained from Mallinckrodt Baker

(Philipsburg, NJ, USA). Water used for the ICP-MS analysis was sterile water for injection

(Aqua B. Braun Medical, Melsungen, Germany). Proteinase K and sodium dodecylsulfate

(SDS) were acquired from Sigma-Aldrich (Steinheim, Germany). Sodium chloride (NaCl),

edta disodium salt, potassium hydrogencarbonate (KHCO3), ammonium chloride (NH4Cl)

and iridium chloride were purchased from Merck (Darmstadt, Germany). Ammonium

hydrogencarbonate (NH4HCO3) was purchased from VWR (Fontenay-sous-Bois, France).

Calf thymus DNA, tris-HCl, phosphate buffered saline (PBS), and triton X-100 were

acquired from Sigma-Aldrich (St. Louis, MO, USA). Absolute ethanol was obtained from

Biosolve (Valkenswaard, the Netherlands).

Instruments

Pt analyses were performed on an ICP-quadrupole-MS (Varian 810-MS) equipped with a

90° reflecting ion mirror (Varian, Mulgrave, Victoria, Australia). The sample introduction

system consisted of a Micromist glass low-flow nebuliser (sample uptake 0.4 mL/min), a

peltier-cooled (4 °C) double pass glass spray chamber and a quartz torch. Sample

transport from the SPS-3 autosampler (Varian) to the nebuliser was performed using a

peristaltic pump. The instrument was cooled by using a Kühlmobil 142 VD (Van der

Heijden, Dörentrup, Germany). Hoek Loos (Schiedam, The Netherlands) provided argon

gas (4.6) with a 99.996% purity. The instrument settings of the ICP-MS are listed in Table

1. Data were acquired and processed using the ICP-MS Expert Software version 1.1 b49

(Varian). Further data handling was performed using Excel 2000 (Microsoft, Redmond,

WA, USA).

The Pt isotope used for calculation of Pt concentrations was 194Pt. Pt determination at this

isotope can be subject to the interference of hafnium-oxides [16]. But, since oxide

formation and observed hafnium counts were low in all of the analysed samples, these

oxides were considered insignificant and no corrections were necessary. The detection

Analysis of Pt-DNA adducts

137

mode was based on peak jumping with peak dwell times of 50 ms, 25 scans per

replicate, and three replicates per sample. Internal standardisation was performed on

each replicate using iridium (191Ir). Quantification was based on the mean concentration

of three replicates analysed against a calibration curve using weighted linear regression

analysis with 1/%RSD as the weight factor.

Table 1. ICP-MS instrument settings

Flow parameters (L/min) Ion optics (volts)

Plasma flow 18.0 First extraction lens -12

Auxiliary flow 1.65 Second extraction lens -220

Sheath gas 0.25 Third extraction lens -230

Nebuliser flow 1.05 Corner lens -240

Mirror lens left 37

Torch alignment (mm) Mirror lens right 35

Sampling depth 5.0 Mirror lens bottom 20

Entrance lens 5

Other Fringe bias -3

RF power (kW) 1.30 Entrance plate -50

Pump rate (mL/min) 0.2 Detector focus -500

Stabilisation delay 40 Pole bias 0.0

Calibration standards and quality control samples

Calibration standards containing 0.769 - 154 ng/L cisplatin in 1% HNO3 (corresponding

to 0.500 - 100 ng/L Pt) were prepared and analysed at the start of each analytical run.

Quality control (QC) samples at three concentration levels; 2.30, 15.4, and 108 ng/L

cisplatin (corresponding to 1.50, 10.0, 70.0 ng/L Pt) were used to check the analytical

runs. To 1.5 mL of calibration standard or QC sample, 15 µL of a 1.00x104 ng/L internal

standard solution was added. The lower limit of quantification (LLOQ) of the assay was

set at a concentration of 0.769 ng/L cisplatin in 1% HNO3 (corresponding to 0.500 ng/L

Pt). The analyte response at the LLOQ level was at least five times the response of a blank

1% HNO3 solution. Because the minimal sample volume for one single ICP-MS

measurement was 1.5 mL, the absolute LLOQ corresponded with 0.75 pg Pt, which was

lower than the LLOQ in previous investigations [9,14,15]. The Pt concentrations in DNA

were obtained as pg Pt/µg DNA. When 100 µg DNA samples were used, the LLOQ was

7.5 fg Pt/µg.

Chapter 3.1

138

Method development and validation procedures

A schematic outline of the full sample pretreatment procedure for the quantification of

Pt-DNA adducts is depicted in Figure 1. At first, the DNA hydrolysis and quantification of

Pt bound to DNA (step 11 and 12) were optimised and validated. Subsequently, the full

sample pretreatment procedure was validated for PBMCs isolated from 10 mL whole

blood and for various amounts of gastric tissue.

Optimisation and validation of the hydrolysis of DNA

Optimisation of the hydrolysis of DNA

The hydrolysis procedure was optimised by assessment of the concentration of HNO3

required for complete hydrolysis and by the evaluation of the hydrolysis time needed to

prevent matrix effects. Therefore, twenty millilitres of a 100 mg/L calf thymus DNA

solution (2,000 µg DNA) were incubated with 123 µg cisplatin (80 µg Pt) for 72 h at 37 °C.

The DNA was then precipitated with 100% ethanol, washed twice with 75% ethanol, and

dissolved in 20 mL of water. The resulting solution was diluted 10-fold, 100-fold, 103-fold,

and 104-fold with 100 mg/L Pt-free DNA to obtain solutions with variable amounts of Pt-

DNA adducts and constant concentrations of DNA. Subsequently, HNO3 was added to

aliquots of each solution (containing 35 µg DNA), to obtain HNO3 concentrations of 1%

(v/v) or 35% (v/v). These solutions were left at 70 °C and Pt concentrations were

determined after several incubation times (0, 1, 2, 4, 5.5, 7, 24, and 30 h). The hydrolysis is

considered to be complete when no matrix effect was observed and when the Pt

concentration had reached a constant level.

Precision of the Pt determination after reaction with DNA

The precision of the hydrolysis procedure was assessed by analysing different amounts

of Pt-DNA adducts in triplicate. Therefore, 60 µg (sample A), 300 µg (sample B), and 1,500

µg (sample C) of DNA were incubated with 18.5 µg cisplatin (12.0 µg Pt) in 3 mL solution

for 72 h at 37 °C. The DNA was then precipitated with 100% ethanol, washed twice with

75% ethanol, and dissolved in 3 mL of water. The resulting solutions were diluted 10-

fold, 100-fold, 103-fold, 104-fold, and 105-fold with solutions containing equivalent

concentrations of DNA to obtain three series of solutions (A, B, and C) with variable

amounts of Pt-DNA adducts and constant concentrations of DNA. Subsequently, 1% (v/v)

HNO3 was added to aliquots of sample A, B, and C, containing respectively 7, 35, and 175

µg of DNA. These solutions were left at 70 °C for 24 h. The hydrolysis was performed in

triplicate. The precision was defined as the relative standard deviation of a triplicate

sample.

Analysis of Pt-DNA adducts

139

Whole blood Tumour tissue

Isolated cells

DNA

Platinum analysis ICP-MS

Figure 1. Schematic outline of sample pretreatment procedure for PBMCs and tumour tissue

Accuracy of the Pt determination after reaction with DNA

Because no reference material of Pt-DNA adducts is available, the accuracy of the Pt

determination after hydrolysis had to be assessed in an alternative way. Various amounts

of DNA (40, 200, and 1,000 µg) were incubated with 0.123, 1.23, and 12.3 µg cisplatin

(0.080, 0.800, and 8.00 µg Pt) in 2 mL solution for 72 h at 37 °C in duplicate.

Subsequently, HNO3 was added to the solutions to obtain HNO3 concentrations of 1%

Chapter 3.1

140

(v/v). These solutions were hydrolysed for 24 h at 70 °C. The resulting solutions vary in

the amount of Pt-DNA adducts, unreacted cisplatin, and DNA. Because the exact amount

of total Pt was known, the accuracy of the Pt determination in the presence of variable

Pt-adduct and DNA concentrations could be assessed. The accuracy was expressed as a

percentage of the Pt concentration that was added to the solution.

Validation of the determination of Pt-DNA adducts in PBMCs

Precision of the complete sample pretreatment procedure in PBMCs

The precision of the complete sample pretreatment procedure was evaluated by the

determination of Pt in DNA which was isolated in triplicate from PBMCs after incubation

of PBMCs with various amounts of cisplatin. Therefore, 30 millilitres of heparinised whole

blood from normal volunteers were treated with 2.15 (low), 21.5 (mid), and 215 (high) µg

cisplatin (1.40, 14.0, and 140 µg Pt) for 2.5 h at 37 °C. After incubation, the 30 mL were

divided equally over three 10 mL heparin-containing tubes (Becton Dickinson

Vacutainer Systems, Plymouth, UK). Following centrifugation for 15 min at 1,000 g and 4

°C, the PBMC fraction was isolated as described earlier [8,17]. Briefly, contaminating red

blood cells were lysed by incubation with 0.83% (w/v) NH4Cl, 0.1% (w/v) KHCO3, and 1

mM edta disodium salt for 20 min at 4 °C. PBMCs were washed twice with icecold PBS

and resuspended in 9 mL of a buffer containing 10 mM Tris-HCl, 2.3% (w/v) NaCl, and 2

mM edta disodium salt at pH 7.3.

DNA was isolated from PBMCs as described previously [8,17]. In short, 0.9 mL 1.1 M

NH4HCO3, 0.45 mL 20% (w/v) SDS, and 150 µL 1% (w/v) proteinase K solution were added

successively, followed by overnight incubation at 42 °C. After the digestion was

complete, 3.3 mL of saturated 6 M NaCl was added to each tube and the tubes were

shaken vigorously to precipitate proteins. The tubes were centrifuged and the

supernatant containing the DNA was transferred to another tube. Subsequently, the

supernatant was shaken, centrifuged, and transferred to another tube. Following this, 20

mL of absolute ethanol were added to precipitate the DNA. The DNA was washed twice

with 75% ethanol and was subsequently dissolved in 1 mL of water. DNA concentrations

were analysed after dilution in 10 mM Tris-HCl pH 8 by measuring the absorbance at 260

nm using a Biophotometer (Eppendorf, Hamburg, Germany). The purity of the DNA was

checked by determining the absorbance ratio at 260 and 280 nm. Ratios between 1.8

and 2.0 were routinely obtained.

The DNA was hydrolysed by incubation in 1% (v/v) HNO3 at 70 °C for 24 h. The resulting

solutions were diluted in 1% HNO3 to within the calibration range of the ICP-MS. A

volume of at least 1.5 mL should be available for one ICP-MS measurement. After the

addition of internal standard, samples were analysed by ICP-MS.

Analysis of Pt-DNA adducts

141

The precision at each cisplatin incubation level was defined as the relative standard

deviation of a triplicate sample.

Comparison of the ICP-MS analysis and the 32P- postlabeling assay

In addition to analysis with ICP-MS, DNA samples isolated from PBMCs were analysed

using the 32P-postlabeling assay. Therefore, DNA solutions were digested and purified as

described by Pluim et al [8]. Pt-DNA adducts were separated from the unmodified

nucleosides by strong cation exchange chromatography. Pt-DNA adducts were

subsequently deplatinated using NaCN followed by labeling with [γ-32P]ATP. The 32P-

labelled dinucleosides GG and AG were then separated by HPLC and quantified by

online radioisotope detection.

Validation of the determination of Pt-DNA adducts in tissue

Precision of the complete sample pretreatment procedure in tissue

Tissue biopsies of normal gastric and gastric tumour tissue of three gastric cancer

patients were used for the validation of the determination of Pt-DNA adduct in tissue.

Patients were treated with 60 mg/m2 cisplatin as a 4 h intravenous infusion. The Medical

Ethics Committee of the hospital approved the study protocol and all patients gave their

informed consent. Biopsy samples were acquired during a gastroscopy which was

performed approximately 24 h after the start of the infusion.

The acquired tissue samples, which weighed approximately 50 mg, were suspended and

homogenised in a buffer (containing 10 mM Tris-HCl, 2.3% (w/v) NaCl, and 2 mM edta

disodium salt at pH 7.3), so that the final tissue concentration was 20 mg/ml. Samples

were then divided in aliquots containing 12, 6, 3, and 1 mg tissue. DNA was isolated from

tissue as described for PBMCs with modifications taking into account the lower amount

of tissue available. The volumes of the reagents added are depicted in Table 2. After

hydrolysis, dilution, and the addition of internal standard, the levels of Pt-DNA adducts

were analysed. The precision of the method was defined as the relative standard

deviation of the adduct levels measured in the 12, 6, 3, and 1 mg sample.

Application of method

In addition to collection of tissue samples, patients donated whole blood samples up to

24 h after start of the infusion. From these samples, PBMCs were processed and Pt-DNA

adducts were determined with the ICP-MS method. Pt-DNA adduct levels in tissue

samples and PBMCs in samples drawn 24 h after the start of the infusion were compared.

Chapter 3.1

142

Table 2. Sample pretreatment of tissue

Amount of reagent added (µL) Amount of tissue (mg)

NH4HCO3 SDS Proteinase K NaCl Ethanol Water

12 60 30 10 220 1,320 200

6 30 15 5 110 660 200

3 15 7.5 2.5 55 330 150

1 15 7.5 2.5 55 330 150

Results and discussion

Optimisation and validation of the hydrolysis of DNA

Optimisation of the hydrolysis of DNA

Figure 2a illustrates the Pt recovery with time dependent hydrolysis of DNA on a semi-

logaritmic scale for all the solutions. Figure 2b shows the Pt recovery for the undiluted

solution on a normal scale. The hydrolysis in 1% HNO3 was complete after 24 h of

incubation. No difference in adduct concentration was observed between 24 h and 30 h

of incubation (Figure 2b). The Pt concentration in the solutions decreased linearly with

increasing dilution factors, indicating that the efficiency of the hydrolysis was not

affected by the amount of Pt-DNA adduct present in the solution. The decrease in Pt

concentration observed from 0 to 1 h of incubation can be explained by precipitation of

DNA in the 1 h sample, because 1% HNO3 was added to the 1 h sample, whereas no

HNO3 was added to the 0h sample. This resulted in a decrease of the dissolved Pt

concentration in the 1 h sample.

When 35% of HNO3 was used, constant platinum levels were already reached after 1 h of

incubation (data not shown) and Pt concentrations were similar to those after

incubation using 1% HNO3 for 24 h. However, incubation of adduct solutions using 1%

HNO3 for 24 h was preferred, because 1% HNO3 could be analysed by ICP-MS directly in

contrast to 35% HNO3 samples, resulting in a LLOQ.

Analysis of Pt-DNA adducts

143

1.0E+00

1.0E+02

1.0E+04

1.0E+06

1.0E+08

0 10 20 30 40

Time (h)

Plat

inum

con

cent

ratio

n (n

g/L)

undiluted

10-fold diluted

100-fold diluted

1,000-fold diluted

10,000-fold diluted

Figure 2a. Pt recovery with time dependent hydrolysis of DNA using 1% HNO3 (v/v) (logarithmic scale)

5.0E+05

6.0E+05

7.0E+05

8.0E+05

9.0E+05

1.0E+06

1.1E+06

1.2E+06

0 10 20 30 40

Time (h)

Plat

inum

con

cent

ratio

n (n

g/L)

undiluted

Figure 2b. Pt recovery with time dependent hydrolysis of DNA using 1% HNO3 (v/v) (normal scale)

Precision of the Pt determination after reaction with DNA

The precision of the hydrolysis procedure and Pt determination ranged between 0.62

and 11.5% for all Pt-DNA adduct levels and amounts of DNA (Table 3). In addition, the Pt

concentrations within each of the three series of DNA solutions were linear, indicating

that the hydrolysis efficiency was independent from the ratio of Pt-DNA adducts versus

unbound DNA.

Chapter 3.1

144

Table 3. Precision* and linearity of the hydrolysis procedure and the determination of Pt

Sample Amount of DNA (µg)

Dilution Average amount of recovered Pt-DNA adduct (pg Pt/µg DNA)

Precision (RSD (%), n=3)

7 0 1.03 x104 4.03

7 10-fold 1.02 x103 4.75

7 100-fold 9.94 x101 5.39

7 103-fold 1.03 x101 1.42

7 104-fold 1.04 x100 5.38

A

7 105-fold 9.26 x10-2 3.10

35 0 8.69 x103 2.78

35 10-fold 8.62 x102 8.94

35 100-fold 8.96 x101 0.616

35 103-fold 9.17 x100 3.01

35 104-fold 9.36 x10-1 3.93

B

35 105-fold 8.16 x10-2 11.5

175 0 4.86 x103 6.45

175 10-fold 4.98 x102 0.560

175 100-fold 4.86 x101 11.1

175 103-fold 5.84 x100 1.79

C

175 104-fold 5.83 x10-1 4.32

175 105-fold 5.45 x10-2 1.20 *

The precision was defined as the relative standard deviation of a sample hydrolysed in triplicate

Accuracy of the Pt determination after reaction with DNA

The accuracy of the Pt determination in the presence of variable amounts of hydrolysed

DNA ranged between 89.8 and 99.9% for all solutions (Figure 3). No relationship

between the amount of DNA and the accuracy was observed. Furthermore, iridium

signals were constant during the measurements, indicating that no matrix effects were

present. These results confirmed that the DNA hydrolysis was complete and that a

complete recovery of Pt was achieved.

Analysis of Pt-DNA adducts

145

0

20

40

60

80

100

120

140

0 40 200 1000

Amount of DNA (µg)

Acc

urac

y (%

)

Cisplatin Low

Cisplatin Mid

Cisplatin High

Figure 3. The accuracy of Pt determination in the presence of variable amounts of hydrolysed DNA

Validation of the determination of Pt-DNA adducts in PBMCs

Precision of the complete sample pretreatment procedure

Pt-DNA adduct concentrations and relative standard deviations for the determination of

adducts from PBMCs isolated from 10 mL whole blood are summarised in Table 4. The

sample with the lowest amounts of Pt-DNA adducts contained 35.6 fg Pt/µg DNA, which

was still five-fold higher than the LLOQ of the method. The precision for all

concentration levels ranged between 2.36 and 5.54%. These values are excellent,

especially considering the fact that the precision values are affected by both the

uncertainty in the determination of the DNA concentration and the uncertainty in the Pt

determinations.

Chapter 3.1

146

Table 4. Precision of Pt-DNA adduct determination in PBMCs using ICP-MS and 32P-postlabeling

Sample ICP-MS 32P-postlabeling

Total Pt-DNA adduct Pt-DNA adduct (fmol/µg DNA)

(fg Pt /µg DNA)

(fmol Pt /µg DNA)

RSD (%)

Pt-GG Pt-AG Pt-GG + Pt-AG

RSD (%)

Deviation from ICP-MS (%)

Low 35.6 0.182 4.72 0.161 <LLOQ 0.161 8.03 11.5

Mid 309 1.58 2.36 1.48 0.224 1.70 3.93 8.00

High 3,240 16.6 5.54 14.1 1.78 15.9 2.02 4.20

Comparison of the ICP-MS analysis and the 32P- postlabeling assay

Results for the Pt-DNA adduct determination using the 32P-postlabeling assay are also

depicted in Table 4. The adduct levels assessed by the 32P-postlabeling assay were

expected to be approximately 90% of the ICP-MS results because the 32P-assay

exclusively quantifies Pt-GG and Pt-AG in contrast to ICP-MS, which analyses the total

amount of Pt-DNA adducts. The results of the two methods were in close agreement.

The 32P-assay resulted in adduct levels which were 88.5, 108, and 95.8% of the adduct

levels analysed by ICP-MS for the low, mid, and high concentrated samples, respectively.

For the low concentrated sample, Pt-AG could not be quantified because the

concentration was lower than the LLOQ (0.053 fmol Pt-AG/µg DNA in samples

containing 100 µg DNA). Precision data for the 32P-postlabeling assay were within the

same range as for the ICP-MS method.

The LLOQ for the total adduct determination by ICP-MS was 7.5 fg Pt/µg DNA (or 0.038

fmol adduct/µg DNA), which was lower than the LLOQ for the 32P-assay which had LLOQs

of 0.087 fmol adduct/µg DNA and 0.053 fmol adduct/µg DNA for Pt-GG and Pt-AG

respectively [8].

Validation of the determination of Pt-DNA adducts in tissue

Precision of the complete sample pretreatment procedure in tissue

Recovered amounts of DNA, Pt-DNA adduct concentrations, and relative standard

deviations for the determination of adducts from various amounts of tissue samples are

given in Table 5. The sample with the lowest amount of tissue (1 mg) contained on

average 4.82 µg DNA. The Pt-DNA adduct concentration in this sample was on average

2.20 pg Pt µg DNA (10.6 pg Pt absolute), which was still 14-fold higher than the LLOQ

(0.75 pg Pt absolute). The precisions ranging from 0.930 to 11.0% indicate that the

results were not affected by the amount of tissue sample that was processed. These

Analysis of Pt-DNA adducts

147

results demonstrate that the quantification of Pt-DNA adducts in only 1 mg of tissue

leads to excellent results.

Application of the method

Figure 4 shows the Pt-DNA adduct concentrations analysed in PBMCs up to 24 h after

the start of the cisplatin infusion. Comparison of adduct levels in tissue samples and

PBMCs (Table 5) shows that gastric tumour tissue contained 4.3- to 7.7-fold higher

adduct levels than PBMCs, whereas normal gastric tissue contained 4.2- to 6.3-fold

higher adduct levels than PBMCs. The differences in adduct levels could be a result of

the physical environment and cellular characteristics that differ substantially between

tissue and PBMCs. To draw any definite conclusions from these results, however, more

patients will need to be studied. The observation that in two patients the adduct levels

in tumour tissue were slightly (1.2-fold) higher than in normal tissue, could be explained

by the better blood perfusion of tumour tissue. To draw any conclusions, however, this

also has to be evaluated further in a larger sample set. The obvious hypothesis to test

would be that patients with a high level of Pt-DNA adducts in (gastric) tumour tissue

compared to the PBMC adduct level, have a higher response rate to Pt-containing

chemotherapy regimens.

0

100

200

300

400

500

600

700

800

0 5 10 15 20 25 30

Time (h)

Plat

inum

-DN

A b

indi

ng in

PBM

Cs (f

g/µg

) Patient 1

Patient 2

Patient 3

Figure 4. Pt-DNA adduct concentrations in PBMCs versus time

Chapter 3.1

148

Table 5. DNA recovery of tissue and precision of Pt-DNA adduct determination in tissue. Pt-DNA adduct levels in tissue were compared to Pt-DNA adduct levels in PBMCs at 24 h

Amount of tissue (mg)

Average amount of DNA (µg)

Patient 1

pg Pt/µg DNA

Patient 2

pg Pt/µg DNA

Patient 3

pg Pt/µg DNA

Tissue Normal Tumour Normal Tumour Normal Tumour

12 43.4 2.54 3.14 2.06 2.81 1.85 2.03

6 23.5 2.49 3.05 2.25 2.60 1.95 1.91

3 12.5 2.54 3.17 2.49 2.70 1.92 2.07

1 4.82 2.53 2.97 1.93 2.32 1.74 1.69

Mean tissue 2.52 3.08 2.18 2.61 1.86 1.92

Precision (RSD (%), n=4)

0.930 2.96 11.0 7.98 4.98 8.70

PBMC 0.402 0.454 0.448

Ratio tissue level to PBMC level

6.27 7.67 4.80 5.75 4.15 4.29

Conclusion

We have developed and validated an ICP-MS method for the determination of Pt-DNA

adducts in PBMCs and tissue samples. The method proved to be applicable for the

determination of Pt-DNA adducts in PBMCs isolated from 10 mL of blood and in 1 mg of

tissue. The possibility to analyse Pt-DNA adducts in extremely small tissue samples

creates the opportunity to study the levels of adducts in biopsy samples from e.g. fine

needle aspirates and to investigate the distribution of adducts across the tumour.

Analysis of Pt-DNA adducts

149

References 1. Rosenberg B, Vancamp L, Krigas T. Inhibition of cell division in escherichia coli by electrolysis products

from a platinum electrode. Nature 1965; 205: 698-9.

2. Schellens JH, Ma J, Planting AS, van der Burg ME, van Meerten E, Boer-Dennert M, Schmitz PI, Stoter G, Verweij J. Relationship between the exposure to cisplatin, DNA-adduct formation in leucocytes and tumour response in patients with solid tumours. Br J Cancer 1996; 73: 1569-75.

3. Reed E, Parker RJ, Gill I, Bicher A, Dabholkar M, Vionnet JA, Bostick-Bruton F, Tarone R, Muggia FM. Platinum-DNA adduct in leukocyte DNA of a cohort of 49 patients with 24 different types of malignancies. Cancer Res 1993; 53: 3694-9.

4. Hoebers FJ, Pluim D, Verheij M, Balm AJ, Bartelink H, Schellens JH, Begg AC. Prediction of treatment outcome by cisplatin-DNA adduct formation in patients with stage III/IV head and neck squamous cell carcinoma, treated by concurrent cisplatin-radiation (RADPLAT). Int J Cancer 2006; 119: 750-6.

5. Fichtinger-Schepman AM, van der Velde-Visser SD, Dijk-Knijnenburg HC, van Oosterom AT, Baan RA, Berends F. Kinetics of the formation and removal of cisplatin-DNA adducts in blood cells and tumor tissue of cancer patients receiving chemotherapy: comparison with in vitro adduct formation. Cancer Res 1990; 50: 7887-94.

6. Reedijk J. Why does Cisplatin reach Guanine-N7 with competing S-donor ligands available in the cell? Chem Rev 1999; 99: 2499-510.

7. Centerwall CR, Tacka KA, Kerwood DJ, Goodisman J, Toms BB, Dubowy RL, Dabrowiak JC. Modification and uptake of a cisplatin carbonato complex by Jurkat cells. Mol Pharmacol 2006; 70: 348-55.

8. Pluim D, Maliepaard M, van Waardenburg RC, Beijnen JH, Schellens JHM. 32P-postlabeling assay for the quantification of the major platinum-DNA adducts. Anal Biochem 1999; 275: 30-8.

9. Bonetti A, Apostoli P, Zaninelli M, Pavanel F, Colombatti M, Cetto GL, Franceschi T, Sperotto L, Leone R. Inductively coupled plasma mass spectroscopy quantitation of platinum-DNA adducts in peripheral blood leukocytes of patients receiving cisplatin- or carboplatin-based chemotherapy. Clin Cancer Res 1996; 2: 1829-35.

10. Liu J, Kraut E, Bender J, Brooks R, Balcerzak S, Grever M, Stanley H, D'Ambrosio S, Gibson-D'Ambrosio R, Chan KK. Pharmacokinetics of oxaliplatin (NSC 266046) alone and in combination with paclitaxel in cancer patients. Cancer Chemother Pharmacol 2002; 49: 367-74.

11. Cooper BW, Veal GJ, Radivoyevitch T, Tilby MJ, Meyerson HJ, Lazarus HM, Koc ON, Creger RJ, Pearson G, Nowell GM, Gosky D, Ingalls ST, Hoppel CL, Gerson SL. A phase I and pharmacodynamic study of fludarabine, carboplatin, and topotecan in patients with relapsed, refractory, or high-risk acute leukemia. Clin Cancer Res 2004; 10: 6830-9.

12. McDonald ES, Randon KR, Knight A, Windebank AJ. Cisplatin preferentially binds to DNA in dorsal root ganglion neurons in vitro and in vivo: a potential mechanism for neurotoxicity. Neurobiol Dis 2005; 18: 305-13.

13. Rice JR, Gerberich JL, Nowotnik DP, Howell SB. Preclinical efficacy and pharmacokinetics of AP5346, a novel diaminocyclohexane-platinum tumor-targeting drug delivery system. Clin Cancer Res 2006; 12: 2248-54.

14. Yamada K, Kato N, Takagi A, Koi M, Hemmi H. One-milliliter wet-digestion for inductively coupled plasma mass spectrometry (ICP-MS): determination of platinum-DNA adducts in cells treated with platinum(II) complexes. Anal Bioanal Chem 2005; 382: 1702-7.

15. Bjorn E, Nygren Y, Nguyen TT, Ericson C, Nojd M, Naredi P. Determination of platinum in human subcellular microsamples by inductively coupled plasma mass spectrometry. Anal Biochem 2007; 363: 135-42.

16. Lustig L, Zang S, Michalke B, Schramel P, Beck W. Platinum determination in nutrient plants by inductively coupled plasma mass spectrometry with special respect to the hafnium oxide interference. Fresenius J Anal Chem 1997; 357: 1157-63.

17. Ma J, Verweij J, Planting AS, Boer-Dennert M, van Ingen HE, van der Burg ME, Stoter G, Schellens JH. Current sample handling methods for measurement of platinum-DNA adducts in leucocytes in man lead to discrepant results in DNA adduct levels and DNA repair. Br J Cancer 1995; 71: 512-7.

