You are on page 1of 4

Journal of Antimicrobial Chemotherapy (2002) 49, 597599

The term biocide is increasingly being used to describe


compounds with antiseptic, disinfectant or, sometimes,
preservative activity. A compound might be used in only
one such capacity or possess two or even all of these pro-
perties.
1
Until fairly recently, there were two long-held
general opinions about biocides. The rst was that, as long
as they were effective, there was little reason (apart from
academic value) to determine how they achieved their
inhibitory or lethal effects. The second, widely perceived,
view was that antiseptics and disinfectants acted as general
protoplasmic poisons and, as such, merited little attention.
At the beginning of the twentieth century, there were
few drugs available for the treatment of infections. Anti-
septics and disinfectants had at that stage been employed
for various purposes and in various guises, notable ex-
amples being phenol (carbolic acid), mercuric chloride,
chlorine, hypochlorites and iodine. Quaternary ammonium
compounds (QACs) were described in 1916 but were
not used commercially for another 19 years or so.
2
Early
studies on the action of such compounds concentrated on
the kinetics of bacterial inactivation,
3
although Cooper
4
notably described the relationship between phenolics
(phenol and meta-cresol) and bacterial proteins as being of
importance in their mechanism of disinfection. In particu-
lar, it was considered that the protein structure inside the
bacterial cell was affected. Subsequently, Knaysi et al.
5,6
reported on the manner of death of bacteria, mainly
Escherichia coli, exposed to mild chemical and physical
agents, concluding that the order of death was determined
by the distribution of resistances among the cells. Later,
Jordan & Jacobs
7
found the distribution of resistance of
E. coli treated with phenol to be normal at all phenol
concentrations used. Interestingly, specic enzymes were
considered by some workers
8,9
to be involved in bacterial
inactivation by biocidal agents.
Penicillin (benzylpenicillin, penicillin G) began to be used
as a chemotherapeutic antibiotic in the 1940s. Early studies
on its mode of action were undertaken by Gardner
10
and Duguid,
11
and Eagle & Musselman
12
demonstrated a
paradoxical effect of high concentrations on staphylococci.
However, it was not until the isolation of the Park
nucleotides and subsequent studies, especially those by
Strominger et al.,
13,14
that it was realized that not only was
penicillin an important antibiotic but also that it was a valu-
able tool in studying bacterial peptidoglycan biosynthesis.
Further discoveries, the production of other -lactam
antibiotics and work demonstrating that tetracyclines and
other antibiotics inhibited bacterial protein synthesis,
15
provided a stimulus for additional and comprehensive
investigations to be undertaken on bacterial inactivation
mechanisms (and also, naturally, on bacterial resistance
mechanisms) by such selectively toxic drugs. It was realized
that these aspects, linked to structureactivity relation-
ships, provided the key to the development of many new,
improved antibiotic molecules.
Until recently, there has not been the same enthusiasm
for studying the mode of action of biocidal agents. How-
ever, the literature does contain a surprising, if scattered,
number of publications about the mechanisms of inhibition
and inactivation of Gram-positive non-sporulating bacteria
(although the information on mycobacteria is disappoint-
ing), bacterial spores (as well as germinating and out-
growing ones) and Gram-negative bacteria.
16,17
Less data
are available about the mechanisms of fungal and viral
inactivation, with very few comprehensive studies on the
mechanisms of protozoal inactivation.
1820
It is rather
surprising that information on the mechanisms of viral
inactivation, in particular, is sparse. Furthermore, it is not
known why the MICs of biocides such as chlorhexidine and
QACs are of an equivalent order for both mycobacteria
and staphylococci whereas these cationic biocides possess
low mycobactericidal potency but are rapidly lethal to the
latter.
21
Contrary to early suggestions by Rahn et al.,
8,9
it is
now considered to be highly unlikely that bacterial cells
possess a single type of target enzyme the inhibition or
inactivation of which by a biocide is responsible for a loss of
cell viability.
16
Thus, although triclosan has repeatedly
been shown to inhibit enoyl reductase (see below),
22,23
work from this laboratory has demonstrated that other
events are involved in bacterial inactivation.
24
597
Mechanisms of antimicrobial action of antiseptics and disinfectants:
an increasingly important area of investigation
A. D. Russell*
Welsh School of Pharmacy, Cardiff University, Cardiff CF10 3XF, UK
*Tel: 44-2920-875812; Fax: 44-2920-874149; E-mail: russellD2@cardiff.ac.uk
2002 The British Society for Antimicrobial Chemotherapy
JAC
Leading article
It has become clear that some antiseptics and disinfect-
ants on the one hand and antibiotics on the other have simi-
lar effects on bacteria. For example, (1) lament formation
is induced in Gram-negative bacteria by both antibiotics
(-lactams, novobiocin, uoroquinolones) and biocides
(phenoxyethanol, phenylethyl alcohol, chloroacetamide,
acridines);
25
(2) inhibition of enoyl reductase, involved in
fatty acid synthesis, is inhibited by both isoniazid, an import-
ant antitubercular drug, and the bisphenol (phenylether)
triclosan;
26
and (3) autolysis brought about by low con-
centrations of phenolics and inorganic and organic mercury
compounds has been suggested as being similar to that
following bacterial exposure to penicillin.