Chapter 3.2

The effects of sulfur-containing compounds and gemcitabine on the

binding of cisplatin to plasma proteins and DNA determined by ICP-MS and

HPLC-ICP-MS

Elke E.M. Brouwers Alwin D.R. Huitema Jan H.M. Schellens

Jos H. Beijnen

Submitted for publication

Chapter 3.2

152

Abstract

The aim of this study was to investigate the effect of the compounds sodium thiosulfate

(STS), glutathione (GSH), acetylcysteine (AC), and gemcitabine on the platinum-protein

(Pt-protein) and platinum-DNA (Pt-DNA) binding of cisplatin in whole blood. This was

done to obtain more insight into the platinum (Pt) binding in whole blood and the

effects of modulators on this process.

STS, GSH, AC, and gemcitabine were added before and after the incubation of whole

blood with cisplatin. Pt levels in plasma and plasma ultrafiltrate and bound to DNA in

peripheral blood mononuclear cells were determined using inductively coupled plasma

mass spectrometry (ICP-MS). Additionally, information on the major Pt-DNA adducts was

obtained by separation of the Pt-DNA adducts by high performance liquid chromato-

graphy (HPLC) with off-line ICP-MS detection.

Results showed that the reactive Pt levels in whole blood are reduced by STS, GSH, and

AC. This reduction was demonstrated by a reduced Pt-protein and Pt-DNA binding in the

presence of sulfur-containing compounds. Furthermore, STS and AC appeared to be

able to release Pt from proteins. The compounds could hardly release Pt from the DNA.

Gemcitabine slightly inhibited Pt-DNA binding and did not alter Pt-protein binding. The

type of Pt-DNA adducts were found not altered in the presence of the modulators.

In conclusion, the results of the current study illustrate that STS, GSH, and AC affect the

Pt binding in whole blood, which suggests that these compounds could affect Pt-

binding in patients. By interfering with Pt-DNA and Pt-protein binding, the compounds

could influence side effects and cytotoxicity.

Effect of antidotes on platinum binding

153

Introduction

Cisplatin (cis-diamminedichloridoplatinum(II)) (Figure 1) is a successful anticancer drug

which is applied for the treatment of various malignancies. After intravenous infusion,

cisplatin and its reactive metabolites become rapidly partitioned into plasma protein-

bound platinum (Pt), free plasma Pt, tissue Pt, Pt in peripheral blood mononuclear cells

(PBMCs), and erythrocyte-sequestered Pt. As much as 60-95% of cisplatin and its reactive

metabolites bind to plasma proteins [1]. The role of the platinum-protein (Pt-protein)

complexes in the mechanism of cytotoxicity is, up to now, unknown and the free Pt

fraction is generally considered as pharmacologically active [2,3]. Part of this fraction

ultimately enters the cell and cell nucleus and binds to the DNA. Platinum-DNA (Pt-DNA)

adducts affect DNA replication and transcription and, thereby, inhibit tumour growth.

Figure 1. Structural formula cisplatin

Unfortunately, the application of cisplatin is impaired by severe side effects, such as

nephrotoxicity, ototoxicity, and neurotoxicity. In addition to the direct consequence of

these side effects, their persistent nature can seriously affect the patients quality of life.

The high incidence of severe side effects has led to the development of treatment

strategies aimed to prevent or reduce the side effects without affecting the antitumour

activity. Because cisplatin has high affinity for binding to sulfur donors, sulfur-containing

nucleophiles could serve as Pt neutralisers. Systemically administered sodium thiosulfate

(STS), nowadays, is commonly used in combination with locally administered cisplatin in

patients with e.g. head-neck carcinoma [4] and intraperitoneal tumours [5-7]. STS has a

protective effect against cisplatin induced nephrotoxicity [5] and ototoxicity [8].

Glutathione (GSH), an endogenous sulfur-containing compound, also provided

protection against cisplatin induced nephrotoxicity in patients when administered

intravenously [9]. In addition to STS and GSH, acetylcysteine (AC), a precursor of GSH,

could serve as a cisplatin neutralising agent. This compound decreased the toxicity in

experimental animals when administered in conjunction with cisplatin [10]. To our best

knowledge, to date, no patient studies have been performed on the protective effects of

AC against cisplatin induced toxicity. A major advantage of the use of STS, GSH, and AC

is that they are well tolerated in high doses. So far, sulfur-containing compounds were

always tested for their ability to prevent side effects. Previous studies, however, showed

that sulfur-containing compounds are capable of reversing the Pt-DNA [11] and Pt-

protein binding [12]. Because persistent side effects might be a consequence of Pt which

is accumulated in the body and remains bound to e.g. proteins and DNA, it is plausible,

that sulfur-containing compounds could reduce persistent side effects even when

Pt

Cl

Cl

NH3

NH3

Chapter 3.2

154

administered months after treatment without reducing cytotoxicity. To realise such

effects, however, the compounds should be capable of removing Pt from the cellular

and blood compartment. Up to now, this has not been investigated.

In addition to compounds that are administered deliberately to protect against cisplatin

induced toxicity, cytotoxic compounds that are administered with cisplatin in standard

combination regimens might also affect the cytotoxic action of cisplatin. Gemcitabine,

which was combined with cisplatin for the treatment of non-small cell lung cancer,

appeared to reduce the formation of Pt-DNA adducts in PBMCs [13].

The effect of the modulators STS, GSH, AC, and gemcitabine on Pt-protein binding and

the formation of Pt-DNA adducts can be evaluated in ex vivo studies. Thereby,

information is gained on the potential use of sulfur-containing compounds for the

prevention and reduction of side effects in vivo. In the current study, we investigated the

effect of STS, GSH, AC, and gemcitabine (Figure 2) on the Pt-protein and Pt-DNA binding

of cisplatin in whole blood. Pt concentrations in plasma and plasma ultrafiltrate (pUF)

were determined using a previously validated inductively coupled plasma mass

spectrometry (ICP-MS) method [14]. ICP-MS was also used to assess the total amount of

Pt-DNA adducts in PBMCs. Information on the major Pt-DNA adducts, i.e. Pt-GG

(intrastrand crosslink on pGpG sequences) and Pt-AG (intrastrand crosslink on pApG

sequences) was obtained by separation of the Pt-DNA adducts by high performance

liquid chromatography (HPLC) with off-line ICP-MS detection.

a b

c d

Figure 2. Structural formula of sodium thiosulfate (a), glutathione (b), acetylcysteine (c), and

gemcitabine HCl (d)

Na2S2O3

SH NHOH

O

O

HOH

O

NH

H NH2

O

SH O

OH

NH H

OCH3

N NH+

O

NH2

O

OH

OH F

F

Cl

Effect of antidotes on platinum binding

155

Experimental

Chemicals

Cisplatin reference standard was purchased from Calbiochem (San Diego, CA, USA).

Nitric acid (HNO3) 70% Ultrex II ultrapure reagent was obtained from Mallinckrodt Baker

(Philipsburg, NJ, USA). Water used for the analyses was sterile water for irrigation (Aqua

B. Braun Medical, Melsungen, Germany). Heparinised whole blood was obtained from

healthy volunteers. The STS solution (250 g/L) was supplied by the Hospital Pharmacy of

Haarlem (the Netherlands). AC was purchased from Zambon (Amersfoort, the

Netherlands). Gemcitabine HCl was obtained from Eli Lilly (Houten, the Netherlands).

Proteinase K and sodium dodecylsulfate (SDS) were acquired from Sigma-Aldrich

(Steinheim, Germany). Sodium chloride (NaCl), edta disodium and diammonium salt,

potassium hydrogencarbonate (KHCO3), ammonium acetate, and iridium chloride were

purchased from Merck (Darmstadt, Germany). Ammonium hydrogencarbonate

(NH4HCO3) was purchased from VWR (Fontenay-sous-Bois, France). GSH, zink chloride

(ZnCl2), magnesium chloride (MgCl2), nuclease P1, tris-HCl, phosphate buffered saline

(PBS), and triton X-100 were obtained from Sigma-Aldrich (St. Louis, MO, USA). DNAse I

and alkaline phosphatase were acquired from Roche Applied Science (Basel,

Switzerland). The dinucleosides GpG and ApG were purchased from Metabion (Planegg-

Martinsried, Germany). Absolute ethanol en methanol were obtained from Biosolve

(Valkenswaard, the Netherlands).

Treatment of whole blood with cisplatin and modulators

Two separate experiments were performed to evaluate the effects of STS, GSH, AC, and

gemcitabine on Pt-protein and Pt-DNA binding. In the first experiment, modulators were

added prior to cisplatin incubation to assess whether the modulators could prevent Pt-

protein and Pt-DNA binding. In the second experiment, modulators were added 3 h after

cisplatin incubation to assess whether the modulators were capable of removing Pt from

the proteins and DNA. The administered concentrations of the sulfur-containing

compounds used in the experiments were calculated by dividing the initial dose

generally administered to patients by a blood volume of 5 L [4,15,16]. The cisplatin and

gemcitabine concentrations chosen for the experiments were approximately 30-fold

higher than the maximal concentrations reached after a common intravenous infusion

[17,18]. This was done to achieve Pt-DNA adduct levels that were detectable after HPLC

speciation analysis.

For the first experiment, four samples of 30 mL of heparinised whole blood were

incubated at 37 °C with 3.6 g/L STS, 0.5 g/L GSH, 2.1 g/L AC, or 0.65 g/L gemcitabine. To

Chapter 3.2

156

two 30 mL whole blood samples no modulator was added. After 30 min, cisplatin was

added to all samples up to a final concentration of 0.1 g/L cisplatin (corresponding to

0.065 g/L Pt) and incubation was continued for 6 h.

After 1, 3, and 6 h, aliquots of 10 mL of whole blood were taken from the samples and

transferred to 10 mL heparin-containing tubes (Becton Dickinson Vacutainer Systems,

Plymouth, UK). Following centrifugation for 15 min at 1,000 g and 4 °C, the plasma

fraction was isolated. PUF was obtained by centrifuging the plasma fraction through 30

kDa cut-off ultrafiltrate filters (Centriplus Millipore Corporation, Bedford, MA, USA) for 30

min (1,000 g, 20 °C). Plasma samples were diluted in a 0.01% (g/v) edta diammonium salt

and Triton-X solution in water. PUF samples were diluted using a 1% (v/v) HNO3 solution

in water. After the addition of an internal standard, the Pt contents of the plasma and

pUF samples were analysed using ICP-MS.

The PBMC fraction was isolated from the centrifuged whole blood as described earlier

[19,20]. The sample pretreatment procedure for PBMCs is outlined in Figure 3. Briefly,

contaminating red blood cells were lysed by incubation with 0.83% (w/v) NH4Cl, 0.1%

(w/v) KHCO3, and 1 mM edta disodium salt for 20 min at 4 °C. PBMCs were washed twice

with icecold PBS and resuspended in 9 mL of a buffer containing 10 mM Tris-HCl, 2.3%

(w/v) NaCl, and 2 mM edta disodium salt at pH 7.3.

DNA was isolated from PBMCs as described previously [19,20]. In brief, 0.9 mL 1.1 M

NH4HCO3, 0.45 mL 20% (w/v) SDS, and 150 µL 1% (w/v) proteinase K solution were added

successively, followed by overnight incubation at 42 °C. After the digestion was

complete, 3.3 mL of saturated, 6 M NaCl was added to each tube and the tubes were

shaken vigorously to precipitate proteins. The tubes were centrifuged and the

supernatant containing the DNA was transferred to another tube. Subsequently, the

supernatant was shaken, centrifuged, and transferred to another tube. Following this, 20

mL of absolute ethanol were added to precipitate the DNA. The DNA was washed twice

with 75% ethanol and was subsequently dissolved in 1 mL of water. DNA concentrations

were analysed after dilution in 10 mM Tris-HCl pH 8 by measuring the absorbance at 260

nm using a Biophotometer (Eppendorf, Hamburg, Germany). The purity of the DNA was

checked by determining the absorbance ratio at 260 and 280 nm. Ratios between 1.8

and 2.0 were routinely obtained.

Aliquots of 75 µg of DNA were hydrolysed by incubation in 1% (v/v) HNO3 at 70 °C for 24

h. The resulting solutions were diluted in 1% HNO3 to concentrations within the

calibration range of the ICP-MS. After the addition of internal standard, the total amount

of Pt-DNA adducts was analysed by ICP-MS. The total Pt-adduct concentrations were

expressed in pg Pt per µg DNA. Another 100 µg of DNA were processed to quantitate the

individual adducts by HPLC-ICP-MS as described below.

For the second experiment, six samples of 30 mL of whole blood were incubated with 100

mg/L cisplatin at 37 °C and after 3 h samples were taken for Pt analyses. Subsequently,

Effect of antidotes on platinum binding

157

STS, GSH, AC, and gemcitabine were added in similar concentrations as described above.

To two samples, no modulator was added. The effects of the modulators were evaluated

after 1.5 and 3 h. Plasma, pUF, PBMCs, and DNA were obtained and processed as

described above.

Whole blood

Isolated PBMCs

DNA

Platinum analysis ICP-MS

Speciation digested DNAHPLC

Platinum analysis fractions ICP-MS

Figure 3. Schematic outline of sample pretreatment procedure for PBMCs

Chapter 3.2

158

Enzymatic digestion of DNA

In addition to the determination of the total amount of Pt-DNA adduct, Pt-GG and Pt-AG

adducts were assessed using HPLC-ICP-MS. Furthermore, chromatograms were

investigated to see whether adducts peaks with deviating retention times were formed

in presence of the modulators compared to the samples containing only cisplatin.

Therefore, DNA was digested as described by Pluim et al [19]. Briefly, 100 µg of the DNA

were diluted to 500 µL with water. Subsequently, 150 µL ammonium acetate pH 5 and 6

µL nuclease P1 solution (0.5 U/µL) were added and the solution was incubated for 2 h at

60 °C. Then, 12 µL of a solution containing 1 M Tris-HCl, 10 mM MgCl2, and 1 mM ZnCl2

and 5 µL DNAse I (10 U/µL) were added and incubated for 2 h at 37 °C. Finally, 10 µL of a

alkaline phosphatase solution (1 U/µL) were added and incubated overnight at 37 °C.

The resulting solution which contains unmodified nucleosides and Pt-DNA adducts was

injected directly into the HPLC system.

Separation of Pt-DNA adducts

Analytical separation of the two major cisplatin-DNA adducts Pt-GG and Pt-AG was

carried out with an HPLC system consisting of an 1100 Series liquid chromatograph

binary pump and degasser (Agilent technologies, Palo Alto, CA, USA), a Spectra Series

AS3000 autosampler with column oven equipped with a 20 µL injection loop (Thermo

Separation Products (TSP), Fremont, CA, USA), and a photo-diode-array (PDA) detector

Model Waters 996 (Waters Chromatography BV, Etten-Leur, The Netherlands).

Separation was achieved using a Polaris 5 C18-A chromsep column (150 x 2 mm ID,

particle size 5 µm; Varian BV, Middelburg, The Netherlands). The temperature of the

column was kept at 35 °C. Chromatograms were processed using Chromeleon software

(Dionex Corporation, Sunnyvale, CA, USA). The mobile phase consisted of 5 mM

ammonium acetate pH 4 in 2.5% methanol (buffer A) and in 25% methanol (buffer B).

Following injection of the digested DNA (approximately 1.5 µg in 10 µL), the analytes

were eluted off the column by a gradient increasing from 0 to 100% buffer B (Table 1)

with a flow of 0.2 mL/min. Pt-DNA adduct concentrations were below the detection limit

of the UV detector. Therefore, eluting fractions were collected at intervals of 1 min and

the Pt content was analysed using ICP-MS after a tenfold dilution in 1% (v/v) HNO3.

The identity of the individual Pt-DNA adducts was confirmed by chromatography of the

reaction products of GpG and ApG with cisplatin. Hence, respectively 596 mg/L and 580

mg/L of the dinucleosides GpG and ApG were incubated with 300 mg/L cisplatin (195

mg/L Pt). Adduct concentrations in the Pt-GG and Pt-AG incubation solutions were

higher than in the digested DNA solutions obtained from whole blood. Therefore,

Effect of antidotes on platinum binding

159

adduct peaks could be identified based on absorption spectrum, retention time, and Pt

content of the peaks.

Table 1. HPLC gradient

Time (min) Buffer A (%) Buffer B (%)

0-20 100 0

20-40 100 → 0 0 → 100

40-41 0 → 100 100 → 0

41-50 100 0

Determination of Pt concentrations by ICP-MS

Pt analyses were performed on an ICP-quadrupole-MS (Varian 810-MS) equipped with a

90° reflecting ion mirror (Varian, Mulgrave, Victoria, Australia). The sample introduction

system consisted of a Micromist glass low-flow nebuliser (sample uptake 0.4 mL/min), a

peltier-cooled (4 °C) double pass glass spray chamber and a quartz torch. Hoek Loos

(Schiedam, The Netherlands) provided argon gas (4.6) with a 99.996% purity. Data were

acquired and processed using the ICP-MS Expert Software version 1.1 b49 (Varian). The

Pt isotope used for calculation of Pt concentrations was 194Pt. Internal standardisation

was performed on each replicate using iridium (191Ir).

Results

Effect of STS, GSH, AC, and gemcitabine on Pt-protein binding

The effects of the modulators on the recovery of Pt in pUF when added prior to cisplatin

addition (first experiment) are illustrated in Figure 4a. Between 1 and 6 h after start of

incubation, Pt continued to bind resulting in an ongoing reduction of the ultrafiltrable Pt

fraction. After 6 h, the ultrafiltrable fraction in the samples that were solely incubated

with cisplatin (controls 1 and 2 in Figure 4) was reduced to 16%. Gemcitabine did not

affect this Pt binding. GSH and AC both appeared to partly prevent Pt-protein binding.

The administered GSH and AC concentrations resulted in an increase of ultrafiltrable Pt

of respectively 25 and 57% at 6 h. Pt-protein binding was almost completely prevented

by STS. When the modulators were added 3 h after the samples were incubated with

cisplatin (second experiment), initially, 31% of Pt was recovered in the pUF (Figure 4b).

After 6 h of incubation, the samples incubated solely with cisplatin or with cisplatin and

gemcitabine, showed a Pt recovery in pUF of 12%. Thus, again, gemcitabine did not

Chapter 3.2

160

affect Pt-protein binding. GSH appeared to limit Pt-protein binding with 13% at 6 h,

resulting in an ultrafiltrable fraction of 25%. In the AC and STS containing samples, after

6 h of incubation, the ultrafiltrable Pt fraction was raised from the initial 31% to 39 and

46%, respectively. Interestingly, AC and STS seemed to be capable of releasing Pt from

the proteins, resulting in a larger ultrafiltrable Pt fraction.

Effect of STS, GSH, AC, and gemcitabine on Pt-DNA binding

The total amounts of Pt-DNA adducts in PBMCs were analysed by ICP-MS without prior

separation. The effects of the modulators on the formation of the total amount of Pt-

DNA adducts when they were added prior to cisplatin addition (first experiment) are

shown in Figure 5a. After 1 h of incubation, PBMCs contained 24 pg Pt/µg DNA in the

samples solely incubated with cisplatin. At this time point, gemcitabine and GSH did not

affect Pt-DNA binding. AC and STS, however, inhibited Pt-DNA binding with respectively

45% and 80% at 1 h. After 3 and 6 h of incubation, samples containing only cisplatin

showed a Pt-DNA binding of 65 and 93 pg Pt/µg DNA, respectively. Gemcitabine seemed

to slightly inhibit Pt-DNA binding (7-16%). For GSH, no inhibitory effect was observed

after 3 h of incubation. After 6 h, however, Pt-DNA binding appeared to be 18% lower

compared to the samples containing only cisplatin. AC, obviously decreased the extent

and rate of Pt-DNA adduct formation. After 6 h, Pt-DNA adduct levels were 80% lower

than levels in the samples with only cisplatin. In the STS incubated samples, no increase

of the Pt-DNA adduct levels with time was observed. When the modulators were added

3 h after the samples were incubated with cisplatin (second experiment), initially, on

average 104 pg Pt was bound to 1 µg DNA (Figure 5b). After 6 h of incubation, samples

with solely cisplatin or cisplatin and gemcitabine contained on average 137 pg Pt/µg

DNA. Pt-DNA binding was not affected by gemcitabine. Pt-DNA binding was, similar to

the first experiment, inhibited by GSH and after 6 h, Pt-DNA levels were 17% lower than

levels in the samples incubated with only cisplatin. AC and STS appeared to completely

prevent further Pt-DNA binding. After 6 h, Pt-DNA adduct levels even seemed to be 8%

lower than before the addition of the modulating compounds.

Effect of antidotes on platinum binding

161

0

20

40

60

80

100

120

1 2 3 4 5 6 7

Time (h)

Plat

inum

reco

vere

d in

pU

F (%

)

Control 1Control 2GemcitabineAcetylcysteineSodium thiosulfateGlutathione

Figure 4a. Percentage of ultrafiltrable Pt in plasma versus time after incubation of whole blood with

modulators for ½ h, followed by cisplatin incubation (first experiment)

0

10

20

30

40

50

60

1 2 3 4 5 6 7

Time (h)

Plat

inum

reco

vere

d in

pU

F (%

)

Control 1Control 2GemcitabineAcetylcysteineSodium thiosulfateGlutathione

Figure 4b. Percentage of ultrafiltrable Pt in plasma versus time after incubation of whole blood with

cisplatin for 3 h followed by the addition of modulators (second experiment)

Chapter 3.2

162

0

50

100

1 2 3 4 5 6 7

Time (h)

pg P

t/ug

DN

A

Control 1Control 2GemcitabineAcetylcysteineSodium thiosulfateGlutathione

Figure 5a. Time dependent formation of Pt-DNA adducts in PBMCs after incubation of whole blood

with modulators for ½ h, followed by cisplatin incubation (first experiment)

50

75

100

125

150

175

1 2 3 4 5 6 7

Time (h)

pg P

t/ug

DN

A

Control 1Control 2GemcitabineAcetylcysteineSodium thiosulfateGlutathione

Figure 5b. Time dependent formation of Pt-DNA adducts in PBMCs after incubation of whole blood

with cisplatin for 3 h followed by the addition of modulators (second experiment)

Effect of antidotes on platinum binding

163

In addition to the analyses of the total amount of Pt-DNA adducts, analytical separation

of the two major cisplatin-DNA adducts Pt-GG and Pt-AG was applied to obtain

information regarding the type of adducts formed in the samples and the ratio of Pt-GG

and Pt-AG. Pt-GG and Pt-AG formation was followed through time. Of two samples,

HPLC-ICP-MS chromatograms are depicted in Figure 6. Figure 6a shows the reaction of

solely cisplatin with DNA after 1, 3, and 6 h of incubation and Figure 6b shows the

reaction of cisplatin with DNA after the addition of AC. The peak at 15 min represents Pt-

GG, whereas the small peak at 27 min corresponds to Pt-AG. The increase of the levels of

Pt-GG and Pt-AG analysed by HPLC-ICP-MS with time were in agreement with the

increase observed for the total amount of Pt-DNA adducts as analysed by ICP-MS. Due to

the low injection volume (10 µL) and the dilution prior to ICP-MS analyses, samples,

however, were diluted 250-fold compared to ICP-MS alone. Therefore, unfortunately,

sensitivity decreased and for the samples containing low concentrations of adducts no

adequate determination of Pt-GG and Pt-AG could be performed. For that reason, the

separation method, was used only to gain information on the retention time and the

ratios of the adduct peaks formed with or without the presence of modulators. None of

the chromatograms revealed a different pattern of peaks and ratios remained constant,

which suggests that the type of adducts formed were similar for all the samples.

The identities of the Pt-GG and Pt-AG peaks were confirmed by chromatography of the

in vitro reaction products of cisplatin with GpG and ApG. UV and ICP-MS chromatograms

are shown in Figure 7a and b. The large figures show the chromatograms of the

incubation mixture, whereas the small figures represent the chromatograms of the

dinucleosides GpG and ApG. The major peaks visible in the chromatograms correspond

to Pt-GG (15.5 min) and Pt-AG (27 min). The large peak at one min is caused by cisplatin,

whereas the other peaks, most probably, are caused by other adducts formed in the

incubation mixtures such as GG-Pt-GG and AG-Pt-AG. The identities of the Pt-GG and Pt-

AG peak were confirmed by the absorption maximum of the peaks which were shifted to

lower energy (higher λmax) compared to the unreacted dinucleosides [21]. Furthermore,

the retention times of the adduct peaks were shorter than the unreacted dinucleosides

indicating that the hydrophobicity of the dinucleosides was reduced by cisplatin.

Chapter 3.2

164

0

50

100

150

200

250

300

0 5 10 15 20 25 30 35 40

Retention (min)

Plat

inum

sig

nal I

CP-M

S

Figure 6a. Time dependent formation of Pt-DNA adducts in PBMCs after incubation of whole blood

with cisplatin

0

50

100

150

200

250

300

0 5 10 15 20 25 30 35 40

Retention (min)

Plat

inum

sig

nal I

CP-M

S

Figure 6b. Time dependent formation of Pt-DNA adducts in PBMCs after incubation of whole blood

with cisplatin preceded by the incubation with AC

Pt-GG

1 h 3 h

6 h

Pt-AG

1 h 3 h

6 h

Effect of antidotes on platinum binding

165

0 10 20 30 40 50

Retention (min)

UV absorption 256nmPt signal ICP-MS

Figure 7a. HPLC-UV and ICP-MS chromatogram of GpG incubated with cisplatin

0 10 20 30 40 50

Retention (min)

UV absorption 256nmPt signal ICP-MS

Figure 7b. HPLC-UV and ICP-MS chromatogram of ApG incubated with cisplatin

20 30 40

20 30 40

Pt-GG

λmax=259.1

GG λmax=251.9

Pt-AG

λmax=261.7

AG λmax=253.7

Chapter 3.2

166

Discussion

Cisplatin has developed into a frequently used chemotherapeutic agent. Its use,

however, is hampered by severe toxicities, which can seriously affect the quality of life.

Treatment strategies aimed to reduce the side effects without affecting the antitumour

activity are frequently tested.

The aim of the current study was to investigate the effect of the compounds STS, GSH,

AC, and gemcitabine on the Pt-protein and Pt-DNA binding of cisplatin in whole blood.

This was done to gain more insight into the Pt binding in whole blood and the effect of

modulators on this process. Thereby, information is gained on the effect of the

modulators on the activity of cisplatin. Furthermore, the potential use of sulfur-

containing compounds for the reduction or prevention of side effects and, potentially,

for assisting the recovery of these side effects in vivo.

Initially, the effect of STS, GSH, AC, and gemcitabine on Pt-protein binding was

evaluated. A reduced Pt-protein binding suggests a lower reactivity of Pt present in the

blood stream and this could lead to reduced cytotoxicity. After 6 h of incubation of

cisplatin 85% was bound to blood constituents which is in accordance with the protein

binding in the human body [1]. Gemcitabine did not show an effect on Pt-protein

binding, which was expected, because the nitrogen groups in gemcitabine do not have

a stronger nucleophilic character than the sulfur-groups in proteins [22]. Therefore, Pt-

protein binding will be preferred over Pt-gemcitabine binding. GSH and AC, on the other

hand appeared to be capable of inhibiting Pt-protein binding to a large extent.

Interestingly, Pt-protein binding was found to be almost completely inhibited by STS. A

large inhibition (46%) of the ex vivo Pt-protein binding by STS was also described by

Elferink et al. [23]. The superior effect of STS on Pt-protein binding compared to GSH and

AC may be explained by the electrostatic interaction between the partially positively

charged hydrated product of cisplatin and the doubly negatively charged thiosulfate.

STS has a more nucleophilic character than GSH and AC [24]. Another issue which might

explain the differences in effects between the compounds, could be the deviating molar

concentrations used in the current experiments. The STS, GSH, and AC concentrations

chosen for the current research were in accordance with the doses of the compounds

that are generally administered to patients. Molar ratios were 23 mM STS : 1.6 mM GSH :

13 mM AC.

In addition to the inhibition of Pt-protein binding, AC and STS appeared to be capable to

remove Pt from proteins. Because the thermodynamic strength of a typical coordination

bond such as the Pt-ligand bond is much weaker than a covalent bond, the ligands of Pt

compounds can usually be exchanged easily [25]. The observation that Pt in the

ultrafiltrable fraction was increased after the addition of AC and STS, suggests that these

compounds might have shifted the reaction equilibrium of the Pt-protein coordination

complex, resulting in a decreased level of Pt-protein complexes. It might be that AC and

Effect of antidotes on platinum binding

167

STS are stronger Lewis bases than most of the sulfur groups present in the proteins, or

that interactions of AC and STS with Pt are impeded less by sterical effects and are thus

thermodynamically more stable than the interactions of proteins with Pt. The possibility

to release Pt from proteins might result in an increased rate of elimination of Pt from the

body and in that way in a reduced toxicity [15]. The ability of STS to remove Pt from

proteins shown in the current study, however, is in contrast to an earlier ex vivo

investigation by Elferink et al., who mentioned that STS was not able to reverse Pt-

protein binding to a major extent [23]. Hence, before drawing definite conclusions from

the current results, the effect of sulfur-containing compounds on protein structures and

proteolysis should be investigated. If sulfur-containing compounds could affect the

tertiary protein structure, or when they could induce proteolysis, an increased amount of

bound Pt would be recovered in the ultrafiltrate.

The ultrafiltrable Pt fraction observed after incubation with sulfur-containing

compounds, was expected to contain smaller amounts of reactive Pt compared to the

samples without sulfur-containing compounds. A previous investigation showed a

decreased reactivity of the ultrafiltrable Pt in patients who received co-administration of

STS, indicating that the Pt-STS complex was not reactive [26]. The Pt reactivity in this

study was measured as the ability of ultrafiltrable Pt to bind to diethyldithiocarbamate.

Other investigations reported that the ultrafiltrable fraction after STS co-administration

contained significantly less unchanged cisplatin than without STS treatment [12,27].

Lower concentrations of reactive Pt can of course lead to lower toxicity.

In order to assess the reactivity of Pt present in the samples of the current research, we

investigated the Pt-DNA binding reactivity in PBMCs. Pretreatment of the samples with

gemcitabine appeared to slightly inhibit Pt-DNA binding. Currently, no explanation for

this interaction is known. No effect of gemcitabine on the protein level was observed

and the interaction with Pt is expected to occur intracellulary. The highly significant

effect as was observed by Crul et al. [13], however, was absent in the present

investigation. A reason for this inconsistency could be the difference in experimental

circumstances. Crul et al. evaluated the effect of gemcitabine on Pt-DNA binding in vivo,

whereas the current experiments were performed ex vivo. The Pt-DNA binding observed

for GSH and AC were respectively 18 and 80% lower than observed for cisplatin alone.

STS appeared to completely prevent Pt-DNA adduct binding. The inhibition of Pt-DNA

by the modulators is a relevant parameter for the reactivity of Pt and thus for the

potential to prevent toxicity. The current data suggest that all of the tested sulfur-

containing compounds may reduce toxicity and that STS is the most potent compound.

Concurrent administration of cisplatin with sulfur-containing compounds might affect

side effects, as well as efficacy. Therefore, dosing regimens and administration routes

should be selected carefully, as illustrated by the intra-arterially cisplatin treatment with

intravenously administered STS for the treatment of head and neck cancer [4].

Chapter 3.2

168

In addition to the ability of the modulators to prevent Pt-DNA binding, it was also

investigated whether the modulators could release Pt from the DNA. For STS, the

distribution of which is limited to the extracellular space [23], it is unlikely that it can

directly interact with the Pt-DNA adducts and thereby release Pt. AC, however, is able to

enter cells [28] and, thus, might be able to release Pt from the DNA. The current

experiments showed that, although, AC and STS completely inhibited further Pt-DNA

binding when they were administered 3 h after the start of cisplatin incubation, they

could hardly release Pt from the DNA. GSH also limited Pt-DNA binding, but did not

prevent it. These observations are in line with observations in previous investigations

which suggested that platinum species might migrate from S to N donor ligands [29,30],

implying that Pt-DNA bonds at GG sites are thermodynamically more stable than Pt-S

bonds.

These data imply that, to be protective, the time of administration of the compound is

relevant, because none of the compounds could obviously release Pt from the DNA. The

cytotoxic protective effect should be initiated before the binding of cisplatin to the DNA.

As mentioned before for STS, protection against nephrotoxicity was accomplished only

when STS was given within one hour before and 30 min after cisplatin administration

[31]. Furthermore, the observation that no Pt was released from the DNA, implies that

the tested compounds are probably not capable of reducing persistent toxicity.

In addition to the total amount of Pt-DNA adducts, the different adducts formed after

separation using HPLC were studied. Incubation with STS, GSH, AC, and gemcitabine did

not result in different adducts. Furthermore, the ratio of Pt-GG and Pt-AG remained

constant under all conditions. These results suggest that the DNA binding mode of

cisplatin is not modified by STS, GSH, AC, and gemcitabine.

In conclusion, the current study showed that the reactive Pt levels in whole blood are

reduced by STS, GSH, and AC after ex vivo incubation of whole blood with cisplatin. This

reduction was demonstrated by a reduced Pt-protein and Pt-DNA binding in the

presence of the sulfur-containing compounds. Consequently, STS, GSH, and AC could

prevent cisplatin induced side effects. It is, however, not expected that these

compounds might reduce persistent toxicities, because they were not able to release Pt

from the DNA to a large extent. Hence, a large activity of the compounds is only

expected when they are administered during or shortly after cisplatin treatment, which

may have a large impact on cytotoxicity. The minor effect of gemcitabine on Pt-DNA

binding needs to be evaluated further in future experiments. The ex vivo effects

observed in the current study are expected to be larger than the in vivo effects, because

the exposure time to the modulators will be shorter in vivo, due to elimination of the

compounds from the blood. This study, however, provides further insight into the

potential effects and use of STS, GSH, AC, and gemcitabine in patients treated with

cisplatin.

Effect of antidotes on platinum binding

169

References 1. Mandal R, Teixeira C, Li XF. Studies of cisplatin and hemoglobin interactions using nanospray mass

spectrometry and liquid chromatography with inductively-coupled plasma mass spectrometry. Analyst 2003; 128: 629-34.

2. Calvert H, Judson I, van der Vijgh WJ. Platinum complexes in cancer medicine: pharmacokinetics and pharmacodynamics in relation to toxicity and therapeutic activity. Cancer Surv 1993; 17: 189-217.

3. Graham MA, Lockwood GF, Greenslade D, Brienza S, Bayssas M, Gamelin E. Clinical pharmacokinetics of oxaliplatin: a critical review. Clin Cancer Res 2000; 6: 1205-18.

4. Hoebers FJ, Pluim D, Verheij M, Balm AJ, Bartelink H, Schellens JH, Begg AC. Prediction of treatment outcome by cisplatin-DNA adduct formation in patients with stage III/IV head and neck squamous cell carcinoma, treated by concurrent cisplatin-radiation (RADPLAT). Int J Cancer 2006; 119: 750-6.

5. Howell SB, Pfeifle CE, Wung WE, Olshen RA. Intraperitoneal cis-diamminedichloroplatinum with systemic thiosulfate protection. Cancer Res 1983; 43: 1426-31.

6. Markman M, Howell SB, Lucas WE, Pfeifle CE, Green MR. Combination intraperitoneal chemotherapy with cisplatin, cytarabine, and doxorubicin for refractory ovarian carcinoma and other malignancies principally confined to the peritoneal cavity. J Clin Oncol 1984; 2: 1321-6.

7. van Rijswijk RE, Hoekman K, Burger CW, Verheijen RH, Vermorken JB. Experience with intraperitoneal cisplatin and etoposide and i.v. sodium thiosulphate protection in ovarian cancer patients with either pathologically complete response or minimal residual disease. Ann Oncol 1997; 8: 1235-41.

8. Madasu R, Ruckenstein MJ, Leake F, Steere E, Robbins KT. Ototoxic effects of supradose cisplatin with sodium thiosulfate neutralization in patients with head and neck cancer. Arch Otolaryngol Head Neck Surg 1997; 123: 978-81.

9. Di Re F, Bohm S, Oriana S, Spatti GB, Zunino F. Efficacy and safety of high-dose cisplatin and cyclophosphamide with glutathione protection in the treatment of bulky advanced epithelial ovarian cancer. Cancer Chemother Pharmacol 1990; 25: 355-60.

10. Dickey DT, Wu YJ, Muldoon LL, Neuwelt EA. Protection against cisplatin-induced toxicities by N-acetylcysteine and sodium thiosulfate as assessed at the molecular, cellular, and in vivo levels. J Pharmacol Exp Ther 2005; 314: 1052-8.

11. Filipski J, Kohn KW, Prather R, Bonner WM. Thiourea reverses cross-links and restores biological activity in DNA treated with dichlorodiaminoplatinum (II). Science 1979; 204: 181-3.

12. Brouwers EEM, Huitema ADR, Beijnen JH, Schellens JHM. Long-term platinum retention after treatment with cisplatin and oxaliplatin. Submitted 2007.

13. Crul M, Schoemaker NE, Pluim D, Maliepaard M, Underberg RW, Schot M, Sparidans RW, Baas P, Beijnen JH, van Zandwijk N, Schellens JH. Randomized phase I clinical and pharmacologic study of weekly versus twice-weekly dose-intensive cisplatin and gemcitabine in patients with advanced non-small cell lung cancer. Clin Cancer Res 2003; 9: 3526-33.

14. Brouwers EEM, Tibben MM, Rosing H, Hillebrand MJX, Joerger M, Schellens JHM, Beijnen JH. Sensitive inductively coupled plasma mass spectrometry assay for the determination of platinum originating from cisplatin, carboplatin, and oxaliplatin in human plasma ultrafiltrate. J Mass Spectrom 2006; 41: 1186-94.

15. Leone R, Fracasso ME, Soresi E, Cimino G, Tedeschi M, Castoldi D, Monzani V, Colombi L, Usari T, Bernareggi A. Influence of glutathione administration on the disposition of free and total platinum in patients after administration of cisplatin. Cancer Chemother Pharmacol 1992; 29: 385-90.

16. Prescott LF, Donovan JW, Jarvie DR, Proudfoot AT. The disposition and kinetics of intravenous N-acetylcysteine in patients with paracetamol overdosage. Eur J Clin Pharmacol 1989; 37: 501-6.

17. Siegel-Lakhai WS, Crul M, Zhang S, Sparidans RW, Pluim D, Howes A, Solanki B, Beijnen JH, Schellens JH. Phase I and pharmacological study of the farnesyltransferase inhibitor tipifarnib (Zarnestra, R115777) in combination with gemcitabine and cisplatin in patients with advanced solid tumours. Br J Cancer 2005; 93: 1222-9.

18. Rademaker-Lakhai JM, Crul M, Pluim D, Sparidans RW, Baas P, Beijnen JH, van Zandwijk N, Schellens JH. Phase I clinical and pharmacologic study of a 2-weekly administration of cisplatin and gemcitabine in patients with advanced non-small cell lung cancer. Anticancer Drugs 2005; 16: 1029-36.

19. Pluim D, Maliepaard M, van Waardenburg RC, Beijnen JH, Schellens JHM. 32P-postlabeling assay for the quantification of the major platinum-DNA adducts. Anal Biochem 1999; 275: 30-8.

Chapter 3.2

170

20. Ma J, Verweij J, Planting AS, Boer-Dennert M, van Ingen HE, van der Burg ME, Stoter G, Schellens JH. Current sample handling methods for measurement of platinum-DNA adducts in leucocytes in man lead to discrepant results in DNA adduct levels and DNA repair. Br J Cancer 1995; 71: 512-7.

21. Hann S, Zenker A, Galanski M, Bereuter TL, Stingeder G, Keppler BK. HPIC-UV-ICP-SFMS study of the interaction of cisplatin with guanosine monophosphate. Fresenius J Anal Chem 2001; 370: 581-6.

22. Hay RW, Porter DS. The reaction of sulphur and nitrogen nucleophiles with [Pt(dien)Cl]+. Transition Met Chem 1999; 24: 186-8.

23. Elferink F, van der Vijgh WJ, Klein I, Pinedo HM. Interaction of cisplatin and carboplatin with sodium thiosulfate: reaction rates and protein binding. Clin Chem 1986; 32: 641-5.

24. Videhult P, Laurell G, Wallin I, Ehrsson H. Kinetics of Cisplatin and its monohydrated complex with sulfur-containing compounds designed for local otoprotective administration. Exp Biol Med (Maywood ) 2006; 231: 1638-45.

25. Reedijk J. New clues for platinum antitumor chemistry: kinetically controlled metal binding to DNA. Proc Natl Acad Sci U S A 2003; 100: 3611-6.

26. Goel R, Cleary SM, Horton C, Kirmani S, Abramson I, Kelly C, Howell SB. Effect of sodium thiosulfate on the pharmacokinetics and toxicity of cisplatin. J Natl Cancer Inst 1989; 81: 1552-60.

27. Nagai N, Hotta K, Yamamura H, Ogata H. Effects of sodium thiosulfate on the pharmacokinetics of unchanged cisplatin and on the distribution of platinum species in rat kidney: protective mechanism against cisplatin nephrotoxicity. Cancer Chemother Pharmacol 1995; 36: 404-10.

28. McLellan LI, Lewis AD, Hall DJ, Ansell JD, Wolf CR. Uptake and distribution of N-acetylcysteine in mice: tissue-specific effects on glutathione concentrations. Carcinogenesis 1995; 16: 2099-106.

29. Reedijk J. Why does Cisplatin reach Guanine-N7 with competing S-donor ligands available in the cell? Chem Rev 1999; 99: 2499-510.

30. Chen Y, Guo ZD, Murdoch PD, Zang EL, Sadler PJ. Interconversion between S- and N-bound L-methionine adducts of Pt(dien)2+ (dien=diethylenetriamine) via dien ring-opened intermediates. J Chem Soc , Dalton Trans 1998; 1503-8.

31. Gandara DR, Wiebe VJ, Perez EA, Makuch RW, DeGregorio MW. Cisplatin rescue therapy: experience with sodium thiosulfate, WR2721, and diethyldithiocarbamate. Crit Rev Oncol Hematol 1990; 10: 353-65.

Chapter 4 Persistent effects of

platinum agents

Chapter 4.1

Long-term platinum retention after treatment with cisplatin and oxaliplatin

Elke E.M. Brouwers Alwin D.R. Huitema

Jos H. Beijnen Jan H.M. Schellens

Submitted for publication

Chapter 4.1

176

Abstract

The purpose of this study was to evaluate long-term platinum (Pt) retention in patients

previously treated with oxaliplatin and cisplatin. Forty-five patients, treated 8-75 months

before participating in this study, were included. Pt levels in plasma and plasma

ultrafiltrate (pUF) were determined using a highly sensitive inductively coupled plasma

mass spectrometry method. In addition, the reactivity of Pt species recovered in pUF was

evaluated. Relationships between long-term Pt retention and possible determinants

were evaluated using non linear mixed effects modeling.

Pt plasma concentrations ranged between 142-2.99x103 ng/L. On average 15% and 24%

of plasma Pt was recovered in pUF for cisplatin and oxaliplatin, respectively. No Pt-DNA

adducts in peripheral blood mononuclear cells (PBMCs) could be detected. Ex vivo

incubation of DNA with pUF of patients revealed that up to 10% of the reactivity of Pt

species in patients pUF was retained. Protein binding was shown to proceed during

sample storage. Sodium thiosulfate (STS) appeared to be capable of releasing Pt from

the plasma proteins, suggesting that protein binding was not irreversible. Pt levels were

related to follow-up time, age, cumulative dose, co-administration of STS with intra-

arterial cisplatin administration, and glomerular filtration rates before start of

chemotherapy. No relationship between glutathione S-transferase genotypes and Pt

levels was observed.

In conclusion, our data suggest that plasma Pt levels are related to follow-up time, age,

cumulative dose, GFR at time of treatment, and STS use. Pt levels in plasma, most

probably, represent Pt eliminated from regenerating tissue. Although no Pt-DNA

adducts could be detected in PBMCs, it was shown that Pt species in pUF were still

present in a reactive form. The clinical consequences of the observed reactivity remain

to be established.

Long-term platinum retention

177

Introduction

Since its discovery as an effective anticancer agent in the 1960s [1], cisplatin is used

extensively in oncology. The use of platinum (Pt) agents has had an enormous impact on

the prognosis of several cancer types. After the introduction of cisplatin, mortality of e.g.

testicular cancer reduced significantly, leading to an increased number of survivors. The

success of cisplatin has led to the development of other Pt-based compounds, such as

oxaliplatin, which are effective in cancer types resistant to cisplatin. Oxaliplatin has

found a widespread use in the treatment of cisplatin resistant colorectal cancer [2].

The improved life expectancy of cancer patients treated with Pt-based compounds, has

led to an increased interest in the long-term side effects of these drugs. Among these

side effects are peripheral neuropathy, nephrotoxicity, ototoxicity, and secondary

malignancies [3-6]. The presence of long-term side effects has led to the investigation of

long-term pharmacokinetics, distribution, and elimination of Pt-based drugs. Studies

have shown that with a standard cisplatin-containing chemotherapy, plasma and tissue

Pt levels are still remarkably elevated years after chemotherapy [7-12]. For oxaliplatin, no

data on long-term pharmacokinetics are available yet. In addition, no studies have been

performed to investigate the potential reactivity of retained Pt species years after

treatment.

In the current study the long-term Pt retention in plasma and plasma ultrafiltrate (pUF)

of patients treated with cisplatin or oxaliplatin up to 6 years before participating in this

study was investigated. The in vivo reactivity of circulating Pt was studied by testing the

DNA- and protein binding activity of ultrafilterable Pt and the ability of sodium

thiosulfate (STS) to release Pt from the plasma proteins. For quantification of Pt levels in

plasma, pUF, and for quantification of the level of Pt-DNA adducts, we used inductively

coupled plasma mass spectrometry (ICP-MS). The low detection limits of this technique

allows to even assess natural Pt background levels in biological matrices [13-16] and

thereby enabled us to study the long-term retention and remaining reactivity of Pt

agents. Finally, potential relationships between Pt exposure and follow-up time, age,

cumulative dose, route of administration, renal function, glutathione S-transferase (GST)

genotypes, and co-administration of STS with intra-arterial cisplatin were investigated.

Methods

Participants

For cisplatin, patients were selected at random from all patients who started treatment

between 2000 and 2004, received cumulative cisplatin doses of ≥ 300 mg/m2, and were

available for follow-up. For this pilot study, 20 patients of the 400 eligible patients were

Chapter 4.1

178

included. To select the patients, random selections were performed on the 400 eligible

patients until 20 patients agreed to participate in the study. SPSS (SPSSinc, version 11.0,

Chicago, IL, USA) was used for random sample selection. For oxaliplatin, all available

patients who started treatment between 2000 and 2005 and received cumulative

oxaliplatin doses of ≥ 600 mg/m2 were approached for participation in the current study.

This led to an inclusion of 25 patients. The Medical Ethics Committee of the hospital

approved the study protocol and all patients gave their written informed consent.

Additionally, we included 20 cancer patients who were not treated with cisplatin and 20

normal volunteers, both as a control for Pt background levels in plasma.

Blood sampling

Whole blood samples for Pt analysis were collected in 10 mL edta-containing tubes

(Becton Dickinson Vacutainer Systems, Plymouth, UK). Plasma was obtained by

centrifuging the whole blood samples for 15 min (1,000 g, 4 °C). Plasma ultrafiltrate was

obtained by centrifuging the plasma fraction through 3 and 30 kDa cut-off ultrafiltrate

filter (Centriplus Millipore Corporation, Bedford, MA, USA) for 30 min (1,000 g, 20 °C). The

fraction containing peripheral blood mononuclear cells (PBMCs) was isolated from the

whole blood sample using the method described by Pluim et al [17].

Additionally, from each patient, 5 ml blood samples were obtained for genetic analysis.

Lymphocyte DNA was isolated according to the method of Boom [18]. All samples were

stored at –20 °C until analysis.

Determination of Pt levels

Pt analyses were performed using an ICP-quadrupole-MS (Varian 810-MS, Varian,

Mulgrave, Victoria, Australia) and a validated method described previously [19]. The Pt

isotope used for calculation of the validation parameters was 194Pt. 191Ir (Merck,

Darmstadt, Germany) was used for internal standardisation. From patient samples, Pt

levels were assessed in plasma, pUF, and bound to DNA in PBMCs. Plasma samples were

diluted in a 0.01% (g/v) edta diammonium salt (Merck) and 0.01% (g/v) Triton-X solution

(Sigma-Aldrich, St. Louis, MO, USA) in water. PUF samples were diluted using a 1% (v/v)

nitric acid (HNO3) solution in water (Mallinckrodt Baker, Philipsburg, NJ, USA). DNA was

isolated from PBMCs using a method described by Pluim et al [17]. Following isolation,

the DNA content was determined after dilution in 10 mM Tris-HCl pH 8 by measuring the

absorbance at 260 nm using a Biophotometer (Eppendorf, Hamburg, Germany). The

purity of the DNA was checked by measurement of the absorbance ratio at 260 and 280

nm. Ratios between 1.8 and 2.0 were routinely obtained. Before ICP-MS analysis, the

DNA was hydrolysed for 24 h in 1% (v/v) HNO3 solution at 70 °C. The limit of

Long-term platinum retention

179

quantification (LLOQ) of the method was 20 ng/L for plasma, 7.5 ng/L for pUF, and the

absolute sensitivity for Pt bound to DNA was 0.75 pg Pt (7.5 fg Pt per µg DNA when

using 100 µg DNA). The LLOQs were determined on the basis of five times the noise in

blank matrix solutions. The noise was assessed by analysing different batches of control

plasma and pUF. The accuracies at the LLOQ levels were between 80-120%, whereas the

precisions were lower than 20% [19].

All dilutions were prepared using plastic pipettes (Falcon, Becton Dickinson Labware,

Franklin Lakes, NJ, USA) and polypropylene tubes of 10 mL (Plastiques-Gosselin,

Hazebrouck Cedex, France) and 30 mL (Sarstedt AG&Co, Nümbrecht, Germany), which

were all evaluated thoroughly for Pt contamination prior to method development [19].

Ex vivo assessment of DNA binding activity of Pt in pUF

PUF samples (500 µL) of two cisplatin and two oxaliplatin treated patients, which

contained relatively high Pt concentrations, were incubated with 500 µg of calf thymus

DNA (Sigma-Aldrich). Additionally, pUF from normal volunteers was incubated with 500

µg of calf thymus DNA and cisplatin or oxaliplatin with Pt concentrations equivalent to

the investigated patients samples. The latter was done to assess the binding capacity of

the parent compounds at concentrations in the same range as the patients samples.

After a five-day incubation at 37 °C to achieve a maximal Pt-DNA binding, the DNA was

precipitated using 100% ethanol (Biosolve, Valkenswaard, the Netherlands) followed by

two wash steps with 75% ethanol to remove unbound Pt. The DNA was then dissolved in

water. After determination of the DNA concentrations, the DNA samples were

hydrolysed at 70 °C for 24 h in 1% HNO3 and Pt-DNA binding was analysed.

Ex vivo assessment of protein binding capacity of Pt in pUF

To assess the protein binding capacity of Pt present in pUF, pUF (30 kDa) was prepared

for four cisplatin and oxaliplatin plasma samples after storage at -30 °C for 144-278 days.

Samples contained Pt concentrations in the low, mid, and high region. Pt concentrations

analysed in pUF prepared after storage were compared to Pt concentrations analysed in

pUF which was prepared immediately after blood sampling. Additionally, plasma

samples were reanalysed to assess whether Pt concentrations in plasma were reduced

during storage due to adsorption to the tubes.

Ex vivo activity of STS

Four plasma samples for both cisplatin and oxaliplatin were used to investigate the

ability of STS to remove Pt from proteins. Therefore, 750 µL of plasma were incubated

Chapter 4.1

180

with 25 µL of a 250 g/L STS solution (Hospital Pharmacy of Haarlem, Haarlem, the

Netherlands). As control, 750 µL of plasma were incubated with 25 µL of water. After

incubation, pUF was prepared (30 kDa) and Pt concentrations of the STS incubated

samples and control samples were compared. Protein concentrations in the ultrafiltrates

were analysed using a 2-D Quant Kit (GE healthcare Bio-Sciences AB, Uppsala, Sweden).

In contrast to other methods to assess protein concentrations, this method uses protein

precipitation to avoid interference of reducing agents such as STS.

Genotyping

Polymorphisms in the genes encoding the enzymes GSTM1, GSTT1, and GSTP1 were

determined. In GSTT1 and GSTM1, known inherited homozygous deletions are

equivalent to nonfunctional enzymes [20]. In the GSTP1 gene, a functional SNP between

adenosine (A) and guanosine (G) at base pair 313 leads to the expression of either Ile or

Val at codon 105. This polymorphism significantly affects enzyme activity [21].

PCR amplifications were performed in 50 µL reactions with ~100 ng of genomic DNA,

200 µM dNTPs (Epicentre Technologies, Madison, WI, USA), 10 x PCR Buffer II (Applied

Biosystems, Foster City, CA, USA), magnesium chloride (MgCl2), 0.5-1 U AmpliTaq Gold

(Applied Biosystems), and forward and reverse primers (Metabion, Planegg-Martinsried,

Germany). GSTM1 and GSTT1 deletions were analysed using a gel electrophoresis

method with β-globulin as internal control as described by Sreelekha et al [22]. GSTP1

(exon 5) was genotyped according to Jerónimo et al [23].

Clinical parameters

Information regarding cumulative cisplatin and oxaliplatin dose, follow-up time (time

from end of Pt therapy until inclusion in the current study), route of Pt administration,

co-administration of STS, and serum creatinine before start of chemotherapy were

collected from patient files. Additionally, serum creatinine was assessed at the time of

study.

Statistical analyses

Differences between Pt levels of cisplatin and oxaliplatin treated patients, control cancer

patients, and normal controls were evaluated using the Mann-Whitney U test. The

Wilcoxon signed rank test was used to test the difference between the renal function at

the time of chemotherapy and at follow-up. Correlations between plasma and pUF levels

of Pt treated patients were evaluated by using the Pearson correlation coefficient.

Long-term platinum retention

181

Determination of the Mann-Whitney U test, Wilcoxon signed rank test, and calculation of

the Pearson correlation coefficient were performed using SPSS. Possible relationships

between determinants and Pt levels were evaluated using non-linear mixed effects

modeling using NONMEM software (Version V1) (GloboMax LLC, Ellicott city, MD, USA).

The first order conditional estimation method was used throughout. It was assumed

that, due to the long follow-up, the treatment period was negligible compared to the

follow-up time. The significance of established relationships was assessed using the

likelihood ratio test.

Results

Participants

Table 1 summarises the characteristics of the participants. Participants were treated with

cisplatin for diverse tumour types, whereas all patients treated with oxaliplatin were

diagnosed with colorectal cancer. Five participants were treated with STS in

combination with 600 mg/m2 intra-arterially administered cisplatin. The range in the

follow-up time of patients was between 8 and 75 months.

Pt levels

Figure 1 shows Pt levels in plasma of 20 normal controls, 20 cancer control patients, 20

cisplatin treated cancer patients treated 18-75 months before entering this study, and 25

oxaliplatin treated cancer patients treated 8-33 months before entering this study.

Moreover, pUF levels from Pt treated patients are shown. All plasma and pUF samples of

the Pt treated patients were above the LLOQ of the method (plasma: 142-2.99x103 ng/L

median 647ng/L, pUF: 15.3-565 ng/L,, median 157 ng/L), whereas only three plasma

samples of control patients exceeded the LLOQ (plasma: <LLOQ-55.6 ng/L) (Figure 1).

For the cancer control patients, the signals of eight of the 20 samples were more than

three-fold higher than the noise. For the normal control patients, signals of two of the 20

samples were more than three-fold higher than the noise (Figure 1). Pt concentrations in

plasma of Pt treated patients were significantly higher than those in control patients

(Mann-Whitney U test, p<0.0001). Pt levels in pUF were highly correlated to levels in

plasma. The Pearson correlation coefficients were 0.97 and 0.95 for cisplatin and

oxaliplatin respectively. For cisplatin, on average 14.8% of plasma Pt was recovered in

pUF. This percentage was 24.2% for oxaliplatin. The percentage of Pt in pUF was not

dependent on the amount of plasma Pt. No difference was observed between 3 and 30

kDa filters. The levels of Pt-DNA adducts in PBMCs were below the LLOQ in all samples.

For nine of the 45 samples, Pt signals were higher than three-fold the noise signal.

Chapter 4.1

182

0.1

1

10

100

1000

10000

Plat

inum

con

cent

ratio

n (n

g/L)

Figure 1. Pt concentrations of 20 normal controls (plasma ), 20 cancer control patients (plasma ),

20 cancer patients who were treated with cisplatin 18-75 months before entering this study (plasma

and pUF ), 25 cancer patients who were treated with oxaliplatin 8-33 months before entering this

study (plasma and pUF ).

Ex vivo Assessment of DNA binding activity of Pt in pUF

Cisplatin and oxaliplatin added to pUF and incubated ex vivo with DNA revealed that

after five days, for both compounds, 21% of the added Pt was bound to DNA. The pUF

samples of the cisplatin patients demonstrated a DNA binding of 1.8 and 2.4%. For

oxaliplatin, these percentages were 0.78 and 1.1%. Hence, the experiment showed that

for cisplatin and oxaliplatin patient samples, on average, respectively 10 and 4.3% of the

total binding capacity of equivalent concentrations of parent compound was recovered

more than 8 months after the end of treatment. Pt contents of the DNA samples

incubated with the patient samples were just above the LLOQ of the method (2.1-6.0

pg).