16
However, it is
interesting that E. coli cells exposed to phenoxyethanol,
proavine or chloroacetamide showed the same suscept-
ibility to ampicillin and noroxacin as cells not pre-exposed
to a biocide.
25
It would be interesting to determine whether
biocide-resistant E. coli cells showed differences in
response to these important antibiotics. A correlation was
found earlier by McKellar et al.
27
as a consequence of a
non-specic increase in impermeability. Additionally, it
must be noted that the end result may be brought about by
different inhibitory or lethal mechanisms.
28
It has also become increasingly obvious that insuscept-
ibility mechanisms to biocides and antibiotics may be
similar although not necessarily identical.
15
A natural
(intrinsic) insusceptibility to both groups may be shown by
Gram-negative bacteria and mycobacteria, with outer
membrane or cell wall impermeability being responsible.
Additionally, efux systems may remove toxic drug and
biocide molecules from the cell, although a key issue here
relates to the concentrations at which the compounds are
used. Thus, with drugs it is usually necessary to equate
MICs or minimum bactericidal concentrations (MBCs)
with blood serum levels. By contrast, biocides are essen-
tially used for external purposes at concentrations likely to
be considerably higher than MICs or MBCs; such con-
centrations are unlikely to be efuxed from bacterial cells.
Nevertheless, the possibility remains that low, residual
concentrations could act as a focus for the survival of
organisms containing efux genes or for the gradual or
rapid development of biocide-resistant bacteria.
29
This
reinforces the argument that the effects of biocides on bac-
terial (and, indeed, other types of microbial) cells should be
examined over a wide range of concentrations.
30
There are other reasons for studying the mechanisms of
biocides. At present, comparatively little is known about
the uptake of biocides into bacterial (and other microbial)
cells. The probability exists that targeted drug delivery,
whereby a biocide can readily reach its target site(s) within
a cell, could lead to greater efcacy. An aspect that needs
to be considered is the possible design of new biocidal
molecules based on known effects of current molecules;
there is little evidence that this is happening. The only
signicant new biocides to be introduced in the past few
years have been ortho-phthalaldehyde (OPA)
31
and those
based on peracetic acid.
32
However, both OPA and per-
acetic acid are themselves old molecules that have been
examined in a new, antimicrobial context.
In recent years, rotation of disinfectants in hospitals and
elsewhere, e.g. in the pharmaceutical and food industries,
has been advocated to prevent the development of bac-
terial resistance. It has been claimed that, ideally, one dis-
infectant should be replaced by another having a dissimilar
mechanism of action.
33
Clearly, a knowledge of the ways in
which such agents act is an essential component of such a
policy.
In conclusion, there is an urgent need to investigate
more fully the nature of the inhibitory and lethal effects of
antiseptics and disinfectants on a range of microorganisms
and microbial entities. Possible multiple target sites and
concentration-dependent effects would form an important
aspect of such studies, which would also provide a better
understanding of intrinsic and acquired bacterial resistance
mechanisms and of the possible linkage between biocide
usage and antibiotic resistance.
References
1. Russell, A. D., Furr, J. R. & Maillard, J.-Y. (1997). Microbial
susceptibility and resistance to biocides. ASM News 63, 4817.
2. Russell, A. D. (2002). Introduction of biocides into clinical
practice and the impact on antibiotic resistant bacteria. Journal of
Applied Microbiology, Symp. Suppl. (in press).
3. Kronig, B. & Paul, T. L. (1897). The chemical foundations of the
study of disinfection and of the action of poisons. Zeitschrift fr
Hygiene 25, 1112.
4. Cooper, E. A. (1912). On the relationship of phenol and m-cresol
to proteins: a contribution to our knowledge of the mechanism of
disinfection. Biochemical Journal 6, 36287.
5. Knaysi, G. (1930). Disinfection. I. The development of our know-
ledge of disinfection. Journal of Infectious Diseases 47, 293302.
6. Knaysi, G. & Morris, G. (1930). The manner of death of certain
bacteria and yeast when subjected to mild chemical and physical
agents. Journal of Infectious Diseases 47, 30317.
7. Jordan, R. C. & Jacobs, S. E. (1944). Studies on the dynamics
of disinfection. I. New data on the reaction between phenol and
Bact. coli using an improved technique, together with an analysis of
the distribution of resistance amongst the cells of the bacterial
population studied. Journal of Hygiene (Cambridge) 43, 27589.
8. Rahn, O. & Schroder, W. R. (1941). Inactivation of enzymes as
the cause of death in bacteria. Biodynamica 3, 199208.
9. Roberts, M. H. & Rahn, O. (1946). The amount of enzyme
inactivation at bacteriostatic and bactericidal concentrations of dis-
infectants. Journal of Bacteriology 42, 63944.