LLOQ plasma

LLOQ pUF

Long-term platinum retention

183

Table 1. Characteristics of participants

Cisplatin Oxaliplatin

Gender (m/f) 13 m / 7 f 20 m / 5 f

Age at time of chemotherapy (median)

43 years

62 years

Age at time of follow-up (median)

49 years 64 years

Duration of follow-up 18-75 months (median 41) 8-33 months (median 18)

Tumour type Testicular carcinoma (9)

Yolk sac carcinoma (1)

Non small cell lung cancer (1)

Small cell lung cancer (1)

Head and neck carcinoma (8)

Colorectal carcinoma (25)

Cumulative dose 300-600 mg/m2 cisplatin (median 350)

195-390 mg/m2 Pt (median 227)

585-1170 mg/m2 oxaliplatin (median 878)

287-575 mg/m2 Pt (median 431)

Sodium thiosulfate 5 head and neck carcinoma patients treated intra-arterially with 600 mg/m2 cisplatin

NA

GSTM1 8/20 positive, 12/20 negative 10/25 positive, 15/25 negative

GSTT1 17/20 wildtype, 3/20 negative 21/25 positive, 4/25 negative

GSTP1 12/20 105Ile/105Ile-GSTP1

7/20 105Val/105Ile-GSTP1

1/20 105Val/105Val-GSTP1

9/25 105Ile/105Ile-GSTP1

10/25 105Val/105Ile-GSTP1

6/25 105Val/105Val-GSTP1

NA = not applicable

Ex vivo assessment of protein binding capacity of Pt in pUF

Table 2 shows the results of the experiments in which the plasma protein binding

capacity of Pt in plasma samples of patients was investigated. On average, the reduction

in Pt concentrations in pUF was 9.3 and 6.6% for cisplatin and oxaliplatin, respectively,

after storage for 144-278 days. When the two Pt compounds were considered separately,

no relationship between storage time and Pt reduction in pUF could be observed. For

both compounds, plasma Pt concentrations remained constant over time.

Chapter 4.1

184

Table 2. Pt concentrations in eight pUF samples before and after storage

Compound Percentage of plasma Pt recovered in pUF before storage

Percentage of plasma Pt recovered in pUF after storage

Days of storage

Cisplatin 19.4 8.1 278

19.7 8.4 278

15.1 5.5 276

12.3 7.4 276

Oxaliplatin 27.8 21.2 180

25.9 16.0 161

22.9 15.8 157

28.2 25.2 144

Ex vivo activity of STS

The results for the Pt concentrations in pUF samples with and without prior addition of

STS to plasma are depicted in Table 3. For cisplatin and oxaliplatin, respectively, a 2.3-

and 1.6-fold increase in Pt pUF concentrations was observed in samples prepared from

STS treated plasma. No proteins could be detected in any of the pUF samples.

Table 3. Pt concentrations in eight pUF samples with and without incubation with STS

Compound Percentage of plasma Pt recovered in pUF

Percentage of plasma Pt recovered in pUF after STS incubation

Factor increase of Pt in pUF after STS incubation

Cisplatin 12.4 17.5 1.4-fold

12.7 34.9 2.7-fold

13.2 30.7 2.3-fold

12.5 36.6 2.9-fold

Oxaliplatin 32.2 46.2 1.4-fold

20.0 39.5 2.0-fold

26.5 38.7 1.5-fold

27.2 40.4 1.5-fold

Long-term platinum retention

185

Effects of determinants on in vivo plasma Pt levels

Plasma Pt levels showed a gradual decline over time since the end of treatment (Figure

2). The decrease of the Pt levels followed a first order elimination profile. The observed

overlap of follow-up time for cisplatin and oxaliplatin (18-75 months vs 8-33 months)

justified combination of the data. The elimination profile could be described by a two-

compartment model of which the first elimination half-life (t1/2) was 1.2 months and the

second t1/2 28.5 months. The first t1/2 could only be estimated for oxaliplatin, because

there were no observations for cisplatin in this time period, whereas the second t1/2 could

be estimated for both compounds. No difference in the second t1/2 was observed for the

two compounds. The second t1/2 was related to age, in which a twofold increase in age

resulted in a 1.4-fold longer t1/2 (p=0.002).

0

500

1000

1500

2000

2500

3000

3500

0 20 40 60 80

Time since end treatment (months)

Plas

ma

plat

inum

con

cent

ratio

n (n

g/L)

Cisplatin

Oxaliplatin

Figure 2. Plasma Pt concentrations versus time since end of treatment

Pt levels were proportional to the cumulative dose. STS co-adminstration in combination

with intra-arterial cisplatin administration led to a 71% reduction in Pt levels (p<0.001).

An association between GSTT1, GSTM1, and GSTP1 genotypes and Pt levels could not be

established. The effect of the renal function on Pt levels was evaluated using the

glomerular filtration rate (GFR) which was calculated with the ‘Modification of Diet in

Renal Disease (MDRD)’-formula [24]. Median GRF values before start of the Pt

chemotherapy were 78 and 65 mL/min/1.73 m2 for cisplatin and oxaliplatin, respectively.

Chapter 4.1

186

At the time of the current study, GFR values of cisplatin patients were significantly

decreased to 55 mL/min/1.73 m2 (p<0.001), whereas oxaliplatin GFR values remained

constant (61 mL/min/1.73 m2). Figure 3 shows the MDRD GFR for patients at the start of

their chemotherapy treatment and at follow-up. The GFR at the time of chemotherapy

was significantly and conversely related to plasma Pt levels (p<0.01). Long-term plasma

Pt concentrations were lower when the GFR at the time of chemotherapy was higher.

0

20

40

60

80

100

120

0 20 40 60 80

Follow-up time (months)

MD

RD

GFR

a

0

20

40

60

80

100

120

0 10 20 30 40

Follow-up time (months)

MD

RD

GFR

b

Figure 3. GFR values for cisplatin (a) and oxaliplatin (b) treated patients at time of treatment (t=0) and

at follow-up

Long-term platinum retention

187

Discussion

Since the discovery of the antineoplastic effects of Pt-based compounds, cisplatin and

later oxaliplatin have developed into commonly used anticancer agents. The increased

survival of patients treated with these agents and the associated long-term side effects,

have initiated the investigation of the long-term pharmacokinetics, distribution, and

elimination of Pt-based compounds.

The current study focussed on the long-term pharmacokinetics of cisplatin and

oxaliplatin in plasma and pUF. We showed that plasma Pt levels of patients, treated with

cisplatin or oxaliplatin 8 to 75 months before participating in the current study, were still

>30-fold higher than the mean level of unexposed controls (normal volunteers and

cancer patients not treated with Pt). In earlier studies, raised plasma or serum Pt levels

one to 240 months after treatment with cisplatin were also reported [7,8,10,11].

In addition to elevated plasma Pt levels, the ultrafiltrable fraction of the plasma also

contained Pt species. This has not been reported before. The fraction of plasma Pt

recovered in the pUF was higher for oxaliplatin than for cisplatin, which could be a result

of a higher reactivity of cisplatin and hence, a more extensive protein binding of

cisplatin. Because pUF is generally considered to contain the pharmacologically active Pt

fraction [25], the question arises as to whether ultrafiltrable Pt measured up to 75

months after chemotherapy is composed exclusively of inactive Pt bound to low-

molecular-weight proteins, protein fragments, amino acids, and other molecules smaller

than 3 kDa, or whether it might also contain bound or unbound Pt with retained

reactivity. Because ICP-MS can not distinguish between unchanged cisplatin/oxaliplatin

and its metabolites or adducts, no information on the composition of the Pt species in

the pUF samples could be obtained. Unfortunately, up to now, no other technique is

sensitive enough to elucidate the chemical composition of the pool of Pt metabolites

and adducts that are probably present in the pUF samples.

Therefore, to address the question whether the Pt present in the pUF samples was still

reactive, we attempted to assess the Pt-DNA binding activity in vivo and ex vivo.

Although we were not able to quantify Pt-DNA adduct levels in PBMCs of the patients,

we did show that for cisplatin and oxaliplatin patient pUF samples, respectively 10% and

4.3% of the DNA binding activity of the parent compound was retained. The difference

between the reactivity of cisplatin and oxaliplatin samples is in agreement with the

difference in DNA binding activity of the parent compounds. In vitro studies reported

that cisplatin showed a 1.3 to 10-fold higher DNA binding compared to similar

concentrations of oxaliplatin [26,27]. Whether the Pt-DNA adducts are similar to the

adducts formed by the parent compounds remains to be established.

In addition to the remaining DNA binding properties, Pt in plasma also appeared to have

remaining protein binding capacity, which substantiates the observations that Pt species

recovered years after treatment may still show reactivity.

Chapter 4.1

188

The investigation of the Pt levels in pUF after the ex vivo incubation of plasma with STS

revealed that STS incubation resulted in higher Pt levels in pUF. This suggests that the Pt

protein binding is not irreversible and that a nucleophilic compound such as STS is

capable of releasing Pt from the proteins because the nucleophilic sulfide group,

possibly, has a higher binding affinity for Pt than proteins do. This observation, however,

should be interpreted with caution because the effect of STS on protein structure and

thus of the ability of the proteins to pass the ultrafilteration membrane could not be

evaluated because protein concentrations in pUF were too low to be detected.

Investigation of the effects of determinants on plasma Pt levels revealed a strong

relationship between Pt levels and the follow-up time. The relationship suggested that

plasma Pt was eliminated according to a first order elimination profile, which could be

best described by a two-compartment model for which the first t1/2 (1.2 months) could

be estimated for oxaliplatin and the second t1/2 (28.5 months) could be estimated for

cisplatin and oxaliplatin, with no difference between the two compounds. For cisplatin,

associations between plasma Pt levels and follow-up time were published before [7-11].

Hohnloser et al. reported an elimination half-life of 6.6 months for the time segment of

5.4 to 32 months and 26 months for the time segment of 21 to 107 months [11]. The last

t1/2 is in agreement with our findings. Although the two elimination half-lives found in

this study characterise the data between 8 and 75 months after the end of treatment,

the complete elimination of plasma Pt can presumably be described by numerous half-

lives which increase with a longer follow-up period. This was confirmed by

investigations of Gelevert et al., who calculated a t1/2 of 54 months for patients who were

treated with cisplatin 120 to 240 months before follow-up.

The long-term Pt plasma levels are, most probably, a result of Pt that has accumulated in

the tissue and is released into the bloodstream due to regeneration of tissue. Although,

for e.g. cisplatin, half of the Pt dose will be excreted during the first week after treatment,

it is estimated, based on animal models, that the long-term Pt retention in human might

still exceed 5% months after treatment [7]. Several investigations have shown raised Pt

levels in tissue samples up to years after chemotherapy [28,29]. The two compartments

in the Pt elimination observed in this study, could be explained by the Pt release from

fast regenerating tissue, followed by a release from slower regenerating tissue. This

hypothesis is supported by observations of Heydorn et al. and Gregg et al. Heydorn et al.

who reported that different tissues showed variable elimination half-lives. Gregg et al.

even showed that Pt levels in peripheral nerve tissue such as the dorsal root ganglia, did

not decay with time [29], which was expected considering the slow regeneration of

peripheral nerve tissue [30]. Because the rate of regeneration of tissue decreases with

age, the observation that a higher age was associated with a longer t1/2 was in

correspondence with the hypothesis that Pt levels recovered in the plasma represent the

regeneration of tissue.

Long-term platinum retention

189

In addition to follow-up time, the cumulative dose also appeared to be associated with

plasma Pt levels, which was expected as a higher initial Pt load will result in higher tissue

concentrations [29] and thus higher long-term plasma Pt levels. Few other studies also

suggested a correlation between cumulative Pt dose and long-term plasma or serum Pt

levels [8,11].

The association between renal function and plasma Pt levels was also evaluated. The

main excretion route of Pt agents is via the kidneys. In a previous study it was shown

that urinary Pt concentrations of patients studied 5.3 to 16.8 years after completion of

cisplatin chemotherapy, were strongly correlated to serum concentrations suggesting

rate limiting release of Pt from the tissues followed by fast renal excretion [10]. This

observation implies that no effect of renal function at follow-up on plasma Pt

elimination would be expected. The renal function before start of the treatment,

however, could affect the initial level of Pt accumulation in the tissues and thus long-

term plasma Pt levels. This hypothesis was in accordance with our observation that a

higher GFR was associated with lower plasma Pt levels.

A similar approach counts for the administration of STS at the time of chemotherapy.

The binding of STS to cisplatin could inactivate cisplatin resulting in a reduction of initial

Pt accumulation in tissue. Our results demonstrated that long-term plasma Pt levels

were reduced by 71% with co-administration of STS, which is in agreement with a

reduced tissue accumulation. Although, the intra-arterial administration of cisplatin in

patients who received STS could affect the venous plasma pharmacokinetics after

administration due to a first pass effect in tissue [31], we do not expect that the

administration route affects the total amount of Pt bound to tissue and thus the long-

term plasma Pt levels.

Although no reports have been published on the effects of GST genotypes on Pt

pharmacokinetics, GST genotypes could, considering the detoxification mechanism for

Pt from cells [32], affect the initial Pt elimination from the tissues and thus the long-term

tissue and plasma Pt levels. For the current study, however, no such relationship could

be established.

To summarise, our data suggest that plasma Pt levels are related to follow-up time, age,

cumulative dose, GFR at time of treatment, and STS use. Although no Pt-DNA adducts

could be detected in PBMCs, it was shown that Pt species in pUF were still present in a

reactive form. The clinical consequences of the observed reactivity remain to be

established.

Chapter 4.1

190

References 1. Rosenberg B, Vancamp L, Krigas T. Inhibition of cell division in escherichia coli by electrolysis products

from a platinum electrode. Nature 1965; 205: 698-9.

2. O'Dwyer PJ, Johnson SW. Current status of oxaliplatin in colorectal cancer. Semin Oncol 2003; 30: 78-87.

3. Grothey A. Oxaliplatin-safety profile: neurotoxicity. Semin Oncol 2003; 30: 5-13.

4. Bokemeyer C, Berger CC, Kuczyk MA, Schmoll HJ. Evaluation of long-term toxicity after chemotherapy for testicular cancer. J Clin Oncol 1996; 14: 2923-32.

5. Chaudhary UB, Haldas JR. Long-term complications of chemotherapy for germ cell tumours. Drugs 2003; 63: 1565-77.

6. Travis LB, Fossa SD, Schonfeld SJ, McMaster ML, Lynch CF, Storm H, Hall P, Holowaty E, Andersen A, Pukkala E, Andersson M, Kaijser M, Gospodarowicz M, Joensuu T, Cohen RJ, Boice JD, Jr., Dores GM, Gilbert ES. Second cancers among 40,576 testicular cancer patients: focus on long-term survivors. J Natl Cancer Inst 2005; 97: 1354-65.

7. Tothill P, Klys HS, Matheson LM, McKay K, Smyth JF. The long-term retention of platinum in human tissues following the administration of cisplatin or carboplatin for cancer therapy. Eur J Cancer 1992; 28: 1358-61.

8. Gietema JA, Meinardi MT, Messerschmidt J, Gelevert T, Alt F, Uges DR, Sleijfer DT. Circulating plasma platinum more than 10 years after cisplatin treatment for testicular cancer. Lancet 2000; 355: 1075-6.

9. Gelevert T, Messerschmidt J, Meinardi MT, Alt F, Gietema JA, Franke JP, Sleijfer DT, Uges DR. Adsorptive voltametry to determine platinum levels in plasma from testicular cancer patients treated with cisplatin. Ther Drug Monit 2001; 23: 169-73.

10. Gerl A, Schierl R. Urinary excretion of platinum in chemotherapy-treated long-term survivors of testicular cancer. Acta Oncol 2000; 39: 519-22.

11. Hohnloser JH, Schierl R, Hasford B, Emmerich B. Cisplatin based chemotherapy in testicular cancer patients: long term platinum excretion and clinical effects. Eur J Med Res 1996; 1: 509-14.

12. Stewart DJ, Benjamin RS, Luna M, Feun L, Caprioli R, Seifert W, Loo TL. Human tissue distribution of platinum after cis-diamminedichloroplatinum. Cancer Chemother Pharmacol 1982; 10: 51-4.

13. Barany E, Bergdahl IA, Bratteby LE, Lundh T, Samuelson G, Schutz A, Skerfving S, Oskarsson A. Relationships between trace element concentrations in human blood and serum. Toxicol Lett 2002; 134: 177-84.

14. Barany E, Bergdahl IA, Bratteby LE, Lundh T, Samuelson G, Schutz A, Skerfving S, Oskarsson A. Trace element levels in whole blood and serum from Swedish adolescents. Sci Total Environ 2002; 286: 129-41.

15. Goulle JP, Mahieu L, Castermant J, Neveu N, Bonneau L, Laine G, Bouige D, Lacroix C. Metal and metalloid multi-elementary ICP-MS validation in whole blood, plasma, urine and hair Reference values. Forensic Sci Int 2005; 153: 39-44.

16. Begerow J, Turfeld M, Dunemann L. Determination of physiological palladium, platinum, iridium and gold levels in human blood using double focusing magnetic sector field inductively coupled plasma mass spectrometry. Journal of Analytical Atomic Spectrometry 1997; 12: 1095-8.

17. Pluim D, Maliepaard M, van Waardenburg RC, Beijnen JH, Schellens JHM. 32P-postlabeling assay for the quantification of the major platinum-DNA adducts. Anal Biochem 1999; 275: 30-8.

18. Boom R, Sol CJ, Salimans MM, Jansen CL, Wertheim-van Dillen PM, van der NJ. Rapid and simple method for purification of nucleic acids. J Clin Microbiol 1990; 28: 495-503.

19. Brouwers EEM, Tibben MM, Rosing H, Hillebrand MJX, Joerger M, Schellens JHM, Beijnen JH. Sensitive inductively coupled plasma mass spectrometry assay for the determination of platinum originating from cisplatin, carboplatin, and oxaliplatin in human plasma ultrafiltrate. J Mass Spectrom 2006; 41: 1186-94.

20. Pemble S, Schroeder KR, Spencer SR, Meyer DJ, Hallier E, Bolt HM, Ketterer B, Taylor JB. Human glutathione S-transferase theta (GSTT1): cDNA cloning and the characterization of a genetic polymorphism. Biochem J 1994; 300: 271-6.

21. Ali-Osman F, Akande O, Antoun G, Mao JX, Buolamwini J. Molecular cloning, characterization, and expression in Escherichia coli of full-length cDNAs of three human glutathione S-transferase Pi gene variants. Evidence for differential catalytic activity of the encoded proteins. J Biol Chem 1997; 272: 10004-12.

22. Sreelekha TT, Ramadas K, Pandey M, Thomas G, Nalinakumari KR, Pillai MR. Genetic polymorphism of CYP1A1, GSTM1 and GSTT1 genes in Indian oral cancer. Oral Oncol 2001; 37: 593-8.

Long-term platinum retention

191

23. Jeronimo C, Varzim G, Henrique R, Oliveira J, Bento MJ, Silva C, Lopes C, Sidransky D. I105V polymorphism and promoter methylation of the GSTP1 gene in prostate adenocarcinoma. Cancer Epidemiol Biomarkers Prev 2002; 11: 445-50.

24. Levey AS, Bosch JP, Lewis JB, Greene T, Rogers N, Roth D. A more accurate method to estimate glomerular filtration rate from serum creatinine: a new prediction equation. Modification of Diet in Renal Disease Study Group. Ann Intern Med 1999; 130: 461-70.

25. Calvert H, Judson I, van der Vijgh WJ. Platinum complexes in cancer medicine: pharmacokinetics and pharmacodynamics in relation to toxicity and therapeutic activity. Cancer Surv 1993; 17: 189-217.

26. Goodisman J, Hagrman D, Tacka KA, Souid AK. Analysis of cytotoxicities of platinum compounds. Cancer Chemother Pharmacol 2006; 57: 257-67.

27. Saris CP, van de Vaart PJ, Rietbroek RC, Blommaert FA. In vitro formation of DNA adducts by cisplatin, lobaplatin and oxaliplatin in calf thymus DNA in solution and in cultured human cells. Carcinogenesis 1996; 17: 2763-9.

28. Heydorn K, Rietz B, Krarup-Hansen A. Distribution of platinum in patients treated with cisplatin determined by radiochemical neutron activation analysis. The Journal of Trace Elements in Experimental Medicine 1998; 11: 37-43.

29. Gregg RW, Molepo JM, Monpetit VJ, Mikael NZ, Redmond D, Gadia M, Stewart DJ. Cisplatin neurotoxicity: the relationship between dosage, time, and platinum concentration in neurologic tissues, and morphologic evidence of toxicity. J Clin Oncol 1992; 10: 795-803.

30. Jacobs WB, Fehlings MG. The molecular basis of neural regeneration. Neurosurgery 2003; 53: 943-8.

31. Sileni VC, Fosser V, Maggian P, Padula E, Beltrame M, Nicolini M, Arslan P. Pharmacokinetics and tumor concentration of intraarterial and intravenous cisplatin in patients with head and neck squamous cancer. Cancer Chemother Pharmacol 1992; 30: 221-5.

32. Goekkurt E, Hoehn S, Wolschke C, Wittmer C, Stueber C, Hossfeld DK, Stoehlmacher J. Polymorphisms of glutathione S-transferases (GST) and thymidylate synthase (TS)-novel predictors for response and survival in gastric cancer patients. Br J Cancer 2006; 94: 281-6.

Chapter 4.2

Persistent neuropathy after treatment with cisplatin and oxaliplatin

Elke E.M. Brouwers Alwin D.R. Huitema

Willem Boogerd Jos H. Beijnen

Jan H.M. Schellens

Submitted for publication

Chapter 4.2

194

Abstract

The aims of the current study were to assess persistent neuropathy in 45 patients more

than eight months and up to 6 years after treatment with cisplatin and oxaliplatin and to

determine the most adequate method to evaluate neuropathy. Furthermore, the effect

of demographic, therapy-related, biochemical, and pharmacogenetic characteristics on

persistent neuropathy were investigated. The assessment of neuropathy was performed

using a questionnaire and by neurological tests. In addition, neuropathy was evaluated

quantitatively by vibration threshold (VT) measurements.

In general, questionnaire scores and neurological tests were related to VT

measurements. Because VT determination gives the most objective information, VT

measurements were used for further analyses.

The analyses revealed that neuropathy of the hands was related to follow-up time, with

an observed recovery half-life of 6.8 years. No significant reversibility of neuropathy of

the feet within the observation period could be demonstrated. Furthermore, for

cisplatin, the severity of persistent neuropathy was related to the cumulative dose and

sodium thiosulfate use. Oxaliplatin induced neuropathy did not appear to be related to

the dose within the studied dose range. No relationship with platinum levels, renal

function, glutathione transferase genotypes, diabetes mellitus, alcohol use, or co-

medication could be demonstrated.

Cisplatin and oxaliplatin induced persistent neuropathy

195

Introduction

Cisplatin belongs to one of the most frequently used chemotherapeutics and is applied

extensively in the treatment of several tumour types. Chemotherapy with cisplatin

containing regimens is, however, often accompanied by severe side effects, such as

nephrotoxicity, ototoxicity, and peripheral sensory neuropathy. The search for platinum

(Pt) anticancer agents with less severe side effects and increased efficacy has led to the

development of several Pt-based compounds, including oxaliplatin. Oxaliplatin was first

introduced into clinical trials in 1986 [1] and is now part of the standard first-line

treatment in patients with colorectal cancer [2-4]. Unfortunately, patients treated with

this compound also frequently suffer from persistent peripheral sensory neuropathy.

Considering the fact that the life expectancy of patients treated with cisplatin and

oxaliplatin has improved, persistent side effects, such as neuropathy, can greatly affect

the quality of life of patients.

Clinical manifestations of cisplatin and oxaliplatin induced persistent neuropathy are

primarily sensory in nature and are mainly a consequence of effects on large myelinated

sensory fibers [5]. Parasthesia and dysesthesia in the hands and feet are the most

prominent symptoms [4,6]. The occurrence of L’hermitte’s sign is also reported,

indicating the involvement of the dorsal columns of the spinal cord [7-9]. For cisplatin, it

is mentioned that symptoms often occur or increase after completion of treatment [10]

with off-therapy deterioration up to approximately 6 months after withdrawal of

cisplatin [7,8]. In general, symptoms tend to recover slowly and the process is often

incomplete [6]. Oxaliplatin induced persistent neuropathy has been less well

characterised. Long-term follow-up is difficult because the prognosis of patients treated

with this drug is often less than 20 months [11]. It has, however, been mentioned that

the recovery from oxaliplatin neuropathy is faster and more complete than recovery

from cisplatin neuropathy [6].

Many studies have been performed to unravel the mechanism behind Pt-induced

persistent neuropathy, but still, the exact mechanism has not been clarified. The dorsal

root ganglion appears to be the primary site at which neural damage occurs. For

cisplatin, it was shown that Pt was retained in the dorsal root ganglia of patients who

were analysed post mortem [12,13] and that there were morphological changes visible

in these ganglia following treatment [12,14].

Despite the hypotheses on the mechanisms of Pt-induced sensory neuropathy, it still is

not fully elucidated why neuropathy has such a chronic character. It was suggested that

neuropathy may be persistent due to an irreversible damage at the time of

chemotherapy or due to a persistent Pt binding in the dorsal root ganglia.

The evaluation of sensory neuropathy is complicated by the subjective nature of most

assessment methods. Several approaches for the determination of the presence and

severity of neuropathy have been used in the past, among which were clinical

Chapter 4.2

196

neurological examinations [7,11,15], nerve conduction studies [7], questionnaires

[16,17], and the determination of the vibration threshold (VT measurements) [11,18,19].

Assessment of neuropathy by these methods has usually been done during or shortly

after cessation of treatment.

The aims of the current study were to assess persistent neuropathy in 45 patients more

than eight months after treatment with cisplatin and oxaliplatin and to determine the

most adequate method to evaluate neuropathy. The assessment of neuropathy was

done using a questionnaire and neurological tests. Besides, neuropathy was evaluated

quantitatively by vibration threshold (VT) measurements. In addition, we investigated

possible determinants of persistent neuropathy, such as Pt agent, follow-up time (time

from end of therapy until inclusion in the current study), cumulative dose, plasma Pt

levels, age, route of administration, renal function, glutathione S-transferase (GST)

genotypes, co-administration of calcium/magnesium with oxaliplatin or sodium

thiosulfate (STS) with intra-arterial cisplatin, co-medication, and co-morbidity

(alcoholism and diabetes).

Methods

Patients

For cisplatin, patients were selected at random from all patients who started treatment

between 2000 and 2004, received cumulative cisplatin doses of ≥ 300 mg/m2, and were

available for follow-up. For this pilot study, 20 patients of the 400 eligible patients were

included. To select the patients, random selections were performed on the 400 eligible

patients until 20 patients agreed to participate in the study. SPSS (SPSSinc, version 11.0,

Chicago, IL, USA) was used for random sample selection. For oxaliplatin, all available

patients who started treatment between 2000 and 2005 and received cumulative

oxaliplatin doses of ≥ 600 mg/m2 were approached for participation in the current study.

This led to an inclusion of 25 patients. The Medical Ethics Committee of the hospital

approved the study protocol and all patients gave their written informed consent.

Evaluation of nerve function

Neuropathic symptoms were assessed qualitatively using a questionnaire. The

questionnaire was composed of items from previously applied questionnaires which

were specifically associated with chemotherapy induced persistent sensory neuropathy

[20] and persistent Pt-induced neuropathy [16]. Questions addressed sensory symptoms

in the upper (10 questions) extremities (hands), lower (7 questions) extremities (feet),

Cisplatin and oxaliplatin induced persistent neuropathy

197

and the orofacial area (3 questions). Questionnaire subjects are summarised in Table 1.

The severity of the symptoms was graded as 0 (patient did not suffer from symptoms at

all), 1 (patient suffered from symptoms a little), 2 (patient suffered from symptoms pretty

much), and 3 (patient suffered from symptoms very much). When all the questions were

answered with a score 0, patients were considered as not having clinical neuropathy.

Patients who answered one or more questions with a score of 1 or more were

designated as suffering from clinical neuropathy. In addition to this dichotomous

scaling, the scores of the individual questions for each location (hands, feet, and

orofacial area) were summed to get an impression of the severity of the neuropathy. A

sum-score of 0 indicated the absence of clinical neuropathy.

Table 1. Questionnaire items

Hands/arms Feet/legs Orofacial

Numbness/tingling Numbness/tingling Numbness/tingling

Oversensitivity

Burning pain

Difficulty handling/feeling

Objects

L’hermitte’s sign

Disturbances in recognition of hot/cold/rough/smooth

Oversensitivity

Burning pain

Instability

L’hermitte’s sign

Disturbances in recognition of objects

The sensory nerve function was also assessed by two neurological balance tests: the

Sensitized Romberg Test (SRT) and Tandem Gait (TG). The SRT was used to determine

how long a patient was able to stand steady approximated (the toes of one foot

touching the heel of the foot in front of it), eyes open and then closed. The length was

recorded in seconds. More than ten seconds was considered as normal balance (0),

whereas less than ten seconds was considered as abnormal balance (1). The TG was used

to determine if a patient was capable to walk on a straight line with the heel of the first

foot touching the toes of the foot behind it. The results of the TG were recorded as

normal stability (0), difficulty to remain stable (1) and tendency to fall (2).

Neuropathy was assessed quantitatively by vibration threshold (VT) determination

[21,22] using a Vibrameter type IV device (Somedic AB, Stockholm, Sweden).

Measurements were performed in triplicate at the dorsum of the metacarpal bone of the

right and left index finger and of the dorsomedial aspect of the first metatarsal bone of

the right foot. The VT was recorded as micrometers of skin displacement and was

assessed using the method of limits, which showed an acceptable measure of

Chapter 4.2

198

reproducibility and validity [21,23,24]. All VT measurements were determined by the

same investigator to prevent inter-observer variability.