10. Gardner, A. D. (1940). Morphological effects of penicillin on
bacteria. Nature 146, 8378.
11. Duguid, J. P. (1946). Sensitivity of bacteria to the action of peni-
cillin. Edinburgh Medical Journal 53, 40112.
12. Eagle, H. & Musselman, A. D. (1948). Rate of bactericidal
action of penicillin in vitro as a function of its concentration and its
598
Leading article
paradoxically reduced activity at high concentrations against certain
organisms. Journal of Experimental Medicine 88, 99131.
13. Park, J. T. & Strominger, J. L. (1957). Mode of action of peni-
cillin. Biochemical basis for the mechanism of action of penicillin and
its selective toxicity. Science 125, 99101.
14. Tipper, D. J. & Strominger, J. L. (1965). Mechanism of action of
penicillin: a proposal based on their structural similarity to acyl-D-
alanyl-D-alanine. Proceedings of the National Academy of Sciences,
USA 54, 113341.
15. Russell, A. D. & Chopra, I. (1996). Understanding Antibacterial
Action and Resistance, 2nd edn. Ellis Horwood, Chichester, UK.
16. Hugo, W. B. (1999). Disinfection mechanisms. In Principles
and Practice of Disinfection, Preservation and Sterilization, 3rd edn,
(Russell, A. D., Hugo, W. B. & Ayliffe, G. A. J., Eds), pp. 25883.
Blackwell Science, Oxford.
17. Russell, A. D. (1990). The bacterial spore and chemical spori-
cidal agents. Clinical Microbiology Reviews 3, 99119.
18. McDonnell, A. D. & Russell, A. D. (1999). Antiseptics and dis-
infectants: activity, action and resistance. Clinical Microbiology
Reviews 12, 14779.
19. Maillard, J.-Y. & Russell, A. D. (1997). Viricidal activity and
mechanisms of action of biocides. Science Progress 80, 287315.
20. Turner, N. A., Russell, A. D., Furr, J. R. & Lloyd, D. (1999).
Editorial: Acanthamoeba spp., antimicrobial agents and contact
lenses. Science Progress 82, 18.
21. Russell, A. D. (1996). Activity of biocides against mycobacteria.
Journal of Applied Bacteriology, Symp. Suppl. 81, 87S101S.
22. McMurry, L. M., Oethinger, M. & Levy, S. B. (1998). Triclosan
targets lipid synthesis. Nature 394, 5312.
23. Heath, R. J., Rubin, J. R., Holland, D. R., Snow, M. E. & Rock,
C. O. (1999). Mechanism of triclosan inhibition of bacterial fatty acid
synthesis. Journal of Biological Chemistry 274, 111104.
24. Suller, M. T. E. & Russell, A. D. (2000). Triclosan and antibiotic
resistance in Staphylococcus aureus. Journal of Antimicrobial
Chemotherapy 46, 118.
25. Ng, E. G.-L., Jones, S., Leong, S. H. & Russell, A. D. (2002).
Biocides and antibiotics with apparently similar actions on bacteria:
is there the potential for cross-resistance? Journal of Hospital
Infection (in press).
26. McMurry, L. M., McDermott, P. F. & Levy, S. B. (1999). Genetic
evidence that InhA of Mycobacterium smegmatis is a target for
triclosan. Antimicrobial Agents and Chemotherapy 43, 7113.
27. McKellar, R. C., McKenzie, C. N. & Kushner, D. J. (1976).
Correlation of resistance to proavine and penicillin in Escherichia
coli. Antimicrobial Agents and Chemotherapy 10, 7657.
28. Morris, A. & Russell, A. D. (1971). The mode of action of
novobiocin. In Progress in Medicinal Chemistry, (Ellis, G. P. & West,
G. B., Eds), Vol. 41, pp. 3959. Butterworths, London.
29. Thomas, L., Lambert, R. J. W., Maillard, J.-Y. & Russell, A. D.
(2000). Development of resistance to chlorhexidine diacetate in
Pseudomonas aeruginosa and the effect of a residual concentration.
Journal of Hospital Infection 46, 297303.
30. Russell, A. D. & McDonnell, G. (2000). Concentration: a major
factor in studying biocidal action. Journal of Hospital Infection 44,
13.
31. Walsh, S. E., Maillard, J.-Y. & Russell, A. D. (1999). Ortho-
phthalaldehyde: possible alternative to glutaraldehyde for high-level
disinfection. Journal of Applied Microbiology 86, 103946.
32. Block, S. S. (2001). Peroxygen compounds. In Disinfection,
Sterilization and Preservation, 5th edn, (Block, S. S., Ed.),
pp. 185204. Lippincott, Williams & Williams, Philadelphia, PA.
33. Murtough, S. M., Hiom, S. J., Palmer, M. & Russell, A. D.
(2001). Biocide rotation in the healthcare setting: is there a case for
policy implementation? Journal of Hospital Infection 48, 16.
599

You might also like