The dichotomous results from the questionnaires and VT method were compared and

the questionnaire was used to select an optimal cut-off value for the VT method.

Furthermore, the questionnaire sum-scores were compared to the continuous VT

measurements to see whether the VT measurements could predict the patients

experience of neuropathy. Besides, the associations between questionnaire sum-scores,

VT measurements, and neurological tests were evaluated.

Genotyping

Blood samples were drawn in 5 mL edta-containing tubes (Becton Dickinson Vacutainer

Systems, Plymouth, UK) for genotyping of detoxifying GST enzymes. Lymphocyte DNA

was isolated according to the method of Boom [25]. All samples were stored at –20 °C

until analysis. Polymorphisms in the genes encoding the enzymes GSTM1, GSTT1, and

GSTP1 were determined. In GSTT1 and GSTM1, known inherited homozygous deletions

are equivalent to nonfunctional enzymes and are encoded as positive and negative [26].

In the GSTP1 gene, a functional SNP between adenosine (A) and guanosine (G) at base

pair 313 leads to the expression of either Ile or Val at codon 105. This polymorphism

significantly affects enzyme activity [27].

PCR amplifications were performed in 50 µL reactions with ~100 ng of genomic DNA,

200 µM dNTPs (Epicentre Technologies, Madison, WI, USA), 10 x PCR Buffer II (Applied

Biosystems, Foster City, CA, USA), magnesiumchloride (MgCl2), 0.5-1 U AmpliTaq Gold

(Applied Biosystems), and forward and reverse primers (Metabion, Planegg-Martinsried,

Germany). GSTM1 and GSTT1 deletions were analysed using a gel electrophoresis

method with β-globulin as internal control as described by Sreelekha et al [28]. GSTP1

(exon 5) was genotyped according to Jerónimo et al [29].

Clinical parameters

Data regarding determinants that might affect sensory nerve function were collected.

Cumulative dose of the Pt agents, age, follow-up time, route of administration, co-

administration of calcium/magnesium or STS, co-morbidity e.g. alcoholism and diabetes,

co-medication, and serum creatinine before start of chemotherapy were collected from

patient files. Additionally, serum creatinine was assessed at the time of study. The

glomerular filtration rate (GFR) was estimated from serum creatinine using the

‘Modification of Diet in Renal Disease (MDRD)’-formula [30]. Plasma Pt levels at the time

of follow-up were described in a previous investigation [31].

Cisplatin and oxaliplatin induced persistent neuropathy

199

Statistical analyses

The Cronbach’s alpha coefficient was used to test the internal consistency of the

questionnaire using SPSS. It is allowed to use a sum-score when alpha values higher than

0.70 were achieved [32]. Receiver operating characteristic curves (ROC curves) were used

to assess the sensitivity and specificity of the VT method. ROC curves were composed

using SPSS. Boxplots (SPSS) were used to evaluate the association between VT values

and neurological tests. Possible relationships between determinants and neuropathy

were evaluated using non-linear mixed effects modeling using NONMEM software

(Version V1) (GloboMax LLC, Ellicott city, MD, USA). The first order conditional estimation

method was used throughout. It was assumed that, due to the long follow-up, the

treatment period was negligible compared to the follow-up time. The significance of

established relationships was assessed using the likelihood ratio test.

Results

Patients

Table 2 summarises the characteristics of the participants. Participants were treated with

cisplatin for diverse tumour types, whereas all patients treated with oxaliplatin were

diagnosed with colorectal cancer. Five participants were treated with STS in

combination with 600 mg/m2 intra-arterially administered cisplatin. For oxaliplatin, 24 of

25 participants received co-administration of Ca/Mg. The range in the follow-up time of

patients was between 8 and 75 months.

Scores of questionnaire and neurological tests

The questionnaire revealed that 26 patients (11/20 for cisplatin, 15/25 for oxaliplatin)

were classified as experiencing symptoms of sensory neuropathy in the hands.

Neuropathy in the feet was experienced by 32 patients (9/20 for cisplatin, 23/25

oxaliplatin). No patients showed orofacial symptoms. The Cronbach’s alpha coefficient

indicated that the internal consistency for the questions, which concentrated on the

hands (10 questions, α=0.71) and feet (7 questions, α=0.81), was suitable. Therefore, it

was allowed to sum the individual scores of the questions. To our opinion, however, one

should be cautious with the interpretation of the sum-scores. A higher score can not

automatically be designated as a higher level of clinical neuropathy. The sum-scores

ranged between 0 (no clinical neuropathy) and 9 (median: 2) for the hands and between

0 (no clinical neuropathy) and 18 (median: 3) for the feet.

Chapter 4.2

200

The SRT test revealed that 29 patients (11/20 for cisplatin, 18/24 for oxaliplatin) were

classified as having an abnormal balance. For one oxaliplatin patient the SRT test could

not be tested. The TG test was normal in 24 patients, whereas 15 (4/20 for cisplatin,

11/22 for oxaliplatin) patients experienced difficulties to remain stable and 3 (2/20 for

cisplatin, 1/22 for oxaliplatin) showed the tendency to fall. For three oxaliplatin patients

the TG could not be tested.

Table 1. Characteristics of participants

Cisplatin Oxaliplatin

Gender (m/f) 13 m / 7 f 20 m / 5 f

Age at time of chemotherapy 22-66 years (median 43) 40-73 years (median 62)

Age at time of follow-up 25-68 years (median 49) 41-76 years (median 64)

Duration of follow-up 18-75 months (median 41) 8-33 months (median 18)

Tumour type Testicular carcinoma (9)

Yolk sac carcinoma (1)

Non small cell lung cancer (1)

Small cell lung cancer (1)

Head and neck carcinoma (8)

Colorectal carcinoma (25)

Cumulative dose 300-600 mg/m2 cisplatin (median 350)

195-390 mg/m2 Pt (median 227)

585-1170 mg/m2 oxaliplatin (median 878)

287-575 mg/m2 Pt (median 431)

Sodium thiosulfate 5 head and neck carcinoma patients treated intra-arterially with 600 mg/m2 cisplatin

NA

Ca/Mg infusion NA 24/25

GSTM1 8/20 positive, 12/20 negative 10/25 positive, 15/25 negative

GSTT1 17/20 wildtype, 3/20 negative 21/25 positive, 4/25 negative

GSTP1 12/20 105Ile/105Ile-GSTP1

7/20 105Val/105Ile-GSTP1

1/20 105Val/105Val-GSTP1

9/25 105Ile/105Ile-GSTP1

10/25 105Val/105Ile-GSTP1

6/25 105Val/105Val-GSTP1

Plasma Pt levels 142-1.15x103 ng/L 460-2.99x103 ng/L

GFR at time of chemotherapy 51.4-89.5 mL/min/1.73m2 (median 77.6)

36.5-74.1 mL/min/1.73m2

(median 55)

GFR at time of follow-up 26.7-142 mL/min/1.73m2

(median 64.8) 33.6-79.6 mL/min/1.73m2 (median 61.2)

NA = not applicable

Cisplatin and oxaliplatin induced persistent neuropathy

201

Vibration perception results

In Figures 1a and b, the individual VT values of the patients for, respectively, the hands

and feet are plotted versus the age of the patients. It has been shown that age is an

important confounder of the relation between VT and neuropathy [23]. The figures also

show the mean VT values (+ 2 standard deviations (SD)) plotted versus age for a normal

population consisting of 110 controls [23]. The VT-cut-off of 2 SD has been used to

classify patients as having a normal or abnormal sensory nerve function [15,33]. As can

be derived from Figure 1, hand-VT values which deviated ≥ 2 SD from the mean, were

observed in 15 patients (8/20 for cisplatin, 7/25 for oxaliplatin). For the feet-VT, this value

was 18 (7/20 for cisplatin, 11/25 for oxaliplatin).

Feet

0.0

0.1

1.0

10.0

100.0

10 30 50 70 90Age (years)

Am

plitu

de (u

m)

Figure 1. Hands and feet VT values for cisplatin ( ) and oxaliplatin ( ) treated patients and the mean

values ( ) (+2 SD ( )) for 110 control patients [23] plotted against the age.

Comparison of questionnaire scores with vibration perception results

The classification of patients on the questionnaire data (neuropathy or no neuropathy)

showed that a higher number of patients actually experienced neuropathy than the

number that was designated as neuropathic using the VT test. Sensitivity and specificity

values of the dichotomous VT scale with a cut-off of 2 SD were calculated considering

the dichotomous questionnaire score as the correct values (Table 3a). Obviously, not all

patients were classified correctly. Therefore, it was investigated whether the application

of the VT test could be improved by using other VT cut-off values. The sensitivity and

specificity of several cut-off values were investigated by the composition of ROC curves.

The dichotomous questionnaire scores were used as state variables. The natural

logarithms (ln) of the VT values which were normalised for age, were used as test

variables. Normalised VT values were obtained by dividing the VT values of the patients

by the mean VT values assessed in controls with the same age as the patient [23]. The

natural logarithms (ln) of these values ranged from -0.342 to 4.20 for the hands and from

Hands

0.0

0.1

1.0

10.0

100.0

10 30 50 70 90Age (years)

Am

plitu

de (u

m)

Chapter 4.2

202

-1.39 to 4.30 for the feet. Optimal sensitivity and specificity values were obtained with an

ln VT value of 0.860 for the hands and 1.35 for the feet. These values corresponded to

cut-offs of 1.4 and 1.7 SD, respectively for the hands and feet. Using these cut-offs,

sensitivity and specificity values were respectively 77 and 63% for the hands and 72 and

83% for the feet (Table 3b). These values were significantly better than those obtained

with a cut-off value of 2 SD. Using these cut-off values elevated VT values were observed

in 27 patients for the hands and 25 patients for the feet.

The use of a dichotomous VT scale solely provides information regarding the presence

or absence of neuropathy and quantitative information on further grading is not

obtained. Hence, in addition to the generally used dichotomous scale, the ln VT values

were also used to describe the neuropathy based on a continuous scale. To assess

whether the ln VT values could predict the subjective experience of the patients, the

sum-scores of the questionnaire were plotted versus the ln VT values (Figure 2). It was

observed that, in general, raised questionnaire scores were coupled to high ln VT scores

for the hands as well as for the feet.

Table 3a. Sensitivity and specificity for a VT-cut-off value of 2 SD

Questionnaire

Hands (2 SD) Present Absent Feet (2 SD) Present Absent

Present 10 5 Present 16 2

VT te

st

Absent 16 14 Absent 16 10

Total 26 19 Total 32 12 Hands sensitivity: 38%, specificity: 74% Feet sensitivity: 50%, specificity: 83%

Table 3b. Sensitivity and specificity for VT-cut-off values of 1.4 (hands) and 1.7 (feet) SD

Questionnaire

Hands (1.4 SD) Present Absent Feet (1.7 SD) Present Absent

Present 20 7 Present 23 2

VT te

st

Absent 6 12 Absent 9 10

Total 26 19 Total 32 12 Hands sensitivity: 77%, specificity: 63% Feet sensitivity: 72%, specificity: 83%

Cisplatin and oxaliplatin induced persistent neuropathy

203

Feet

-5

0

5

10

15

20

-3 -2 -1 0 1 2 3 4 5

Ln VT

Scor

e qu

estio

nnai

re

Figure 2. Questionnaire sum-scores for the hands and feet for cisplatin patients without neuropathy ( ), cisplatin patients with neuropathy (▲), oxaliplatin patients without neuropathy (□), and oxaliplatin patients with neuropathy ( )

Comparison of neurological tests with questionnaire scores and vibration perception

results

To assess whether the questionnaire scores were associated to the neurological tests,

the sum-scores of the questionnaire results for the feet were plotted versus the SRT and

TG scores (Figure 3a). The association between the ln VT values for the feet and SRT and

TG scores was evaluated by boxplots (Figure 3b). It was observed that, in general, high ln

VT values were coupled to abnormal SRT and TG scores. This relationship, however, was

not observed for the questionnaire and SRT and TG scores. Therefore, VT measurements

were used for further data analyses.

0

5

10

15

20

-1 0 1 2 3

Tandem Gait

Scor

e qu

estio

nnai

re

Figure 3a. Questionnaire sum-scores for the feet versus Sensitized Romberg Test (0=normal balance,

1= abnormal balance) and Tandem Gait scores (0=normal stability, 1=difficulty to remain stable,

2=tendency to fall)

0

5

10

15

20

-1 0 1 2

Sensitized Romberg Test

Scor

e qu

estio

nnai

re

Hands

-202468

10

-1 0 1 2 3 4 5

Ln VT

Scor

e qu

estio

nnai

re

Chapter 4.2

204

0 1

-2-1

01

23

45

Sensitized Romberg Test

lnVT

0 1 2

-2-1

01

23

45

Tandem Gait

lnVT

Figure 3b. Feet ln VT values versus Sensitized Romberg Test (0=normal balance, 1= abnormal balance)

and Tandem Gait scores (0=normal stability, 1=difficulty to remain stable, 2=tendency to fall)

Effects of determinants on ln VT

Possible determinants that could affect the sensory nerve function were evaluated using

the VT measurements. Ln VT values for the hands and feet were plotted against the

follow-up time in Figure 4a and b, respectively. The ln VT values for the hands showed a

small decline with follow-up time following a first order decay with a half-life (t1/2) of 6.8

(± 3.1) years. The observed recovery t1/2 was similar for cisplatin and oxaliplatin and was

related to age, in which a twofold increase in age resulted in a three-fold longer recovery

t1/2 (p=0.02). For the feet, no decline of the ln VT value with time was observed.

Plotting of the dichotomous results of the VT test versus the follow-up time did not

result in any additional information. The continuous scale for neuropathy assessment

obtained from the ln VT measurement was considered more informative than the

dichotomous classification and was, therefore, chosen for further data analysis.

The ln VT values for cisplatin were higher than for oxaliplatin. For cisplatin, ln Vt values

were proportional to the dose. On the contrary, no dose dependency was observed for

oxaliplatin. STS co-administration with intra-arterial cisplatin administration led to 56%

reduction of ln VT for cisplatin (p=0.002) both in the hands and feet. No effect of

concomitant infusion of calcium and magnesium on VT measurements for oxaliplatin

treated patients could be demonstrated.

Curves of modelled ln VT values against the follow-up time for a median aged patient

who received 600 mg/m2 cisplatin with STS, a patient who received 300 mg/m2 cisplatin,

and a patient who received 878 mg/m2 oxaliplatin are depicted in figure 4.

Cisplatin and oxaliplatin induced persistent neuropathy

205

-5.0

0.0

5.0

0 20 40 60 80 100

Follow-up time (months)

Ln V

T ha

nds

Figure 4a. Ln VT values for the hands for cisplatin (▲) and oxaliplatin ( ) treated patients versus time

and a median patients treated with 300 mg/m2 cisplatin ( ), with 600 mg/m2 cisplatin combined

with STS (- - - - -), and with oxaliplatin 878 mg/m2 ( )

-5.0

0.0

5.0

0 20 40 60 80 100

Follow-up time (months)

Ln V

T fe

et

Figure 4b. Ln VT values for the feet for cisplatin ( ) and oxaliplatin (□) treated patients versus time

and a median patients treated with 300 mg/m2 cisplatin ( ), with 600 mg/m2 cisplatin combined

with STS (- - - - -), and with oxaliplatin 878 mg/m2 ( )

Chapter 4.2

206

The observed plasma Pt levels at the time of follow-up and GST allele frequencies are

shown in Table 2. It was not considered plausible that Pt levels at the time of follow-up

were related to nerve function, because plasma Pt levels decrease with time [31],

whereas Pt levels in ganglia remain constant over time [12]. Therefore, Pt levels were

extrapolated to the time of treatment using the elimination half-life [31]. No relationship

between plasma Pt levels and ln VT values was observed. Furthermore, an association

between GSTT1, GSTM1, and GSTP1 genotypes and ln VT values could not be established.

The effect of the renal function on ln VT values was evaluated using the MDRD-GFR.

Median GRF values before start of the Pt chemotherapy were 78 and 65 mL/min/1.73m2

for cisplatin and oxaliplatin, respectively (Table 2). At the time of the current study, GFR

values of cisplatin patients were significantly decreased to 55 mL/min/1.73m2 (p<0.001),

whereas oxaliplatin GFR values remained constant (61 mL/min/1.73m2). No association

was observed between GFR and ln VT values. Furthermore, alcohol use, diabetes

mellitus, or co-medication did not seem to affect ln VT values.

Discussion

Cisplatin and oxaliplatin are frequently used chemotherapeutic agents. Their use,

however, is hampered by a peripheral sensory neuropathy, which can, even years after

cessation of treatment, seriously affect the quality of life.

The aims of the current study were to assess persistent neuropathy in 45 patients more

than eight months and up to 75 months after treatment with cisplatin and oxaliplatin

and to determine the most adequate method to evaluate neuropathy. The assessment

of neuropathy was performed using a questionnaire and by neurological tests. Besides,

neuropathy was assessed quantitatively by VT measurements.

Comparison of the questionnaire results and VT values showed that the generally used 2

SD cut-off value for the VT measurement may not lead to adequate classification of

neuropathy. Therefore, classification was improved by using different cut-off values.

However, by classifying the dichotomous VT measurements, quantitative information

regarding the nerve function was lost. Hence, the ln VT values were used to describe the

neuropathy based on a continuous scale. It was observed that, in general, raised

questionnaire sum-scores were coupled to high ln VT scores for the hands as well as for

the feet. When questionnaire sum-scores and ln VT values were plotted versus the scores

of the neurological tests, ln VT values were more obviously associated to the

neurological tests than the questionnaire sum-scores. This illustrates that the VT

determination gives more objective information than questionnaire sum-scores.

Because quantitative measurements may provide more precise and objective measures

of neuropathy, the VT determination was used to perform further data analyses.

Cisplatin and oxaliplatin induced persistent neuropathy

207

Investigation of the effects of determinants on ln VT values revealed that the ln VT values

for the hands declined with follow-up time. The recovery t1/2 of 6.8 years, however,

suggests that for both cisplatin and oxaliplatin, the reversibility is slow and incomplete.

The observation that a twofold increase in age resulted in a three-fold longer recovery

t1/2, suggests that the effect of age on VT values in our population was larger than the

effect observed in the population of Goldberg et al. [23]. Probably, the sensory nerve

function of patients treated with Pt agents is more susceptible to age than the nerve

function of normal individuals.

The observations that the ln VT values for the feet did not decline with time, suggests

that neuropathy in the feet is an even more important issue than neuropathy in the

hands. This was in accordance with an investigation performed by Land et al. who

mentioned that for oxaliplatin, 18 months after the start of treatment, neuropathy was

primarily present in the feet [17]. The apparent larger effect of Pt agents on the feet

compared to the hands is in accordance with the observation that other determinants

such as age also affect the sensory nerve function in the feet more than in the hands

[23].

The higher ln VT values observed for cisplatin compared to oxaliplatin could, probably,

be explained by the higher activity and thus toxicity of cisplatin.

The ln VT values for cisplatin were proportional to the cumulative dose. This association

was already extensively described in previous studies [8,34,35]. In contrast, no

association between dose and ln VT for oxaliplatin was observed. This might be a

consequence of the limited dose range studied in these patients. In contrast to our

observations, other studies mentioned the presence of a dose dependency for persistent

oxaliplatin induced neuropathy [6,11].

The ln VT values were not related to the plasma Pt levels at the time of chemotherapy.

Plasma Pt levels at the time of chemotherapy were estimated by extrapolating the

plasma Pt levels at follow-up using the estimated elimination t1/2. Because the

elimination t1/2 was estimated using only one data point per patient, this could have

resulted in an inaccurate estimation of Pt levels at the time of chemotherapy. Therefore,

to evaluate the relationship between plasma Pt levels and ln VT, more data points per

patient are indispensable.

The association between STS, calcium/magnesium co-administration, and ln VT values

was also evaluated. STS co-administration with intra-arterial cisplatin administration led

to 56% reduction in ln VT for cisplatin. By binding to cisplatin, STS can inactivate

cisplatin. This might lead to a reduction of Pt load in tissue and thereby a reduction of

toxicity. Because 24 of the 25 evaluated oxaliplatin patients received concomitant

calcium/magnesium infusions, the group of patients not treated with calcium and

magnesium was too small to be able to observe an effect on the ln VT values.

Chapter 4.2

208

It would be reasonable that the renal function at the time of chemotherapy affected the

severity of neuropathy. A better renal function could imply a faster initial elimination of

Pt, leading to lower tissue levels and thus less toxicity. This association, however, was not

observed in the current population. This could be due to the small sample size and a

limited variation in GFR.

Genetic factors, that potentially influence the pharmacokinetics of an anticancer drug

and thereby the development of drug toxicity in patients could be relevant

determinants in this study. The detoxifying GST enzymes are thought to participate

collectively in the intracellular metabolism and detoxification [36,37]. Recent

investigations showed that patients homozygous or heterozygous for the GSTP1 105Val

allele were less susceptible for developing severe oxaliplatin-induced neurotoxicity [38].

The presence of both alleles of 105Val-GSTP1 offered protection against cisplatin-induced

hearing impairment [39]. Although it might be suspected that mutations in the GSTs

genes could affect sensory neuropathy, we did not observe an association between ln VT

and any of the three (GSTM1, GSTT1 and GSTP1) genes investigated, which is probably

due to the small sample size of the current population.

To summarise, the effect of possible determinants on neuropathy could be evaluated

using continuous VT values, which provides the most extensive information about the

severity of neuropathy. Our data suggest that neuropathy in the hands is slowly

reversible, whereas no reversibility of neuropathy in the feet could be demonstrated

during follow-up in the current study. Furthermore, for cisplatin, the severity of

neuropathy was associated with cumulative dose and STS use. Oxaliplatin induced

neuropathy did not appear to be related to the dose. No relationship with renal function,

GST genotypes, diabetes mellitus, alcohol use, or co-medications could be

demonstrated.

Cisplatin and oxaliplatin induced persistent neuropathy

209

References 1. Mathé G, Kidani Y, Triana K, Brienza S, Ribaud P, Goldschmidt E, Ecstein E, Despax R, Musset M, Misset JL. A

phase I trial of trans-1-diaminocyclohexane oxalato-platinum (l-OHP). Biomed Pharmacother 1986; 40: 372-6.

2. Pendyala L, Creaven PJ. In vitro cytotoxicity, protein binding, red blood cell partitioning, and biotransformation of oxaliplatin. Cancer Res 1993; 53: 5970-6.

3. Schmidt W, Chaney SG. Role of carrier ligand in platinum resistance of human carcinoma cell lines. Cancer Res 1993; 53: 799-805.

4. Chaudhary UB, Haldas JR. Long-term complications of chemotherapy for germ cell tumours. Drugs 2003; 63: 1565-77.

5. Alberts DS, Noel JK. Cisplatin-associated neurotoxicity: can it be prevented? Anticancer Drugs 1995; 6: 369-83.

6. Grothey A. Oxaliplatin-safety profile: neurotoxicity. Semin Oncol 2003; 30: 5-13.

7. Boogerd W, Bokkel Huinink WW, Dalesio O, Hoppenbrouwers WJ, van der Sande JJ. Cisplatin induced neuropathy: central, peripheral and autonomic nerve involvement. J Neurooncol 1990; 9: 255-63.

8. Siegal T, Haim N. Cisplatin-induced peripheral neuropathy. Frequent off-therapy deterioration, demyelinating syndromes, and muscle cramps. Cancer 1990; 66: 1117-23.

9. Taieb S, Trillet-Lenoir V, Rambaud L, Descos L, Freyer G. Lhermitte sign and urinary retention: atypical presentation of oxaliplatin neurotoxicity in four patients. Cancer 2002; 94: 2434-40.

10. LoMonaco M, Milone M, Batocchi AP, Padua L, Restuccia D, Tonali P. Cisplatin neuropathy: clinical course and neurophysiological findings. J Neurol 1992; 239: 199-204.

11. Pietrangeli A, Leandri M, Terzoli E, Jandolo B, Garufi C. Persistence of high-dose oxaliplatin-induced neuropathy at long-term follow-up. Eur Neurol 2006; 56: 13-6.

12. Gregg RW, Molepo JM, Monpetit VJ, Mikael NZ, Redmond D, Gadia M, Stewart DJ. Cisplatin neurotoxicity: the relationship between dosage, time, and platinum concentration in neurologic tissues, and morphologic evidence of toxicity. J Clin Oncol 1992; 10: 795-803.

13. Heydorn K, Rietz B, Krarup-Hansen A. Distribution of platinum in patients treated with cisplatin determined by radiochemical neutron activation analysis. The Journal of Trace Elements in Experimental Medicine 1998; 11: 37-43.

14. McKeage MJ, Hsu T, Screnci D, Haddad G, Baguley BC. Nucleolar damage correlates with neurotoxicity induced by different platinum drugs. Br J Cancer 2001; 85: 1219-25.

15. Postma TJ, Hoekman K, van Riel JM, Heimans JJ, Vermorken JB. Peripheral neuropathy due to biweekly paclitaxel, epirubicin and cisplatin in patients with advanced ovarian cancer. J Neurooncol 1999; 45: 241-6.

16. Leonard GD, Wright MA, Quinn MG, Fioravanti S, Harold N, Schuler B, Thomas RR, Grem JL. Survey of oxaliplatin-associated neurotoxicity using an interview-based questionnaire in patients with metastatic colorectal cancer. BMC Cancer 2005; 5: 116.

17. Land SR, Kopec JA, Cecchini RS, Ganz PA, Wieand HS, Colangelo LH, Murphy K, Kuebler JP, Seay TE, Needles BM, Bearden JD, III, Colman LK, Lanier KS, Pajon ER, Jr., Cella D, Smith RE, O'Connell MJ, Costantino JP, Wolmark N. Neurotoxicity from oxaliplatin combined with weekly bolus fluorouracil and leucovorin as surgical adjuvant chemotherapy for stage II and III colon cancer: NSABP C-07. J Clin Oncol 2007; 25: 2205-11.

18. van der Hoop RG, Vecht CJ, van der Burg ME, Elderson A, Boogerd W, Heimans JJ, Vries EP, van Houwelingen JC, Jennekens FG, Gispen WH, . Prevention of cisplatin neurotoxicity with an ACTH(4-9) analogue in patients with ovarian cancer. N Engl J Med 1990; 322: 89-94.

19. Petersen PM, Hansen SW. The course of long-term toxicity in patients treated with cisplatin-based chemotherapy for non-seminomatous germ-cell cancer. Ann Oncol 1999; 10: 1475-83.

20. Postma TJ, Aaronson NK, Heimans JJ, Muller MJ, Hildebrand JG, Delattre JY, Hoang-Xuan K, Lanteri-Minet M, Grant R, Huddart R, Moynihan C, Maher J, Lucey R. The development of an EORTC quality of life questionnaire to assess chemotherapy-induced peripheral neuropathy: the QLQ-CIPN20. Eur J Cancer 2005; 41: 1135-9.

21. Gerr F, Letz R. Vibrotactile threshold testing in occupational health: a review of current issues and limitations. Environ Res 1993; 60: 145-59.

22. Chong PS, Cros DP. Technology literature review: quantitative sensory testing. Muscle Nerve 2004; 29: 734-47.

Chapter 4.2

210

23. Goldberg JM, Lindblom U. Standardised method of determining vibratory perception thresholds for diagnosis and screening in neurological investigation. J Neurol Neurosurg Psychiatry 1979; 42: 793-803.

24. Chuang HY, Schwartz J, Tsai SY, Lee ML, Wang JD, Hu H. Vibration perception thresholds in workers with long-term exposure to lead. Occup Environ Med 2000; 57: 588-94.

25. Boom R, Sol CJ, Salimans MM, Jansen CL, Wertheim-van Dillen PM, van der NJ. Rapid and simple method for purification of nucleic acids. J Clin Microbiol 1990; 28: 495-503.

26. Pemble S, Schroeder KR, Spencer SR, Meyer DJ, Hallier E, Bolt HM, Ketterer B, Taylor JB. Human glutathione S-transferase theta (GSTT1): cDNA cloning and the characterization of a genetic polymorphism. Biochem J 1994; 300: 271-6.

27. Ali-Osman F, Akande O, Antoun G, Mao JX, Buolamwini J. Molecular cloning, characterization, and expression in Escherichia coli of full-length cDNAs of three human glutathione S-transferase Pi gene variants. Evidence for differential catalytic activity of the encoded proteins. J Biol Chem 1997; 272: 10004-12.

28. Sreelekha TT, Ramadas K, Pandey M, Thomas G, Nalinakumari KR, Pillai MR. Genetic polymorphism of CYP1A1, GSTM1 and GSTT1 genes in Indian oral cancer. Oral Oncol 2001; 37: 593-8.

29. Jeronimo C, Varzim G, Henrique R, Oliveira J, Bento MJ, Silva C, Lopes C, Sidransky D. I105V polymorphism and promoter methylation of the GSTP1 gene in prostate adenocarcinoma. Cancer Epidemiol Biomarkers Prev 2002; 11: 445-50.

30. Levey AS, Bosch JP, Lewis JB, Greene T, Rogers N, Roth D. A more accurate method to estimate glomerular filtration rate from serum creatinine: a new prediction equation. Modification of Diet in Renal Disease Study Group. Ann Intern Med 1999; 130: 461-70.

31. Brouwers EEM, Huitema ADR, Beijnen JH, Schellens JHM. Long-term platinum retention after treatment with cisplatin and oxaliplatin. Submitted 2007.

32. Bland JM, Altman DG. Statistics note: Cronbach's alpha. BMJ 1997; 314: 572.

33. Somedic AB Stockholm Sweden. Manual for vibrameter Type III.1985

34. van der Hoop RG, van der Burg ME, Bokkel Huinink WW, van Houwelingen C, Neijt JP. Incidence of neuropathy in 395 patients with ovarian cancer treated with or without cisplatin. Cancer 1990; 66: 1697-702.

35. Bokemeyer C, Berger CC, Kuczyk MA, Schmoll HJ. Evaluation of long-term toxicity after chemotherapy for testicular cancer. J Clin Oncol 1996; 14: 2923-32.

36. Tew KD. Glutathione-associated enzymes in anticancer drug resistance. Cancer Res 1994; 54: 4313-20.

37. Zhang K, Chew M, Yang EB, Wong KP, Mack P. Modulation of cisplatin cytotoxicity and cisplatin-induced DNA cross-links in HepG2 cells by regulation of glutathione-related mechanisms. Mol Pharmacol 2001; 59: 837-43.

38. Lecomte T, Landi B, Beaune P, Laurent-Puig P, Loriot MA. Glutathione S-transferase P1 polymorphism (Ile105Val) predicts cumulative neuropathy in patients receiving oxaliplatin-based chemotherapy. Clin Cancer Res 2006; 12: 3050-6.

39. Oldenburg J, Kraggerud SM, Cvancarova M, Lothe RA, Fossa SD. Cisplatin-induced long-term hearing impairment is associated with specific glutathione s-transferase genotypes in testicular cancer survivors. J Clin Oncol 2007; 25: 708-14.

Chapter 5 Environmental monitoring

of platinum agents

Chapter 5.1

Monitoring of platinum surface contamination in seven Dutch hospital pharmacies using inductively coupled

plasma mass spectrometry

Elke E.M. Brouwers Alwin D.R. Huitema

Eke N. Bakker Jan Willem Douma

Kirsten J.M. Schimmel Geke van Weringh

Paul J. de Wolf Jan H.M. Schellens

Jos H. Beijnen

Submitted for publication

Chapter 5.1

216

Abstract

Objective: To develop, validate, and apply a method for the determination of platinum

(Pt) surface contamination, originating from cisplatin, oxaliplatin, and carboplatin.

Methods: Inductively coupled plasma mass spectrometry was used to determine Pt in

wipe samples. The sampling procedure and the analytical conditions were optimised

and the assay was validated. The method was applied to measure surface contamination

in seven Dutch hospital pharmacies.

Results: The developed method allowed reproducible quantification of 0.50 ng/L Pt (5

pg/wipe sample). Recoveries for stainless steel and linoleum surfaces ranged between

50.4% and 81.4% for the different Pt compounds tested. Pt contamination was reported

in 88% of the wipe samples. Although a substantial variation in surface contamination of

the pharmacies was noticed, in most pharmacies, the laminar-airflow (LAF) hoods, the

floor in front of the LAF hoods, door handles, and handles of service hatches showed

positive results. This demonstrates that contamination is spread throughout the

preparation rooms. Conclusion: We developed and validated an ultrasensitive and

reliable ICP-MS method for the determination of Pt in surface samples. Surface

contamination with Pt was observed in all hospital pharmacies sampled. The

interpretation of these results is, however, complicated.

Environmental monitoring

217

Introduction

Cytotoxic drugs are widely used for the treatment of cancer. Occupational exposure to

these drugs has been recognised as a potential health hazard since 1970s [1,2]. Because

cytotoxic drugs can affect the DNA, RNA, or protein synthesis, many of these drugs are

classified as being carcinogenic, mutagenic, or teratogenic to humans [3]. Skin contact

with cytotoxic drugs, due to contamination of the work area or contamination of

packaging material, seems to play an important role in the uptake of these drugs by

hospital personnel [4,5]. Therefore, strict health and safety rules have been established

and applied for the handling of these agents. Evidently, the potential health risks for

persons manipulating cytotoxic drugs, such as pharmacists, pharmacy technicians,

nurses, and cleaners, however, still are a concern. This concern is consolidated by a

number of recent publications demonstrating workplace contamination [6-12] and

contamination of packaging of cytotoxic drugs [7,13-15]. Moreover, detection of

cytotoxic agents in urine [4,16-22] and blood [23] of personnel who were involved in

preparation or administration has been reported with increasing frequency.

The relationship between prolonged exposure to small quantities of cytotoxic drugs and

harmful effects is difficult to establish. Based on current scientific knowledge, it is

impossible to set a level of exposure that, beyond doubt, will not cause adverse effects.

Because no regulations on the maximal acceptable amount of contamination for these

drugs have been set so far, hospitals should aim for the lowest contamination as is

reasonably achievable. Monitoring of contamination, therefore, is essential. This can aid

in the identification of the main exposure routes and in assessing the effectiveness of

cleaning and working procedures. Evaluation of environmental contamination will,

moreover, lead to an increase of the consciousness among personnel, concerning the

handling of the chemotherapeutic agents. This can lead to an improvement of and the

compliance with working and cleaning procedures. Wipe sampling is a common method

to monitor surfaces for the presence of cytotoxic drugs. Hence, sensitive and validated

methods are indispensable to be able to detect the relatively low quantities of drug

present on surfaces.

Platinum (Pt) coordination complexes, such as cisplatin, oxaliplatin, and carboplatin, play

a major role in the treatment of a variety of tumours. As a result, large amounts of these

agents are processed in hospital pharmacies. Several wipe sample methods for Pt

containing drugs have been used in earlier studies and Pt was detected as a surface

contaminant in many of the workplaces [9-12] or drug vials [13,14] investigated. A

description of the validation of the analytical methods, however, has been scarce.

Validation results were mentioned briefly for the method of Ziegler et al., using electro

thermal vaporisation coupled to inductively coupled plasma mass spectrometry (ICP-

MS) [9]. Raghavan et al. described the validation of a high-performance liquid

chromatography method for the determination of cisplatin in cleaning validation

Chapter 5.1

218

samples [24]. The lower limit of quantification (LLOQ) of this method, was 500 ng/L,

which is high compared to the limits achievable with for example ICP-MS or

voltammetry. Schmaus et al. reported the validation of a voltammetric method with a

limit of quantification of 40 pg of Pt per sample [10].

In the present study, we describe the development and validation of an ICP-MS method

for the evaluation of surface contamination by Pt originating from cisplatin, oxaliplatin,

and carboplatin. ICP-MS assures an ultra high sensitivity and specificity and requires

relatively simple sample pretreatment procedures. The validated method has been

applied to measure surface contamination in seven Dutch hospital pharmacies.

Experimental

Chemicals

Cisplatin and carboplatin reference standards were purchased from Calbiochem (San

Diego, CA, USA). Oxaliplatin was obtained from Sigma-Aldrich (St. Louis, MO, USA).

Chloroplatinic acid, containing 1,000 mg/L Pt in 3.3% hydrochloric acid (HCl), used for

preparation of calibration solutions, was obtained from Inorganic Ventures/IV Labs

(Lakewood, NJ, USA). Iridium chloride, containing 1,000 mg/L iridium (Ir) in 3.3% HCl,

used for internal standardisation, was also purchased from Inorganic Ventures/IV Labs.

Nitric acid (HNO3) 70% and HCl 35% Ultrex II ultrapure reagents were obtained from

Mallinckrodt Baker (Philipsburg, NJ, USA). Water used for the ICP-MS analysis was sterile

water for irrigation (Aqua B. Braun Medical, Melsungen, Germany). Ethanol 80% was

purchased from Fresenius Kabi (Den Bosch, the Netherlands). A multi-element solution

containing 10 mg/L of Ba, Be, Ce, Co, In, Mg, Pb, Th, and Tl (VAR-TS-MS) was purchased

from Inorganic Ventures/IV Labs. Hoek Loos (Schiedam, The Netherlands) provided

argon gas (4.6) with 99.996% purity.

Instrumentation

Analyses were performed on an ICP-quadrupole-MS (Varian 810-MS) equipped with a

90° reflecting ion mirror (Varian, Mulgrave, Victoria, Australia). The sample introduction

system consisted of a Micromist glass low flow nebuliser (sample uptake 0.4 mL/min), a

peltier-cooled (4 °C) double pass glass spray chamber and a quartz torch. The spray

chamber was cooled to reduce the vapour loading on the plasma, increasing the

available energy for atomisation and ionisation of the elements of interest and to reduce

the formation of solvent based interferences. Sample transport from the SPS-3

autosampler (Varian) to the nebuliser was performed using a peristaltic pump (Watson-

Marlow Alitea, Stockholm, Sweden). The instrument was cooled by using a Kühlmobil

Environmental monitoring

219

142 VD (Van der Heijden, Dörentrup, Germany). Data were acquired and processed using

the ICP-MS Expert Software version 1.1 b49 (Varian). Further data handling was

performed using Excel 2000 (Microsoft, Redmond, WA, USA). All measurements were

carried out in a dedicated temperature-controlled, positively pressurised environment in

order to maintain optimum instrument performance and minimise exogenous

contamination. All solutions were prepared using pipettes (Falcon, Becton Dickinson

Labware, Franklin Lakes, NJ, USA) and polypropylene tubes 10 mL (Plastiques-Gosselin,

Hazebrouck Cedex, France) and 30 mL (Sarstedt AG&Co, Nümbrecht, Germany). Filters

(Minisart) used for filtration of wipe samples were obtained from Sartorius (Hannover,

Germany). Prior to method development, tubes were checked thoroughly for Pt, Ir, and

hafnium (Hf) contamination and appeared to be suitable for Pt analyses.

Determination of Pt by ICP-MS

To optimise the ICP-MS signal for the high masses and to reduce the formation of oxides,

a solution containing 1,000 ng/L of Th, In, Ce, Ba, and Pt was used. Typically this 1,000

ng/L solution gave readings of 115In: 7x105 c/s; 232Th: 1x106 c/s, and 194Pt: 2x105 c/s. The

production of [CeO]+ was less than 1.0% of the total [Ce]+ counts. The formation of

doubly charged [Ba]2+ was less than 3%. Performance was checked daily.

The Pt isotope used for calculation of Pt concentrations was 194Pt. Ir was used as internal

standard. Detection of Pt can be subject to the interference of Hf-oxides [25]. Therefore,

Hf signals were monitored for all samples. The detection mode for all isotopes was based

on peak jumping with peak dwell times of 50 ms, 25 scans per replicate, and three

replicates per sample. Quantification was based on the mean concentration of three

replicates analysed against a calibration curve using weighted linear regression analysis.

Assay development

The most suitable wipe material, extraction solvent, and wipe solvent were selected

using one surface sampling and extraction procedure. This will be described below.

Recovery data were assessed for the three most commonly used Pt agents; cisplatin,

oxaliplatin, and carboplatin. The different molecular structures and the variable physical

characteristics might lead to a variation in absorption and extraction characteristics,

therefore, we decided to evaluate all three compounds instead of choosing one

reference compound.

Chapter 5.1

220

Surface sampling and extraction procedure

Each wipe tissue was moistened with 500 µL wipe solvent. In general, sampling was

performed by wiping a defined surface area of 10x10 cm. However, for surfaces for

which it was not possible to take a 10x10 cm sample, the complete top of the device was

sampled and the area was estimated. All wipe samples were collected with a uniform

sampling procedure by wiping in three different directions (vertical, horizontal, and

diagonal). Wipe samples were stored in 50 mL disposable polypropylene flasks (Falcon,

Becton Dickinson Labware, Franklin Lakes, NJ, USA) at –20 °C until further processing.

Prior to analysis, 10 mL of extraction solvent was added to the sample and flasks were

kept in an ultrasonic bath for 60 min. Then, samples were filtered to remove particles

which could possibly obstruct the ICP-MS nebuliser, or could interfere with the analysis.

Two millilitres of sample were, after addition of Ir as internal standard, introduced

directly into the ICP-MS. Samples of locations which were expected to be highly

contaminated, were diluted prior to analysis to minimise washout memory effects of the

sample introduction system of the ICP-MS.

Wipe material

A variety of wipe tissues are available for collecting samples of surface contaminants

These vary in type of material, surface area, and content of Pt contaminants. Three types

of tissues were evaluated for this study; Kimtech Science precision tissues (Kimberley-

Clark Professional, Irving, TX, USA), Whatman glass fiber filters (Schleicher&Schuell

Microscience GmbH, Dassel, Germany), and Klinion non-woven gauzes (Medeco, Oud-

Beijerland, the Netherlands). The tissues were checked for Pt contamination and for their

ability to release Pt from stainless steel surfaces.

Extraction solvent

One percent HNO3 (v/v), 5% HNO3 (v/v), and 1% HCl (v/v) were evaluated as extraction

solutions. Kimtech Science precision tissues were spiked with cisplatin, oxaliplatin, and

carboplatin and the Pt recovery was determined after extraction with 10 mL of each

solvent.

Wipe solvent

Initially, water, 1% HCl, and 80% ethanol were selected as wiping solutions. To

investigate the capability of these solutions to effectively wipe surfaces, 100 cm2

stainless steel surfaces were spiked with cisplatin, oxaliplatin, and carboplatin. These

Environmental monitoring

221

surfaces were subsequently wiped and the material collected by wiping the surface was

extracted from the wipe using a 1% HCl solution.

Validation procedures

Linearity

For calibration, the chloroplatinic acid reference solution containing 1,000 mg/L Pt was

diluted with 1% HCl to obtain working solutions with concentrations ranging from 50.0

to 5.00x103 ng/L Pt. Working solutions were diluted with 1% HCl to obtain calibration

standards, ranging from 0.500 to 100 ng/L Pt. Before analysis, 15 µL of Ir internal

standard solution was added to 1.5 mL of each calibration standard (final internal

standard concentration 100 ng/L). The seven non-zero calibration standards were

processed and analysed in singular in three separate analytical runs. The calibrations

were back-calculated from the responses. Deviations from the nominal concentration

were evaluated.

Recovery and precision

Quality control (QC) samples were prepared to obtain information on the recovery and

precision of the extraction method and Pt analysis. These samples were analysed in the

validation runs and subsequently also during the analysis of the wipe samples of each

hospital pharmacy. Therefore, stock solutions of the Pt agents in water, each containing

a concentration of drug equivalent to 400 mg/L Pt, were prepared. These stock solutions

were further diluted to obtain spiking solutions with concentrations ranging from 10.0

to 2.00x103 ng/L. Tissues were spiked with these solutions serving as QC samples at the

following concentration levels; 5.00x10-3, 2.5x10-2, 0.100, and 1.00 ng Pt on the tissues,

corresponding to 0.500, 2.50, 10.0, and 100 ng/L Pt in the final solution. These tissues

were processed as described earlier. Three replicates of each sample were analysed in

three analytical runs. Recovery was expressed as a percentage of the nominal

concentration. Within-run and between-run precisions were calculated by analysis of

variances (ANOVA) for each test concentration using the analytical run as the grouping

variable.

Two of the most important surfaces, stainless steel and linoleum, were used to obtain

information on the recovery and within-run and between-run precisions of the complete

sampling procedure, including both wiping and extraction. Therefore, 1.00 ng of

cisplatin, oxaliplatin, or carboplatin was pipetted in triplicate on a 100 cm2 stainless steel

and linoleum surface. After drying overnight, surfaces were wiped following the

previously described procedure and analysed using ICP-MS. Three replicates of each

sample were analysed in three analytical runs for both surfaces. The recovery was

Chapter 5.1

222

expressed as a percentage of the nominal concentration. The within-run and between-

run precision were calculated ANOVA for each test concentration using the analytical

run as the grouping variable.

Limit of quantification

The LLOQ was defined as the concentration at which the analyte response was at least

five times the response of a blank wipe sample. Besides, the LLOQ, when spiked on blank

tissues, had to be determined with a precision less than 20% and the mean value should

not deviate more than 20% of the actual value.

Stability

Stability of cisplatin, oxaliplatin, and carboplatin spiked to tissues, at two concentration

levels, was evaluated at ambient temperatures for one week and under storage

conditions (-20 °C) for up to three weeks. From each storage condition two wipe samples

were analysed. Samples were considered stable when 80-120% of the initial

concentration was recovered.

Pt determination in pharmacy facilities where no cytotoxic agents are processed

Pt is an element that not only appears in the environment due to contamination with Pt

containing cytotoxic drugs, but also due to pollution by car exhaust catalysts. As a

consequence, road dust also contains Pt [26]. Even though, in the pharmacy facilities,

precautions (use of slippers/clogs) are taken to reduce the chance of contamination of

the facility, road dust contamination might occur. Because ICP-MS does not differentiate

between the sources of elemental Pt, this should be taken into account when

considering this technique for evaluation of environmental contamination by cytotoxic

Pt agents. Therefore, two additional locations were included in this study to set a

threshold below which it was not possible to address the source of the contamination.

The first location was the laminar-airflow (LAF) hood in a clean room in which no Pt

contamination was expected. The second location was a preparation unit with LAF hood

of a public pharmacy.

Monitoring of surface contamination in seven Dutch hospital pharmacies

The wipe samples were taken in seven hospital pharmacies in the Netherlands.

Characteristics of the facilities are shown in Table 1. The facilities were selected to

provide a representation of the diversity in hospital pharmacies in the Netherlands in

Environmental monitoring

223

terms of size and amount of Pt compounds handled. The facilities of each hospital

consisted of a preparation room with at least one LAF hood and a room for storage and

checking of the prepared drugs and administration purposes. In each facility, samples

were taken at locations that were prone to contamination.

Table 1. Amount of Pt agents processed and years that the facilities are in service

Site

1 2 3 4 5 6 7

Cisplatin use in 2005 (in g) 52.3 6.80 64.0 147 16.8 29.9 104

Oxaliplatin use in 2005 (in g) 47.5 8.65 44.1 109 23.4 98.8 62.0

Carboplatin use in 2005 (in g) 256 48.2 124 635 56.5 223 217

Total amount of Pt processed in 2005 (in g)

192 34.0 129 483 52.1 185 212

Number of years in service 5 10 18 2.5 1.5 10 15

Wipe sampling frequency Once per year

2005 first time

Sporadic: last in 2004

Once per year

Once per year

Once per year

Twice per year

For good comparison of the results 15 standard locations (Figure 1) were selected: (1)

the middle of the bench-top of the LAF hood, (2) front edge of LAF hood, (3) floor in

front of LAF hood, (4) handle of service hatch, (5) door handle, (6) waste bin top, (7)

bench-top on which materials are placed in storage/checking room, (8) floor in front of

(7), (9) mouse computer, (10) handle telephone, (11) storage shelve of cisplatin, (12)

storage shelve oxaliplatin, (13) storage shelve carboplatin, (14) transport box, (15) handle

refrigerator. Locations 1, 2, and 3 were sampled in duplicate to get an impression of the

overall contamination of these locations. A new pair of gloves was used for each wipe

sample. For each facility three blank samples were prepared by moistening tissue with

500 µL water. All samples were stored and processed as described earlier. The storage

time from sampling until work-up procedure was less than two weeks.

Wipe sampling was announced in each facility in advance and was performed after the

daily cleaning procedure of the LAF hoods, but before the daily cleaning procedure of

the rest of the facility. Wipe sampling in all the facilities was performed by the same

person.

Chapter 5.1

224

Figure 1. Sample locations in pharmacy facilities

Results

Assay development

Wipe material

As a result of high Pt backgrounds (10-20 pg Pt per tissue depending on the batch

analysed), Whatman glass fiber filters were found to be not suitable for Pt wipe

sampling. Kimtech Science precision tissues and Klinion non-woven gauzes did not

show Pt contamination. However, Kimtech Science precision tissues showed better

recoveries of Pt compared to Klinion non-woven gauzes and consequently Kimtech

Science precision tissues appeared to be the best choice.

Extraction solvent

The most effective extraction of Pt from the wipe materials was achieved by 1% HCl (94 -

99%). As a result 1% HCl was selected as the extraction solution of choice.

Environmental monitoring

225

Wipe solvent

Recoveries were inadequate for 80% ethanol (< 40% for all three compounds) and

acceptable for water (50 - 77%) and 1% HCl (63 - 78%). Because 1% HCl appeared to be

corrosive for some types of stainless steel, water was selected as wipe solution.

Validation procedures

Linearity

The calibration curve was best described by linear regression, using 1/(RSD% of a

triplicate sample reading) as weight-factor, to avoid bias in favour of samples with high

standard deviations. Deviations from the nominal concentration were between –10.0

and 10.2% for all concentration levels. Relative standard deviations for the calibration

samples were up to 7.84%. Correlation coefficients were higher than 0.99999.

Recovery and precision

The within-run and between-run precision data for spiked tissues, which served as QCs,

are summarised in Table 2. Precision data showed that, for all QC concentration levels,

the reproducibility of the extraction procedure and Pt analysis was excellent. Recoveries

for cisplatin, oxaliplatin, and carboplatin were between 86.7 and 103% for all

concentration levels. These results indicated sufficient recovery.

Table 2. Within and between-run precision data for quality control samples

Cisplatin Oxaliplatin Carboplatin Amount of Pt spiked to tissue (in ng)

Final Pt concentration (in ng/L) Within-

run (%) Between-run (%)

Within-run (%)

Between-run (%)

Within-run (%)

Between-run (%)

5.00x10-3 0.500 7.75 * 8.53 * 8.01 8.75

2.50x10-2 2.50 4.10 4.05 1.66 1.15 2.39 *

0.100 10.0 1.35 7.44 1.75 1.27 2.86 2.50

1.00 100 1.07 7.79 1.63 * 0.84 1.96 * No statistically significant additional value was observed as a result of performing the assay in different runs (mean square within runs is greater than mean square between runs)

For the recovery and within-run and between-run precision data from the spiked

stainless steel and linoleum surfaces see, respectively, Table 3 and 4. Precision data

showed that the reproducibility of the method, including the wiping procedure was

good. Recoveries from the spiked stainless steel surface were 50.4 % for cisplatin, 73.8%

Chapter 5.1

226

for oxaliplatin, and 77.2% for carboplatin (Table 3). Recoveries for the linoleum surface

were, respectively, 76.8, 77.9, and 81.4% (Table 4).

Table 3. Recovery of 1.00 ng Pt from a stainless steel surface

Cisplatin Oxaliplatin Carboplatin

Mean recovery (%) 50.4 73.8 77.2

Within-run precision (%) 2.21 4.63 2.53

Between-run precision (%) 3.36 * *

Number of days 3 3 3

Number of samples per day 3 3 3 * no statistically significant additional value was observed as a result of performing the assay in different runs (mean square within runs is greater than mean square between runs)

Table 4. Recovery of 1.00 ng Pt from a linoneum surface

Cisplatin Oxaliplatin Carboplatin

Mean recovery (%) 76.8 77.9 81.4

Within-run precision (%) 3.62 2.12 3.35

Between-run precision (%) 12.2 5.41 6.82

Number of days 3 3 3

Number of samples per day 3 3 3

Limit of quantification

The LLOQ of the assay was set at a Pt concentration of 0.5 ng/L in 1% HCl, corresponding

to 5 pg per wipe sample or 0.05 pg/cm2 taking into account a surface of 10x10 cm. Signal

to noise ratios at the LLOQ level exceeded 5 during all the experiments, which was in

accordance with the requirement. The acceptance criteria, that the LLOQ was

determined with a precision less than 20% and that the mean value should deviate no

more than 20% from the actual value, were met for all three compounds (Table 2).

Stability

Sample storage at room temperature for one week was not possible. Tissues which were

spiked with cisplatin showed a decrease in Pt levels of 30% after one week. Oxaliplatin

and carboplatin spiked tissues did not reduce under these conditions. Sample storage at

–20 °C was possible for at least three weeks. Pt concentrations of cisplatin spiked tissues

were decreasing more obvious with time than oxaliplatin and carboplatin spiked tissues.

Environmental monitoring

227

However, no decrease of more than 20% of the initial concentration was observed after

three weeks at –20 °C.

Pt determination in pharmacy facilities where no cytotoxic agents are processed

No Pt was detected in wipe samples from the LAF hood of the public pharmacy. Pt levels

of the LAF hood in the clean room and the floor in the public pharmacy ranged between

0.430 and 0.922 ng/L (or 0.0430 - 0.0922 pg/cm2). Therefore, it was recommended to set

a threshold of 1.00 ng/L Pt (0.100 pg/cm2 when wiping a surface of 100 cm2), below

which it was not possible to address the source of the contamination. All surfaces in the

preparation units with Pt levels above this threshold were considered as being

contaminated by Pt containing drugs.

Monitoring surface contamination in seven Dutch hospital pharmacies

In February 2006, wipe samples were collected from seven Dutch hospital pharmacies

with centralised units dedicated to the preparation of intravenous mixtures of cytotoxic

drugs. The amount of Pt which was processed in these facilities ranged from 34.0 to 483

g per year (Table 1). Surface contamination of all sample locations is depicted in Table 5

in pg/cm2. It is important to consider that recoveries of the samples, as assessed in the

validation study, deviate from 100% dependent on the type of surface sampled and on

the type of compounds present on the surface. Therefore, results depicted in Table 5

represent ≥50.4% of the actual contamination present on the surface.

Pt was detected in 94% of the wipe samples and 88% of the samples contained levels

above the threshold set. Six of the 126 samples showed raised Hf signals which,

considering a maximum oxide formation of 1%, might have accounted for up to 20% of

the Pt signals of these samples. None of blank samples prepared for each facility by

moistening tissues with wipe solvent, contained levels of Pt exceeding 20% of the LLOQ

standard. The variation in the level of contamination between pharmacies was high.

Pharmacies of site 1 and 5 showed overall low Pt contaminations. For these sites,

respectively 33% and 39% did not contain Pt levels above the threshold set. Pt levels

detected at pharmacy 3 were relatively low as well, although the wipes taken from the

floor were high at this site. These high values were, most probably, the result of a

accidental spillages in 2005 with a cisplatin infusion mixture that was dropped on the

floor. Most locations wiped at the hospital pharmacies of site 2, 4, 6, and 7 showed high

contaminations. Only one sample from these sites did not contain any detectable Pt. The

high contamination of site 2 seemed to run counter to the quantities of drugs handled,

because in this pharmacy relatively low amounts of Pt were processed. This site,

however, was occasionally used, for preparation of larger amounts of cytotoxic drugs to

Chapter 5.1

228

serve another hospital. Therefore, the amount of drugs processed in 2005, was not fully

representative for the amount of drugs processed in the ten years that this site was in

use. Site 4 showed the highest contamination, which paralleled the relative amount of

drug handled in this unit.

Table 5. Pt contamination in seven Dutch hospital pharmacies

Pt contamination (in pg/cm2)

Site

Number Sampled surface

1 2 3 4 5 6 7

1 Middle of bench LAF hood 0.22 180 0.54 32.7 0.360 7.22 2.94

Duplicate of 1 0.189 124 0.645 18.7 0.328 8.22 2.28

2 Front edge of LAF hood -a 356 3.32 99.5 0.133 28.2 5.12

Duplicate of 2 -a 268 8.34 180 -a 37.0 5.19

3 Floor in front of LAF hood 3.14 173 824 1107 0.228 2.48 21.7

Duplicate of 3 3.20 232 728 2211 0.186 1.91 12.5

4 Handle of service hatch -a 22.7 -a 2055 -a 1.96 11.8

5 Door handle -a 3.17 -a 16.8 -a 21.4 16.1

6 Waste bin top -a 1.02 0.392 10.1 -a 7.38 0.098c

7 Bench-top on which materials are placed

0.829 0.949 0.298 90.6 0.202 0.375 63.4

8 Floor in front of bench 0.105 c 19.7 58.9 38.1 -a 0.311 c 11.9

9 Mouse computer 0.252 0.816 1.34 10.2 -b 0.758 5.41

10 Handle telephone -a 3.22 0.59 12.1 -a 3.06 5.12

11 Storage shelve cisplatin 0.176 c 1.14 0.157 c 4.76 0.536 4.04 336

12 Storage shelve oxaliplatin 0.368 0.916 0.141 c 4.10 82.7 1.15 2.21

13 Storage shelve carboplatin 3.25 1.53 0.147 3.13 0.186 0.989 5760

14 Transport box 74.5 -a 0.285 4.44 -a 0.828 -b

15 Handle refrigerator 0.452 26.3 1.46 36.0 0.948 1.42 5.71 a Recovered Pt concentrations were below the threshold b Device was not present or available for wipe sampling c Hf oxide might have accounted for up to 20% of the Pt content

Environmental monitoring

229

As expected, Pt was found in most wipe samples taken from the middle of the LAF hood

bench. Notable was that, in general, wipe samples of the front edge of the LAF hood

were more contaminated than samples taken from the bench top of the LAF hood.

Furthermore, floor samples usually contained the highest Pt levels. Other locations

showing substantial contamination were storage shelves, door handles, and handles of

service hatches. Duplicate samples of locations 1, 2, and 3 showed similar results,

indicating a homogeneous distribution over de surface area.

The number of years that the seven units were in use, did not parallel contamination

levels and the amount of drug handled in 2005, overall, did not predict the level of

contamination either.

Discussion

The presence of cytotoxic drug contamination in hospital pharmacies is recognised as a

potential health risk. Therefore, it is important to monitor this contamination. Because Pt

coordination complexes belong to the most extensively used anticancer agents, it is

relevant to focus on the occupational exposure of these drugs. The rationale for

evaluation of Pt contamination is also illustrated by several studies showing increased

levels of Pt in blood [23] and urine [16-18,20,21,23] of hospital personnel working with

these agents.

To be able to accurately assess the Pt contamination originating from cisplatin,

oxaliplatin, and carboplatin at different locations, we developed and validated a wipe

sampling method. ICP-MS was used for quantification of Pt, because this technique

assures a high sensitivity and relative simple sample pretreatment. The sensitivity of the

method was excellent. The LLOQ was set at a Pt concentration of 0.5 ng/L,

corresponding to 5 pg per sample or 0.05 pg/cm2 when wiping a surface of 100 cm2. To

our best knowledge, the method described here is 2-300 times more sensitive than

other methods described for determination of Pt in wipe samples [9-11,13,15,24].

Sample pretreatment only involved surface sampling, extraction, and filtration. After

filtration, samples could be analysed immediately. During method development it was

shown that, in addition to tissue material and extraction solvent, the wipe solvent

affected the recovery to a considerable extent. This was in contrast with results

described by Turci et al, who mentioned that the type of wipe solvent would not

influence the recovery, since contaminants would be swept away from the surfaces

independent of the composition or the pH of the solution itself [27]. Best recoveries

were achieved by wiping with Kimtech Science precision tissues moistened with 500 µL

water and subsequent extraction with 1% HCl.

Validation of the method was performed for the three most prominently used Pt agents

in oncology (cisplatin, oxaliplatin, and carboplatin). We decided not to choose one

Chapter 5.1

230

reference compound, because the different molecular structures of the Pt agents and

associated variable physical characteristics, might lead to a variation in adsorption and

extraction characteristics. Excellent reproducibility (imprecision up to 8.75%) and

recoveries (86.7-103%) were demonstrated with spiked tissues, for all concentration

levels and compounds. Up to 13.3 % of the initial amount of Pt added to the tissues was

not recovered after extraction and analysis. This could be due to variation in analysis, as

well as loss due to adsorption to the tissues. Recoveries from the spiked stainless steel

surface were 50.4 % for cisplatin, 73.8% for oxaliplatin, and 77.2% for carboplatin.

Recoveries, for the three compounds, from the linoleum surface were, respectively, 76.8,

77.9, and 81.4%. These results showed that for stainless steel, depending on the

compound analysed, up to 49.6% of the initial amount of Pt spiked to the surface was

lost. This was, for the greater part, caused by the inability of the wipe procedure to

remove all the added Pt and, for a minor part, by the variation in analysis and loss due to

adsorption to the tissues. The lower recovery that was observed for cisplatin, is, most

probably, a consequence of its superior reactivity compared to oxaliplatin and

carboplatin. This might lead to a stronger binding affinity of cisplatin to materials and

surfaces. For linoleum up to 23.2% of the initial amount of Pt spiked to the surface was

not recovered. For this surface, recoveries were similar for all compounds.

To evaluate the stability of spiked samples, recoveries were assessed after storage at

room temperature and –20 °C. Storage of spiked tissues at –20 °C for at least three weeks

was possible. However, storage of tissues spiked with cisplatin for one week at room

temperature led to considerable decrease in Pt levels. Even though Pt concentrations

from oxaliplatin and carboplatin spiked tissues did not reduce under these conditions,

storage at room temperature was not recommended, also because the source of

elemental Pt is not known in wipe samples performed in pharmacies. Differences in

recovery of the cisplatin and the oxaliplatin and carboplatin spiked samples after

storage, again, could be explained by the higher reactivity of cisplatin.

For a correct interpretation of surface sampling results, it is relevant to take into account

that Pt is an element that not only appears in the environment due to contamination

with Pt containing cytotoxic drugs, but also due to pollution by car exhaust catalysts. By

wipe sampling two locations where no cytotoxic drugs were handled, we, therefore,

determined that below a threshold of 1.00 ng/L Pt (0.100 pg/cm2 when wiping a surface

of 100 cm2), it was not possible to address the source of contamination. All surfaces in

the preparation units with Pt levels above the threshold were considered as being

contaminated by Pt containing drugs.

Taking this threshold into consideration, Pt contamination was reported in 88% of the

samples taken in the seven Dutch hospital pharmacies. It is important to consider that

recoveries of the samples, as assessed in the validation study, deviate from 100%

dependent on the type of surface sampled and on the type of compounds present on

Environmental monitoring

231

the surface. Therefore, results depicted in Table 5 represent ≥50.4 of the actual

contamination present on the surface.

The results of this study indicate that there is substantial variation in surface

contamination of the pharmacies tested and that the amount of Pt processed in the

pharmacies did not always parallel the level of contamination. The number of

preparations with Pt drugs was, however, not assessed and might also be related to

surface contamination. This suggests that variation in the application of or compliance

with cleaning and working procedures and the incidence of calamities, rather than the

amount of Pt processed, caused variation in surface contamination.

In general, results revealed that the LAF hoods, the floor in front of the LAF hoods, door

handles, and handles of service hatches were often contaminated. This demonstrates

that contamination is often spread throughout the pharmacy. Notable was that wipe

samples of the front edge of the LAF hood were more contaminated than samples from

the bench top of the LAF hood. This is thought to be due to incorrect application of

working procedures or insufficient cleaning.

By taking duplicate wipe samples of the LAF hood, the front edge of the LAF hood, and

the floor, we demonstrated that these locations were overall contaminated and that

contamination did not appear to be spotty as was mentioned by Zeedijk et al [6].

We also investigated storage shelves and, in most pharmacies, considerable Pt

contamination was found. These elevated levels could be a consequence of elevated

levels of Pt on packaging material [13-15]. Contamination of packaging can lead to a

spread of cytotoxic drugs to locations where the drugs are stored or processed.

Therefore, it was not surprising that the storage shelves of Pt agents were contaminated.

Furthermore, it was noticed that in at least one of the hospital pharmacies (at site 4),

secondary packaging and caps of the vials were discarded onto the floor during

preparation, most probably to prevent the packaging from interfering with preparation

activities. This pharmacy indeed showed considerable contamination of the floor.

In our study, surfaces showed Pt contamination of up to 5,760 pg/cm2. The results are in

the same range as findings of some other studies describing surface contamination of Pt

in hospital pharmacies [10,12]. In the study of Leboucher et al, however, no Pt was found

outside the LAF hood, which could be due to the high detection limit of the atomic

absorption spectrometry method used (10 µg/L) [12]. Schmaus et al. performed a study

in 14 hospital pharmacies and all samples tested positive for Pt [10], even though the

LLOQ of their method was eight times higher than of the method described here.

Although not all samples tested positive for Pt in our study, the highest contamination

found (5760 pg/cm2) was comparable to the highest contamination found by Schmaus

et al. (2700 pg/cm2). Mason et al. showed lower contamination levels than found in our

study [11], even though the amount of drug handled in these pharmacies was higher

than in the pharmacies in our study.

Chapter 5.1

232

When comparing the amounts of Pt detected on different locations in this study (up to

0.576 µg per wipe sample) with the Pt content of one vial (between 6.50 and 237 mg Pt),

contamination seems to be relatively low. Furthermore, the extremely sensitive

technique used in this study, leads to a high percentage of positive samples.

Interpretation of these results is rather complicated. It is important to consider that the

total area of the contamination is large and that pharmacy personnel are exposed to the

contamination daily. Hence, for safety precautions, it is recommended to attempt to

achieve the lowest possible contamination. Environmental monitoring therefore, may be

used to monitor and control contamination and thereby evaluate working and cleaning

procedures, rather than to interpret potential health risks.

In general, when a minimal contamination level is desired, the results of this study

demonstrate that cleaning and working procedures do not sufficiently prevent

contamination in most hospitals. This could be due to an inadequate compliance of

personnel with these procedures. Moreover, contamination can spread out

unconsciously by hands or feet of the personnel. It is also likely that cleaning procedures

as applied in the different pharmacies are not fully optimised and validated, leading to

contamination due to sub-optimal cleaning. With respect to the physical properties of

the different cytotoxic drugs, it is recommended to consider cleaning techniques

appropriate for specific agents. As was shown in this study for Pt, for instance, 80%

ethanol did not effectively remove Pt from a stainless steel surface. Although water gave

better recoveries for Pt, it was not capable of removing all of the added Pt from the

stainless steel surface. This illustrates the importance to evaluate several cleaning

procedures for the different cytotoxic agents handled and to optimise a procedure

which does remove all drugs with acceptable recoveries.

Conclusion

We developed and validated an ultrasensitive and reliable ICP-MS method for the

determination of Pt in surface samples. This method was successfully applied in the

evaluation of Pt contamination in the preparation units of seven Dutch hospital

pharmacies. It was demonstrated that pharmacy personnel is at risk to be exposed to Pt,

despite the use of cleaning and safety procedures. As long as the consequences of long-

term exposure are not known, the aim should be to achieve a contamination levels as

low as possible. This study, therefore, highlights the need to further evaluate cleaning

and safety procedures. Wipe sampling can be applied to quantify improvements made

through changes in procedures.

Environmental monitoring

233

References 1. Donner AL. Possible risk of working with antineoplastic drugs in horizontal laminar flow hoods. Am J

Hosp Pharm 1978; 35: 900.

2. Falck K, Grohn P, Sorsa M, Vainio H, Heinonen E, Holsti LR. Mutagenicity in urine of nurses handling cytostatic drugs. Lancet 1979; 1: 1250-1.

3. International Agency for Research on Cancer (IARC). Monographs on the evaluation of the carcinogenic risks of chemicals to humans: overall evaluations of carcinogenicity. Updation of IARC monographs Vols 1-42. Lyon, France. http://www iarc fr 1997.

4. Sessink PJ, Van de Kerkhof MC, Anzion RB, Noordhoek J, Bos RP. Environmental contamination and assessment of exposure to antineoplastic agents by determination of cyclophosphamide in urine of exposed pharmacy technicians: is skin absorption an important exposure route? Arch Environ Health 1994; 49: 165-9.

5. Fransman W, Vermeulen R, Kromhout H. Occupational dermal exposure to cyclophosphamide in Dutch hospitals: a pilot study. Ann Occup Hyg 2004; 48: 237-44.

6. Zeedijk M, Greijdanus B, Steenstra FB, Uges DRA. Monitoring of cytostatics on the hospital ward: Measuring surface contamination of four different cytostatic drugs from one wipe sample. The European Journal of Hospital Pharmacy Science 2005; 11: 18-22.

7. Hedmer M, Georgiadi A, Bremberg ER, Jonsson BA, Eksborg S. Surface contamination of cyclophosphamide packaging and surface contamination with antineoplastic drugs in a hospital pharmacy in Sweden. Ann Occup Hyg 2005; 49: 629-37.

8. Crauste-Manciet S, Sessink PJ, Ferrari S, Jomier JY, Brossard D. Environmental contamination with cytotoxic drugs in healthcare using positive air pressure isolators. Ann Occup Hyg 2005; 49: 619-28.

9. Ziegler E, Mason HJ, Baxter PJ. Occupational exposure to cytotoxic drugs in two UK oncology wards. Occup Environ Med 2002; 59: 608-12.

10. Schmaus G, Schierl R, Funck S. Monitoring surface contamination by antineoplastic drugs using gas chromatography-mass spectrometry and voltammetry. Am J Health Syst Pharm 2002; 59: 956-61.

11. Mason HJ, Blair S, Sams C, Jones K, Garfitt SJ, Cuschieri MJ, Baxter PJ. Exposure to antineoplastic drugs in two UK hospital pharmacy units. Ann Occup Hyg 2005; 49: 603-10.

12. Leboucher G, Serratrice F, Bertholle V, Thore L, Bost M. [Evaluation of platinum contamination of a hazardous drug preparation area in a hospital pharmacy]. Bull Cancer 2002; 89: 949-55.

13. Connor TH, Sessink PJ, Harrison BR, Pretty JR, Peters BG, Alfaro RM, Bilos A, Beckmann G, Bing MR, Anderson LM, Dechristoforo R. Surface contamination of chemotherapy drug vials and evaluation of new vial-cleaning techniques: results of three studies. Am J Health Syst Pharm 2005; 62: 475-84.

14. Nygren O, Gustavsson B, Strom L, Friberg A. Cisplatin contamination observed on the outside of drug vials. Ann Occup Hyg 2002; 46: 555-7.

15. Mason HJ, Morton J, Garfitt SJ, Iqbal S, Jones K. Cytotoxic drug contamination on the outside of vials delivered to a hospital pharmacy. Ann Occup Hyg 2003; 47: 681-5.

16. Schreiber C, Radon K, Pethran A, Schierl R, Hauff K, Grimm CH, Boos KS, Nowak D. Uptake of antineoplastic agents in pharmacy personnel. Part II: study of work-related risk factors. Int Arch Occup Environ Health 2003; 76: 11-6.

17. Pethran A, Schierl R, Hauff K, Grimm CH, Boos KS, Nowak D. Uptake of antineoplastic agents in pharmacy and hospital personnel. Part I: monitoring of urinary concentrations. Int Arch Occup Environ Health 2003; 76: 5-10.

18. Turci R, Sottani C, Ronchi A, Minoia C. Biological monitoring of hospital personnel occupationally exposed to antineoplastic agents. Toxicol Lett 2002; 134: 57-64.

19. Minoia C, Turci R, Sottani C, Schiavi A, Perbellini L, Angeleri S, Draicchio F, Apostoli P. Application of high performance liquid chromatography/tandem mass spectrometry in the environmental and biological monitoring of health care personnel occupationally exposed to cyclophosphamide and ifosfamide. Rapid Commun Mass Spectrom 1998; 12: 1485-93.

20. Ensslin AS, Huber R, Pethran A, Rommelt H, Schierl R, Kulka U, Fruhmann G. Biological monitoring of hospital pharmacy personnel occupationally exposed to cytostatic drugs: urinary excretion and cytogenetics studies. Int Arch Occup Environ Health 1997; 70: 205-8.

21. Ensslin AS, Pethran A, Schierl R, Fruhmann G. Urinary platinum in hospital personnel occupationally exposed to platinum-containing antineoplastic drugs. Int Arch Occup Environ Health 1994; 65: 339-42.

Chapter 5.1

234

22. Ensslin AS, Stoll Y, Pethran A, Pfaller A, Rommelt H, Fruhmann G. Biological monitoring of cyclophosphamide and ifosfamide in urine of hospital personnel occupationally exposed to cytostatic drugs. Occup Environ Med 1994; 51: 229-33.

23. Nygren O, Lundgren C. Determination of platinum in workroom air and in blood and urine from nursing staff attending patients receiving cisplatin chemotherapy. Int Arch Occup Environ Health 1997; 70: 209-14.

24. Raghavan R, Burchett M, Loffredo D, Mulligan JA. Low-level (PPB) determination of cisplatin in cleaning validation (rinse water) samples. II. A high-performance liquid chromatographic method. Drug Dev Ind Pharm 2000; 26: 429-40.

25. Lustig L, Zang S, Michalke B, Schramel P, Beck W. Platinum determination in nutrient plants by inductively coupled plasma mass spectrometry with special respect to the hafnium oxide interference. Fresenius J Anal Chem 1997; 357: 1157-63.

26. Barefoot RR. Determination of platinum at trace levels in environmental and biological samples. Environmental Science & Technology 1997; 31: 309-14.

27. Turci R, Sottani C, Spagnoli G, Minoia C. Biological and environmental monitoring of hospital personnel exposed to antineoplastic agents: a review of analytical methods. J Chromatogr B Analyt Technol Biomed Life Sci 2003; 789: 169-209.

Conclusions and perspectives

238

Conclusions and perspectives

The introduction of inductively coupled plasma mass spectrometry (ICP-MS) in clinical

pharmacological oncology research resulted in new opportunities in the field of

quantitative analysis of metal-based anticancer agents. The technique is highly sensitive

and can be used to study heavy metal levels in a wide range of sample matrices from

biological and environmental origin.

In this thesis, we described the development and validation of methods for the analysis

of platinum (Pt) and ruthenium (Ru) in various biological matrices. The assays were

optimised and validated according to the latest FDA recommendations for bioanalytical

method validation. ICP-MS proved to be applicable for the determination of metal-based

anticancer agents in biological fluids. Furthermore, platinum-DNA (Pt-DNA) adduct

levels could be assessed in peripheral blood mononuclear cells (PBMCs) and tissue. For

the latter, only 1 mg of tissue was needed, which demonstrates the extreme low

detection capability of the method.

The use of ICP-MS greatly enhances the pharmacokinetic window that can be evaluated.

The availability of the sensitive assays described in this thesis offer the opportunity to

investigate research questions concerning the pharmacokinetics, mechanism of action,

efficacy, and safety of metal-based anticancer agents. Furthermore, the assays can be

used to evaluate the efficacy of drug targeting. In addition to pharmacological

applications, ICP-MS can be used to evaluate environmental contamination with heavy

metals.

Investigation of pharmacokinetics of metal-based anticancer agents

Before the introduction of ICP-MS it was only possible to study unbound Pt

pharmacokinetics during or shortly after treatment. The presence of persistent side

effects, however, has led to an increased demand for the investigation of long-term

pharmacokinetics, distribution, and elimination of metal-based drugs. It was frequently

hypothesized that persistent side effects could be a consequence of heavy metal which

was retained in the body after treatment. In the current thesis we investigated the long-

term pharmacokinetics of Pt in patients treated with cisplatin and oxaliplatin (Chapter

4.1). We observed Pt levels in plasma and plasma ultrafiltrate (pUF) to be significantly

elevated up to 75 months after the end of treatment. Remaining Pt levels decreased with

time. Although no Pt-DNA adducts could be detected in PBMCs, it was shown that Pt

species in pUF were still present in a reactive form. It was hypothesized that persistent

side effects might be a consequence of Pt that accumulates in the body and remains

bound to e.g. proteins and DNA. In the investigations described in this thesis, however,

we were not able to establish relationships between plasma Pt levels and persistent Pt

Conclusions and perspectives

239

induced neuropathy (Chapter 4.2). Hence, the clinical consequences of increased post-

treatment Pt levels remain to be established.

Investigation of efficacy and safety of metal-based anticancer agents

The ability of ICP-MS to investigate metal pharmacokinetics in a small volume of

biological fluid and tissue offers the opportunity to perform preclinical longitudinal

efficacy and safety studies in rodents. Due to the low amount of sample needed, animals

need no longer be sacrificed prior to analysis. This, in turn, enables intervention studies

in rodents. The significant side effects induced by metal-based anticancer agents can

possibly be prevented or reduced by the administration of antidotes. Investigations in

the current thesis revealed that sodium thiosulfate (STS) is capable of reducing

neuropathy and Pt accumulation in the body when administered during cisplatin

treatment (Chapter 4). It is plausible that sulfur-containing compounds can also reduce

persistent side effects when administered months after treatment. To realise such

effects, however, the compounds should be capable of removing Pt from both the

cellular and blood compartment. We investigated the effect of STS and other sulfur-

containing compounds on Pt-DNA and Pt-protein binding ex vivo (Chapter 3.2 and 4.1).

The in vivo effects of these and other antidotes can well be investigated in intervention

studies in rodents.

Investigation of mechanism of action of metal-based anticancer agents

Up to now, the mechanism of action of metal-containing anticancer agents is not fully

understood. There still is considerable interest in the investigation of the mechanism of

action of metal-based anticancer agents. Hyphenation of ICP-MS with HPLC creates the

possibility to further investigate the mechanism of action of metal-based anticancer

agents, as ICP-MS can serve as a very sensitive metal selective detector following

separation using e.g. high performance liquid chromatography (HPLC). In addition to the

exploration of the metabolism of metal-containing anticancer agents, the interaction of

the parent compounds and metabolites with endogenous species can be investigated.

Besides the investigation of the different Pt-DNA adducts formed (Chapter 3.2), future

research will concentrate on the separation and investigation of Pt- and Ru-protein

complexes.

Role in drug targeting

To decrease the side effects and increase the toxic effects to the tumour, drug targeting

is an increasingly applied tool. By modifying the pharmaceutical characteristics of an

240

effective anticancer drug, investigators aim to increase the amount of drug which is

directed to the tumour and decrease the side effects.

The potential of ICP-MS to determine metal levels in small biopsy samples such as fine

needle aspirates (Chapter 3.1), creates the opportunity to trace the pharmaceutically

modified drug and to assess tumour metal levels in rodents and patients. Thereby, the

most adequate modification can be selected, leading to optimisation of treatment

schedules and minimisation of side effects.

Environmental monitoring

In addition to the analysis of clinical samples, a method for the analysis of surface

samples to assess contamination of environments where Pt drugs are processed

(Chapter 5) was developed. This method was applied to evaluate surface contamination

in seven hospital pharmacies. In general, the results of this study revealed considerable

contamination, demonstrating that cleaning and working procedures did not sufficiently

prevent contamination in most hospitals. In future studies, cleaning and working

procedures can be optimised using ICP-MS to test the efficiency of the procedures.

Unfortunately, up to now, the consequences of long-term environmental exposure to Pt-

based anticancer agents are not known. It would be useful to assess a contamination

level below which no harmful effects are expected. This can be done by evaluating the

DNA-binding reactivity and the ability to inhibit cell growth of Pt recovered from the

surface samples. Furthermore, it would be interesting to measure Pt in plasma, urine,

and Pt-DNA adducts in PBMCs of hospital personnel to assess whether the Pt

contamination is actually absorbed into the body and whether it is still in the active

form.

In addition to the environmental monitoring of hospital pharmacies, ICP-MS can be

applied in several areas to evaluate heavy metal contamination. Not only samples from

hospitals and (pharmaceutical) industry, but also surface water, road dust, and food are

relevant sources to investigate for contamination with heavy metals such as Pt, arsenic,

cadmium, and mercury. Furthermore, the uptake of heavy metals by human as a

consequence of environmental exposure is of interest. Not only elevated plasma and

urine levels, but also increased tissue and breast milk levels need attention.

In conclusion, the successful application of ICP-MS in oncology has had an enormous

impact on the field of quantitative analysis of metal-based anticancer agents from

biological and environmental samples. The technique offers the opportunity to study

fairly all applications of metal-based anticancer agents, which could be of interest in

oncology.

Summary Samenvatting

244

Summary

After the discovery of the antiproliferative effects of cisplatin, the drug has developed into one of the most frequently used anticancer agents. Unfortunately, the use of cisplatin is hampered by severe side effects and by the resistance of several tumour types. These limitations have led to the development and evaluation of thousands of metal-containing compounds of which only a few have entered clinical trials. Nowadays, the platinum-containing complexes oxaliplatin and carboplatin have found important applications, whereas satraplatin is under consideration for approval. In addition, ruthenium complexes are regarded as promising alternatives for platinum complexes. Research to unravel the pharmacokinetics and –dynamics of metal-based anticancer agents is required to understand the clinical behaviour of the drugs and to further optimise treatment regimens. Accurate and sensitive methods for the quantitative determination of metal-based anticancer agents are indispensable to study these aspects. Therefore, the introduction of inductively coupled plasma mass spectrometry (ICP-MS) in clinical pharmacological oncology research resulted in new opportunities in the field of quantitative analysis of metal-based anticancer agents. This technique is highly sensitive and can be used to study heavy metals in a wide range of sample matrices. The aim of this thesis (Chapter 1.1) was to develop and validate analytical ICP-MS methods for the analysis of metal-based anticancer agents. These methods were applied to answer research questions concerning long-term pharmacokinetics, platinum-induced side effects, the effects of antidotes on platinum-induced side effects, and environmental monitoring. ICP-MS has now become the method of first choice for the quantitative bioanalysis of metal-based anticancer agents as is demonstrated by the large number of publications that have appeared on the subject so far. ICP-MS provides an extremely high sensitivity. Therefore, in addition to the investigation of pharmacokinetics during or shortly after chemotherapy, other research questions, which are of current interest, can be answered. In Chapter 1.2, an overview of the literature available in this field is presented. The focus is on the determination of the total metal concentration in biological samples (plasma, urine, tissue, DNA and protein adducts) and in environmental samples. Furthermore, the speciation of platinum and ruthenium compounds is described. Chapter 2 describes the development and validation of assays for the analysis of

platinum in plasma and plasma ultrafiltrate (pUF) using atomic absorption spectrometry

(AAS) (Chapter 2.1), of platinum in pUF using ICP-MS (Chapter 2.2), and of ruthenium in

plasma, pUF, and urine by ICP-MS (Chapter 2.3). The benefit of the use of ICP-MS

compared to AAS for the analysis of metal-based anticancer agents in biological fluids

Summary

245

was illustrated. The techniques appeared to be in good agreement. Using ICP-MS,

however, the quantification limit for the analysis of platinum in pUF was 2600-fold lower

than for AAS. For ruthenium, quantification limits for ICP-MS appeared to be 740- (pUF),

1,500- (plasma), and 3,700- (urine) fold lower than for AAS.

Chapter 3 describes the application of ICP-MS for the determination of platinum adducts.

In Chapter 3.1, the development, optimisation, and validation of an ICP-MS method for

the determination of platinum bound to DNA in peripheral blood mononuclear cells

(PBMCS) and tissue was described. The method proved to be applicable for the

determination of platinum-DNA adducts in PBMCs isolated from 10 mL of blood and in 1

mg of tissue. The possibility to analyse platinum-DNA adducts in extremely small tissue

samples creates the opportunity to apply the method to study the levels of adducts in

biopsy samples from e.g. fine needle aspirates and to investigate the distribution of

adducts across the tumour. The current method was applied to study platinum-DNA

adduct levels in PBMCs and tissue from patients treated with cisplatin. To evaluate

pharmacodynamics, platinum-DNA adduct levels in PBMCs are frequently used as a

surrogate marker for the level of platinum-DNA adducts in tumours, because tumour

biopsy samples are often hard to obtain. It is, however, not always feasible to assume

that pharmacodynamics in tissue are in agreement with the dynamics in PBMCs. To

address this issue in patients with gastric cancer, we studied platinum-DNA adduct levels

in gastric tissue and PBMCs 24 h after treatment. Although the number of patients (3)

evaluated was limited and no definite conclusions could be drawn, tissue platinum-DNA

adduct levels were significantly higher than levels in PBMCs. Further research is needed

to evaluate whether there is a correlation between tissue and PBMC adduct levels.

In Chapter 3.2, the effects of gemcitabine and the sulfur-containing compounds sodium

thiosulfate (STS), glutathione (GSH), and acetylcysteine (AC) on platinum-protein and

platinum-DNA adduct levels were quantified in order to investigate the potential of

these compounds to modify platinum cytotoxicity. Gemcitabine seemed to slightly

inhibit the platinum-DNA binding reactivity. STS, GSH, and AC were capable to (partly)

prevent the platinum-protein and -DNA binding. Furthermore, STS and AC appeared to

be able to reverse platinum-protein binding. They, however, could not obviously release

platinum from the DNA.

Chapter 4 describes the long-term effects of cisplatin and oxaliplatin treatment. Long-

term platinum pharmacokinetics were investigated by the analysis of platinum in

biological fluids and platinum bound to DNA in PBMCs in patients treated with cisplatin

or oxaliplatin 8 to 75 months before inclusion in the study (Chapter 4.1). It was observed

that platinum levels in plasma and pUF were still significantly elevated in all 45 patients.

Investigations of the relationships between several determinants and platinum levels

246

revealed that remaining plasma platinum levels were related to follow-up time, age,

cumulative dose, glomerular filtration rate at time of treatment, and STS use.

To evaluate whether the remaining platinum was still reactive, the platinum-DNA and

platinum-protein binding characteristics of the platinum from the patients’ samples

were quantified. Although no platinum-DNA adducts could be detected in PBMCs, it was

shown that platinum species in pUF were still present in a reactive form, because

platinum species in pUF could still bind to DNA and proteins. The remaining DNA

binding capacity, however, was only up to 10% of the binding capacity of the parent

compounds cisplatin and oxaliplatin.

It was hypothesized that persistent side effects might be a consequence of platinum

which is accumulated in the body and remains bound to e.g. proteins and DNA. In the

investigations described in this thesis, however, we were not able to establish

relationships between platinum levels and persistent platinum-induced neuropathy

(Chapter 4.2). This could be due to the fact that only one data point per patient was

available. To evaluate relationships between platinum levels and side effects, more data

points per patient are indispensable. The clinical consequences of the increased

platinum levels, therefore, remain to be established.

Because, the investigations revealed that STS is capable of reducing neuropathy and

platinum accumulation in the body when administered during cisplatin treatment

(Chapter 4), it is plausible that sulfur-containing compounds may also reduce persistent

side effects when administered months after treatment. To realise such effects, however,

the compounds should be capable of removing platinum from the cellular and blood

compartment. Although STS and AC were capable of preventing platinum-protein and -

DNA binding and reversing platinum-protein binding, they could hardly release

platinum from the DNA (Chapter 3.2). It is, therefore, not reasonable that these antidotes

might reduce persistent side effects when administered after treatment.

Besides the analysis of clinical samples, we developed a method for the analysis of

surface samples to assess contamination of environments where platinum drugs are

processed (Chapter 5). This method was applied to evaluate surface contamination in

seven hospital pharmacies. Results showed that the contamination was variable and was

spread through the preparation rooms, which demonstrates that cleaning and working

procedures do not sufficiently prevent contamination in most hospitals. Unfortunately,

up to now, the consequences of long-term platinum exposure are not known. Therefore,

the aim should be to achieve a contamination and exposure levels as low as possible.

In conclusion, the successful application of ICP-MS in oncology has had an enormous

impact on the field of quantitative analysis of metal-based anticancer agents from

biological and environmental samples. The use of this ultrasensitive technique has

Summary

247

contributed to a better insight into long-term platinum pharmacokinetics, the effect of

antidotes on the pharmacokinetics and -dynamics, and the level of environmental

contamination. Furthermore, using ICP-MS, the investigation of pharmacokinetics and –

dynamics is simplified because only a small sample volume is required. This decreases

the inconvenience to the patients. The assays described in the current thesis have paved

the way for further investigation of research questions concerning pharmacokinetics,

mechanism of action, efficacy, and safety of metal-based anticancer agents. In addition

to pharmacological applications, issues regarding environmental monitoring can be

explored.

248

Samenvatting

Na de ontdekking van de remmende werking van cisplatine op de groei van tumoren, heeft het middel zich ontwikkeld tot een van de meest gebruikte antikankermiddelen. Helaas gaat het gebruik van cisplatine vaak gepaard met ernstige bijwerkingen en zijn verschillende type tumoren resistent tegen het middel. Deze beperkingen hebben er toe geleid dat duizenden metaal bevattende verbindingen zijn ontwikkeld en getest. Van deze verbindingen bleken er slechts enkele geschikt om te worden geïntroduceerd in de kliniek. Tegenwoordig worden naast cisplatine ook de platina-bevattende middelen oxaliplatine en carboplatine in de kliniek toegepast. Een ander middel, satraplatine, wordt momenteel onderzocht op klinische toepasbaarheid. Naast platina-bevattende middelen worden ook ruthenium complexen als mogelijke alternatieven ontwikkeld. Om het klinisch gedrag van metaal-bevattende antikankermiddelen te kunnen begrijpen, is onderzoek naar de farmacokinetiek en –dynamiek vereist. Hierdoor wordt verdere optimalisatie van de behandeling met deze geneesmiddelen mogelijk. Om de farmacokinetiek en –dynamiek te onderzoeken is het van belang dat de hoeveelheid van de metaal-bevattende antikankermiddelen in het lichaam wordt gemeten. Inductief-gekoppelde plasma-massa-spectrometrie (ICP-MS) is een zeer gevoelige methode voor de kwantitatieve bepaling van deze middelen. De doelstelling van het onderzoek in dit proefschrift (Hoofdstuk 1.1) was om kwantitatieve ICP-MS-methoden voor metaal-bevattende antikankermiddelen te ontwikkelen en te valideren. Deze methoden zijn vervolgens toegepast om onderzoeksvragen te beantwoorden met betrekking tot de bijwerkingen en lange-termijn farmacokinetiek van deze middelen. De methoden zijn ook gebruikt voor omgevingsmonitoring en voor onderzoek naar effecten van antidota op platina-geïnduceerde bijwerkingen. ICP-MS is nu de techniek van eerste keuze voor de kwantitatieve bioanalyse van metaal-bevattende antikankermiddelen. De methode is zeer gevoelig en kan daardoor gebruikt worden voor vele klinisch farmacologische onderzoeksvragen. Dit blijkt uit het grote aantal publicaties dat op dit gebied is verschenen. In Hoofdstuk 1.2 wordt een overzicht van deze literatuur gepresenteerd. Het hoofdstuk richt zich op de bepaling van de totale hoeveelheid platina en ruthenium in humane biologische monsters (plasma, urine, weefsel en DNA- en eiwitadducten) en omgevingsmonsters. Daarnaast wordt de combinatie van scheidingstechnieken met ICP-MS beschreven. Ontledings- en reactieproducten kunnen van elkaar gescheiden worden en de hoeveelheid metaal in de gescheiden producten kan vervolgens met ICP-MS worden bepaald. Hoofdstuk 2 beschrijft de ontwikkeling en validatie van bepalingsmethoden voor de bepaling van platina in plasma en plasma-ultrafiltraat door middel van atomaire

Samenvatting

249

absorptiespectrometrie (AAS) (Hoofdstuk 2.1), van platina in plasma-ultrafiltraat met behulp van ICP-MS (Hoofdstuk 2.2) en van ruthenium in plasma, plasma-ultrafiltraat en urine met ICP-MS (Hoofdstuk 2.3). Met de ICP-MS methoden werden vele lagere bepalingslimieten bereikt dan met AAS. Voor de bepaling van platina in plasma-ultrafiltraat werd met ICP-MS een bepalingslimiet bereikt die 2600 maal lager is dan met AAS. Voor de bepaling van ruthenium in plasma-ultrafiltraat, plasma en urine waren de bepalingslimieten behaald met ICP-MS respectievelijk zelfs 740, 1500 en 3700 keer lager dan gebruikelijk bij AAS. Door de lage bepalingslimieten kan zowel de farmacokinetiek tijdens de behandeling worden onderzocht, als ook de farmacokinetiek tot jaren na het stoppen van de behandeling. Hoofdstuk 3 beschrijft de toepassing van ICP-MS voor de bepaling van platina-adducten. In Hoofdstuk 3.1 wordt de ontwikkeling, optimalisatie en validatie van een ICP-MS methode beschreven voor de bepaling van de hoeveelheid platina gebonden aan DNA in witte bloedcellen en weefsel. De methode bleek toepasbaar voor de bepaling van platina-DNA-adducten in witte bloedcellen geïsoleerd uit 10 mL bloed en van platina-DNA-adducten in slechts 1 mg weefsel. De mogelijkheid deze adducten te bepalen in een zeer geringe hoeveelheid weefsel maakt ook mogelijk om platina gebonden aan DNA te meten in monsters uit bijvoorbeeld een fijne naald biopsie. Hierdoor kan onder andere de verdeling van adducten over de tumor onderzocht worden. De in Hoofdstuk 3.1 beschreven methode is gebruikt voor het bepalen van concentraties van platina gebonden aan DNA in witte bloedcellen en weefsel van patiënten die behandeld werden met cisplatine. Voor het evalueren van de farmacodynamiek van een platina-bevattend middel worden de hoeveelheden platina-DNA-adducten in witte bloedcellen vaak als surrogaatmerker gebruikt voor de hoeveelheden platina-DNA-adducten in tumoren. Dit wordt gedaan omdat tumorweefsel doorgaans moeilijk te verkrijgen is. Het is echter niet vanzelfsprekend dat de farmacodynamiek in weefsel overeenkomt met die in witte bloedcellen. Om dit te onderzoeken werd de hoeveelheid platina gebonden aan DNA in weefsel uit de maag en uit witte bloedcellen van patiënten met maagkanker bepaald 24 uur na de toediening van cisplatine. Hoewel het aantal onderzochte patiënten (3) gering was, waardoor nog geen definitieve conclusies getrokken konden worden, waren de platina-DNA adductconcentraties in weefsel significant hoger dan de concentraties in witte bloedcellen. Verder onderzoek is vereist om na te gaan of er een correlatie bestaat tussen weefsel en wittebloedcel adductconcentraties. In Hoofdstuk 3.2 is het effect van gemcitabine en de zwavel-bevattende middelen natriumthiosulfaat (STS), glutathion (GSH) en acetylcysteïne (AC) op de platina-eiwit- en platina-DNA-binding gekwantificeerd. Dit onderzoek werd verricht om vast te stellen of deze middelen toegepast zouden kunnen worden voor het modificeren van de cytotoxiciteit van platina-bevattende middelen. Gemcitabine remde de platina-DNA-bindingsreactiviteit in geringe mate. STS, GSH en AC voorkwamen, deels, de binding van platina aan eiwit en DNA. Verder bleken STS en AC in staat om platina-gebonden eiwit vrij te maken. Platina werd echter niet van het DNA vrijgemaakt door deze middelen.

250

Hoofdstuk 4 beschrijft het lange-termijneffect van behandeling met cisplatine en oxaliplatine. De lange-termijnfarmacokinetiek van platina werd onderzocht door de meting van platina in plasma, plasma-ultrafiltraat en gebonden aan DNA in witte bloedcellen in patiënten die 8 tot 75 maanden eerder behandeld waren met cisplatine of oxaliplatine. (Hoofdstuk 4.1). Uit het onderzoek bleek dat platina concentraties in plasma en plasma-ultrafiltraat in alle 45 patiënten significant verhoogd waren ten opzichte van controle patiënten. Uit het onderzoek bleek eveneens dat de platinaconcentraties in plasma gerelateerd zijn aan follow-uptijd, leeftijd, cumulatieve dosis, glomerulaire filtratiesnelheid ten tijde van de behandeling en gebruik van STS. Om te bepalen of het resterende platina nog reactief was, zijn de bindingskarakteristieken van het platina aan eiwit en DNA bepaald in patiëntenmonsters. Hoewel er geen platina-DNA-adducten in witte bloedcellen werden gedetecteerd, bleek het platina in plasma-ultrafiltraat nog steeds reactief te zijn. Platina species aanwezig in plasma-ultrafiltraat waren nog steeds in staat om aan DNA en eiwitten te binden. De resterende DNA-bindingscapaciteit was echter maximaal 10% van de bindingscapaciteit van de intacte verbindingen cisplatine en oxaliplatine. Het zou kunnen dat de persistente bijwerkingen van platina een gevolg zijn van het platina dat in het lichaam opgehoopt is en gebonden is aan bijvoorbeeld eiwitten en DNA. In het onderzoek dat beschreven is in dit proefschrift kon echter geen relatie vastgesteld worden tussen platinaconcentraties en persistente, door platina geïnduceerde, neuropathie (Hoofdstuk 4.2). De reden hiervoor zou kunnen zijn dat er enkel één datapunt per patiënt beschikbaar was en dat voor het bepalen van de relatie tussen platinaconcentraties en bijwerkingen meerdere meetpunten noodzakelijk zijn. De klinische consequenties van de verhoogde platinaconcentraties moet daarom nog vastgesteld worden. Onderzoek liet zien dat STS, wanneer toegediend gedurende cisplatine behandeling, platina-geïnduceerde neuropathie reduceert en platina-ophoping in het lichaam vermindert (Hoofdstuk 4). De vraag is of zwavel-bevattende middelen, wanneer ze maanden na behandeling worden gegeven, ook persistente bijwerkingen kunnen reduceren. Om dergelijke effecten te realiseren moeten de middelen in staat zijn om platina uit de cellulaire- en bloedcompartimenten te verwijderen. Hoewel STS en AC de platina-eiwit- en platina-DNA-binding konden voorkomen en tevens platina van eiwit los konden maken, waren ze niet in staat om platina van het DNA te verwijderen (Hoofdstuk 3.2). Daarom is het niet aannemelijk dat deze antidota de bijwerkingen kunnen reduceren wanneer ze na afloop van de behandeling worden toegediend. Naast de bepaling van metaal in klinische monsters, werd een methode ontwikkeld om oppervlaktemonsters (veegproeven) te analyseren. Het doel hiervan was de verontreiniging vast te stellen in de omgeving waar platinamiddelen toedieningsgereed worden gemaakt (Hoofdstuk 5). De methode is gebruikt voor het bepalen van oppervlakte verontreiniging in zeven Nederlandse ziekenhuisapotheken. De resultaten

Samenvatting

251

lieten zien dat de verontreiniging variabel was en dat deze verspreid voorkwam in de bereidingsruimtes. Deze resultaten duiden erop dat de schoonmaak- en werkprocedures in de meeste ziekenhuizen verontreiniging niet in voldoende mate voorkomen. Helaas zijn de gevolgen van lange-termijnblootstelling aan platina tot nu toe onbekend. Daarom wordt geadviseerd om iedere verontreiniging en blootstelling te voorkomen. Concluderend kan gesteld worden dat de succesvolle toepassing van ICP-MS in de oncologie een grote invloed heeft gehad op de kwantitatieve analyse van metaal-bevattende antikankermiddelen. Het gebruik van deze techniek heeft bijgedragen aan een beter inzicht in de lange-termijnfarmacokinetiek, het effect van antidota op de farmacokinetiek en -dynamiek en het niveau van omgevingsverontreiniging. Verder heeft het gebruik van ICP-MS het onderzoeken van de farmacokinetiek en -dynamiek vereenvoudigd omdat slechts een monster van zeer klein volume nodig is. Dit zorgt voor een verminderde belasting van de patiënten. De methoden die beschreven zijn in dit proefschrift hebben het pad geëffend voor verder onderzoek naar de farmacokinetiek, werkingsmechanismen, effectiviteit en veiligheid van metaal-bevattende antikankermiddelen. Naast klinisch-farmacologische toepassingen kunnen kwesties met betrekking tot omgevingsmonitoring worden onderzocht.

Dankwoord Curriculum Vitae

List of Publications

254

Dankwoord

Ben jij één van die lezers die bij het openen van het proefschrift als eerste het

dankwoord leest? Ik doe dat ook altijd. En nu ik zelf op het punt sta een dankwoord te

schrijven vraag ik me af waarom? Het dankwoord staat immers bijna achteraan en wordt

voorafgegaan door pagina’s interessant onderzoek. Misschien omdat het dankwoord

één van de weinige plekken in het proefschrift is waar je iets persoonlijks kunt laten zien

en wellicht is dat waar je naar op zoek bent. Wie is de persoon die het boek tot stand

heeft gebracht en, niet minder belangrijk belangrijk, wie hebben hem of haar daarbij

ondersteund? Het dankwoord laat zien hoe de promovendus de promotie heeft beleefd.

En dat is waar ik nieuwsgierig naar ben als ik een proefschrift van een andere OIO krijg.

Laat ik dus maar meteen antwoord geven op de vraag: hoe heb ik mijn promotie

eigenlijk beleefd? Ik vond het in één woord geweldig! En als ik eerlijk ben, vind ik het

jammer dat het werk bijna gedaan is. Niet alleen omdat het werk ontzettend interessant

was, maar zeker ook om het plezier dat de (werk)omgeving me heeft gegeven. Ik wil

daarom graag een woord van dank richten tot de mensen die op één of andere manier

bij mijn onderzoek betrokken waren.

Allereerst natuurlijk mijn promotores Prof. Dr J.H. Beijnen en Prof. Dr J.H.M. Schellens.

Beste Jos, jij hebt me de kans gegeven om mezelf te ontwikkelen op het gebied van

onderzoek. Ik heb veel geleerd de afgelopen vier jaar. Jij vertelde me bij één van onze

eerste overleggen dat je als OIO de manager bent van je eigen onderzoek. Die

opmerking zit nog in mijn achterhoofd. Ik wil je bedanken voor het vertrouwen dat je

me hebt gegeven, waardoor ik het gevoel had dat ik ook echt de manager kon zijn. Elk

overleg met jou leidde tot nieuwe ideeën, wat me motiveerde om steeds weer een

stapje verder te gaan, ook als het even niet mee zat. Jos, ik ben blij dat ik nog vier jaar in

het Slotervaartziekenhuis zal zijn en dat ik me verder kan gaan ontwikkelen in de

ziekenhuisfarmacie.

Beste Jan, ik wil je bedanken voor je waardevolle klinische blik. Jij hebt ervoor gezorgd

dat we de ICP-MS konden inzetten voor het beantwoorden van klinisch relevante

vragen. Mijn dank hiervoor. Daarnaast ben ik je zeer erkentelijk voor de kritische

beoordeling van mijn manuscripten.

Veel mensen hebben bijgedragen aan het onderzoek dat beschreven staat in dit

proefschrift.

Beste Willem, bedankt voor je inzichten en voor de discussies met betrekking tot de

neuropathie. Ik heb je aandeel in het opzetten van de klinische studie erg gewaardeerd.

Alwin, het laatste jaar van mijn promotie heb jij me erg geholpen. Ik ben blij dat je de

resultaten van de klinische studie met mij wilde bespreken. Je heldere ideeën en de

daaropvolgende nonmem analyses hebben ertoe geleid dat we meer uit de

neurotoxiciteitsstudie hebben kunnen halen dan ik in eerste instantie had gedacht. Ik

ben blij dat ik de komende vier jaar nog meer van je mag leren.

Dankwoord

255

Matthijs, wat vind ik het jammer dat je niet meer in het Slotervaart werkt. Onze

brainstormsessies maakten mij enthousiast. Ik vind je een goede onderzoeker. Ook jij

neemt geen genoegen met dat wat ‘aangenomen’ wordt en je bent nieuwsgierig naar

de achtergrond van de aanname. Ik ben blij dat ik de ICP-MS kennis met je kon delen en

dat je altijd klaar stond om samen te ‘troubleshooten’. De kritische blik waarmee jij mijn

review hebt bekeken, heb ik zeer gewaardeerd. Hilde, de overleggen met jou waren voor

mij verhelderend. Jouw analytisch inzicht heeft me bij een aantal hoofdstukken uit dit

proefschrift zeer geholpen. Dick, hoeveel e-mails hebben wij over en weer gestuurd de

afgelopen jaren? Dank je voor het kritisch meedenken en het uitvoeren van de 32P-

analyses. Michel, ook al zijn de MS analyses uiteindelijk niet in mijn proefschrift terecht

gekomen, ik wil je toch bedanken voor je hulp daarbij.

I would also like to thank Karel, Henk, Peter, Henk, Ueli, Yolande, and Stephen from

Varian. The intensive contact we had during the first two years have greatly enlarged my

knowledge of ICP-MS. Dear Stephen, thank you for our fruitful discussions. It always

surprised me when I received a reply within 12 hours. It is a great pleasure that you will

come all the way from Australia to attend my defence.

De MDL artsen van het AVL en Maria wil ik graag bedanken voor hun hulp bij de

N06DCM studie. Marja, Brigitte, Cecile, Ninja, Harm, Dilsha en Jolanda, bedankt dat jullie

altijd klaarstonden om me te helpen bij de klinische studies. Het priklab en de

verpleegkundigen van 4C wil ik bedanken voor het afnemen van de bloedmonsters. Ik

wil ook de patiënten bedanken die allemaal mee wilden doen met het onderzoek. De

enorme motivatie die de patiënten uitstraalden, heeft mij verrast. De gesprekken

duurden altijd net iets langer dan noodzakelijk en ik heb veel van hun verhalen geleerd.

De ziekenhuisapotheken die deel hebben genomen aan het veegproefonderzoek wil ik

bedanken voor hun bijdrage.

En dan de mensen van het lab waar ik heel wat uurtjes heb doorgebracht. Bedankt voor

jullie gezelligheid. Abadi, dank je voor je interesse en natuurlijk voor het lenen van je

pipetten. Bas, jouw technisch inzicht heeft me vaak vooruit geholpen. Ciska en Caroline,

ik zal geen herrie meer maken met de centrifuges. Roel, we hebben elkaar de laatste

jaren goed in de gaten kunnen houden. Af en toe voelde ik me alsof ik in een vissekom

zat. Heb je me ooit zien zingen in het ICP-MS hok? Ik hoop het niet. Ik vond het leuk dat

je af en toe binnenwandelde om te vragen hoe het ging.

Dan de collega OIOs. Tja, wat kan ik zeggen. Toen ik ging verhuizen had ik het gevoel dat

ik een warm nest moest verlaten. Judith, Marie-Christine (MC) en Annemieke, ik vond het

leuk om in de zonnetempel met jullie op de kamer te zitten. Het was er rustig en

gezellig, een hele goede combinatie. Judith, wij deelden de koude kant van de kamer.

Succes met de laatste loodjes en ik hoop je met de cursussen te zien. MC, jouw harde

werken gaat beloond worden. Nog even en je proefschrift is klaar. Dan is het tijd voor de

volgende uitdaging. En als je zin hebt in thee, je weet me te vinden. Annemieke, ik hoop

256

je nog vaak te spreken de komende jaren. Ook mijn oud-kamergenoten in de

zonnetempel Nathalie en in de onderwereld Monique en Liesbeth; bedankt.

Anthe, ik mis je hier in het Slotervaart. Dank je voor de leuke, relativerende gesprekken.

Binnenkort weer naar de sauna? Ly, jij hebt de gave om bij mensen een lach op het

gezicht te toveren. Ik hoop nog vele (fiets)uurtjes met je door te brengen. En dank je

voor de vele milliliters bloed. Tussen de vele OIO donoren was jij mijn vaste slachtoffer.

Robert, Rob, Ron en Joost, jullie passie voor het onderzoek werkt aanstekelijk. Dank jullie

voor de leuke gesprekken. En Ron, ik vind het fijn dat ik je af en toe lastig mocht vallen

met computervragen. David en Stijn, de uurtjes in het kinetiekhok waren leuk. Ik ben

ervan onder de indruk hoe jullie het samplen weten te managen. David, dank je voor je

hulp bij het afnemen van de bloedmonsters. Corine, succes met de rest van je promotie

en wat daarna mag komen. Het gaat je lukken! Susanne, volg je hart, dan komt het

helemaal goed. Carola, succes met je verdere promotie. Liia, veel plezier met je nieuwe

baan en natuurlijk succes voor 21 februari! Jolanda, je weet me te vinden als je Ru

vragen hebt. Claudia, succes met je onderzoek. Bas, dank je voor het lenen van je scriptie

en veel plezier met je interessante onderzoek. Roos, Maarten, Sander en Nienke, ook al

zitten jullie niet in de keet, het was leuk om jullie af en toe te spreken. Markus, tot in

december.

Marjolein, toen ik net begon kwam ik bij je op de kamer. Terwijl Liesbeth, Monique en jij

al ver in jullie promotie waren, was voor mij alles nieuw. Ik heb veel van jullie geleerd. Nu

ben ik net weer naar je kamer verhuisd, alleen is die kamer dit keer heel wat kleiner en

lijkt het meer op een hok. Ik heb het erg naar mijn zin en vind het leuk om met je samen

te werken. Sabien, ik vond het jammer dat je wegging. Ik heb onze gesprekken erg

gewaardeerd en ik ben blij dat het zo goed bevalt bij Solvay. Tot snel.

Dear Herman, Adile and the other people of the Folkhälsan Research Center, thank you

for introducing me to science. The seven months that Rafaëlla and I spent in Finland,

opened the world of research to me. You made me think about doing a PhD. This thesis

is the result.

Naast het werk was er ook tijd voor ontspanning. Wat is het leven zonder muziek. Sabijn

en Jorden, jullie twee en onze avonden jammen geven mij energie. Ik ben ontzettend

blij dat ik jullie ken. Let’s sing!

Jos en Hanny: Zoals ik al eerder schreef zijn er slechts een aantal plaatsen in het

proefschrift waar je iets persoonlijks mag laten zien. Naast het dankwoord is de omslag

een tweede plek. Jos, ik heb me laten verrassen door jouw ontwerp. Dat was spannend,

maar je hebt mij perfect ‘gelezen’. Je hebt de vier jaar van mijn onderzoek in één

tekening weten te vangen. Persoonlijker had de omslag niet kunnen zijn. Ik ben je

hiervoor heel dankbaar. Hanny, dank je voor dat ik me altijd welkom voel bij jullie.

Lieve paps en mams, ik vind jullie geweldig en ik ben er ontzettend trots op dat ik jullie

dochter ben. Telkens kijk ik er weer naar uit dat we elkaar zien en spreken. Ik heb het

Dankwoord

257

idee dat we elkaar echt leren kennen en dat vind ik speciaal. En mama, het blijft

bijzonder dat we altijd op hetzelfde moment naar de telefoon grijpen om elkaar te

bellen.

Bart en Willem, mijn grote stoere broers. Ik ben trots op jullie en ik vind het leuk dat jullie

in Claudia en Elisabeth twee bijzondere dames hebben gevonden. En Willem, super dat

je Jos bij het digitaal maken van de omslag hebt geholpen.

Marieke, er zijn weinig mensen die me zo blij maken als jij. En ik blijf het stoer vinden

hoe je de mannen aftroeft als je op de mountainbike zit. Binnenkort naar het strand?

Rafaëlla, wat hebben wij veel geweldige dingen gedaan samen de laatste 10 jaar. Ook de

eerste onderzoeksstappen heb ik samen met jou genomen en dat beviel zo goed dat we

beiden verder zijn gegaan met onderzoek. Ondanks dat onze wegen nu iets meer van

elkaar gescheiden zijn dan voorheen, kruisen ze elkaar gelukkig nog vaak. Ik vind het

geweldig dat jij en Marieke als paranimfen achter me willen staan.

Martien, het leven voelt voor mij als een ontdekkingsreis en ik vind het heerlijk om die

met jou te mogen maken. Ik heb zin in het volgende deel van de reis!

Elke

Amsterdam 2007

Curriculum Vitae

259

Curriculum Vitae

Elke Brouwers werd op 8 november 1977 geboren te Helmond. In 1996 behaalde zij het

atheneum diploma aan het Dr Knippenbergcollege in Helmond. Vervolgens studeerde

zij farmacie aan de Universiteit Utrecht. Ter afsluiting van de doctoraal opleiding volgde

zij een wetenschappelijke stage aan het Folkhälsan Research Center in Helsinki. Er

werden fluoro-immunoassays ontwikkeld voor de bepaling van isoflavonoïden in

plasma en urine onder supervisie van Prof. Dr H. Adlercreutz. In 2002 volgde zij een

stage in de Radcliff Infirmary in Oxford waarna zij het apothekersdiploma behaalde. In

2003 begon zij aan het promotieonderzoek dat beschreven is in dit proefschrift, onder

leiding van Prof. Dr J.H.Beijnen en Prof. Dr J.H.M. Schellens

260

List of publications

1. Brouwers EEM, L’homme RFA, Al-Maharik N, Lapcik O, Hampl R, Wahala K, Mikola H,

Adlercreutz H. Time-resolved fluoro-immunoassay for equol in plasma and urine. J

Steroid Biochem Mol Biol 2003; 84(5): 577

2. L’homme RFA, Brouwers EEM, Al-Maharik N, Lapcik O, Hampl R, Wahala K, Mikola H,

Adlercreutz H. Time-resolved fluoro-immunoassay for plasma and urine O-DMA. J

Steroid Biochem Mol Biol 2002; 81(4-5): 353-61

3. Brouwers EEM, Tibben MM, Joerger M, van Tellingen O, Rosing H, Schellens JHM,

Beijnen JH. Determination of oxaliplatin in human plasma and plasma ultrafiltrate

by graphite-furnace atomic-absorption spectrometry. Anal Bioanal Chem. 2005;

382(7): 1484-90.

4. Brouwers EEM, Tibben MM, Rosing H, Hillebrand, MJX, Joerger M, Schellens JHM,

Beijnen JH. Sensitive inductively coupled plasma mass spectrometry assay for the

determination of platinum originating from cisplatin, carboplatin, and oxaliplatin in

human plasma ultrafiltrate. J Mass Spectrom. 2006; 41(9): 1186-94.

5. Brouwers EEM, Huitema ADR, Bakker EN, Douma JW, Schimmel KJM, van Weringh G,

de Wolf PJ, Schellens JHM, Beijnen JH. Monitoring of platinum surface contamination

in seven Dutch hospital pharmacies using inductively coupled plasma mass

spectrometry. Int Arch Occup Environ Health 2007; 80: 689-699

6. Brouwers EEM, Tibben MM, Rosing H, Schellens JHM, Beijnen JH. Determination of

ruthenium originating from the investigational anti-cancer drug NAMI-A in human

plasma ultrafiltrate, plasma, and urine by inductively coupled plasma mass

spectrometry. Rapid Commun Mass Spectrom 2007; 21(9): 1521-1530

7. Brouwers EEM, Tibben MM, Rosing H, Schellens JHM, Beijnen JH. The application of

inductively coupled plasma mass spectrometry in clinical pharmacological oncology

research. Submitted.

8. Brouwers EEM, Tibben MM, Pluim D, Rosing H, Boot H, Cats A, Schellens JHM,

Beijnen JH. Inductively coupled plasma mass spectrometric analysis of the total

amount of platinum-DNA adducts in peripheral blood mononuclear cells and tissue

from patients treated with cisplatin. Submitted.

List of publications

261

9. Brouwers EEM, Huitema ADR, Schellens JHM, Beijnen JH. The effects of sulfur-

containing compounds and gemcitabine on the binding of cisplatin to plasma

proteins and DNA determined by ICP-MS and HPLC-ICP-MS. Submitted.

10. Brouwers EEM, Huitema ADR, Schellens JHM, Beijnen JH. Long-term platinum

retention after treatment with cisplatin and oxaliplatin. Submitted.

11. Brouwers EEM, Huitema ADR, Boogerd W, Schellens JHM, Beijnen JH. Persistent

neuropathy after treatment with cisplatin and oxaliplatin. Submitted.


Recommended