首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
From the highest mountains to biology''s own Everest—the brain—Reichardt tackles the biggest challenges of climbing and biology.Louis Reichardt''s scientific career has spanned the simple and the complex. As a graduate student, Reichardt helped to uncover the now renowned DNA regulatory mechanisms that allow one of the simplest life forms, lambda phage, either to hide within a cell or to make its presence known via massive replication (1).Open in a separate windowLouis ReichardtLetting his curiosity for the unknown guide him, Reichardt then forayed into a much more intricate system, the brain. As a postdoc, he showed that growth conditions influenced which neurotransmitters are synthesized by isolated neurons (2). Later, as a professor at UCSF, he discovered synaptotagmin, using the first monoclonal antibody that defined a synaptic vesicle membrane protein (3), showed that expression levels of nerve growth factor in target tissues correlate with the density of innervation (4), and characterized the properties of mice lacking genes encoding the neurotrophins and their Trk receptors (5, 6).Now a professor and director of the Neuroscience Program at UCSF, Reichardt''s laboratory still studies the interface of cell biology and neurobiology, including the involvement of cell adhesion molecules in synaptic development (7). He explains that science is not so different from his favorite hobby, mountain climbing.
“The gist of it was ‘Others more foolish might try this, but it was not for us.’ I thought, ‘We''ll obviously have to do it.’”
  相似文献   

2.
3.
Zhang JY 《EMBO reports》2011,12(4):302-306
How can grass-roots movements evolve into a national research strategy? The bottom-up emergence of synthetic biology in China could give some pointers.Given its potential to aid developments in renewable energy, biosensors, sustainable chemical industries, microbial drug factories and biomedical devices, synthetic biology has enormous implications for economic development. Many countries are therefore implementing strategies to promote progress in this field. Most notably, the USA is considered to be the leader in exploring the industrial potential of synthetic biology (Rodemeyer, 2009). Synthetic biology in Europe has benefited from several cross-border studies, such as the ‘New and Emerging Science and Technology'' programme (NEST, 2005) and the ‘Towards a European Strategy for Synthetic Biology'' project (TESSY; Gaisser et al, 2008). Yet, little is known in the West about Asia''s role in this ‘new industrial revolution'' (Kitney, 2009). In particular, China is investing heavily in scientific research for future developments, and is therefore likely to have an important role in the development of synthetic biology.Initial findings seem to indicate that the emergence of synthetic biology in China has been a bottom-up construction of a new scientific framework…In 2010, as part of a study of the international governance of synthetic biology, the author visited four leading research teams in three Chinese cities (Beijing, Tianjin and Hefei). The main aims of the visits were to understand perspectives in China on synthetic biology, to identify core themes among its scientific community, and to address questions such as ‘how did synthetic biology emerge in China?'', ‘what are the current funding conditions?'', ‘how is synthetic biology generally perceived?'' and ‘how is it regulated?''. Initial findings seem to indicate that the emergence of synthetic biology in China has been a bottom-up construction of a new scientific framework; one that is more dynamic and comprises more options than existing national or international research and development (R&D) strategies. Such findings might contribute to Western knowledge of Chinese R&D, but could also expose European and US policy-makers to alternative forms and patterns of research governance that have emerged from a grass-roots level.…the process of developing a framework is at least as important to research governance as the big question it might eventually addressA dominant narrative among the scientists interviewed is the prospect of a ‘big-question'' strategy to promote synthetic-biology research in China. This framework is at a consultation stage and key questions are still being discussed. Yet, fieldwork indicates that the process of developing a framework is at least as important to research governance as the big question it might eventually address. According to several interviewees, this approach aims to organize dispersed national R&D resources into one grand project that is essential to the technical development of the field, preferably focusing on an industry-related theme that is economically appealling to the Chinese public.Chinese scientists have a pragmatic vision for research; thinking of science in terms of its ‘instrumentality'' has long been regarded as characteristic of modern China (Schneider, 2003). However, for a country in which the scientific community is sometimes described as an “uncoordinated ‘bunch of loose ends''” (Cyranoski, 2001) “with limited synergies between them” (OECD, 2007), the envisaged big-question approach implies profound structural and organizational changes. Structurally, the approach proposes that the foundational (industry-related) research questions branch out into various streams of supporting research and more specific short-term research topics. Within such a framework, a variety of Chinese universities and research institutions can be recruited and coordinated at different levels towards solving the big question.It is important to note that although this big-question strategy is at a consultation stage and supervised by the Ministry of Science and Technology (MOST), the idea itself has emerged in a bottom-up manner. One academic who is involved in the ongoing ministerial consultation recounted that, “It [the big-question approach] was initially conversations among we scientists over the past couple of years. We saw this as an alternative way to keep up with international development and possibly lead to some scientific breakthrough. But we are happy to see that the Ministry is excited and wants to support such an idea as well.” As many technicalities remain to be addressed, there is no clear time-frame yet for when the project will be launched. Yet, this nationwide cooperation among scientists with an emerging commitment from MOST seems to be largely welcomed by researchers. Some interviewees described the excitement it generated among the Chinese scientific community as comparable with the establishment of “a new ‘moon-landing'' project”.Of greater significance than the time-frame is the development process that led to this proposition. On the one hand, the emergence of synthetic biology in China has a cosmopolitan feel: cross-border initiatives such as international student competitions, transnational funding opportunities and social debates in Western countries—for instance, about biosafety—all have an important role. On the other hand, the development of synthetic biology in China has some national particularities. Factors including geographical proximity, language, collegial familiarity and shared interests in economic development have all attracted Chinese scientists to the national strategy, to keep up with their international peers. Thus, to some extent, the development of synthetic biology in China is an advance not only in the material synthesis of the ‘cosmos''—the physical world—but also in the social synthesis of aligning national R&D resources and actors with the global scientific community.To comprehend how Chinese scientists have used national particularities and global research trends as mutually constructive influences, and to identify the implications of this for governance, this essay examines the emergence of synthetic biology in China from three perspectives: its initial activities, the evolution of funding opportunities, and the ongoing debates about research governance.China''s involvement in synthetic biology was largely promoted by the participation of students in the International Genetically Engineered Machine (iGEM) competition, an international contest for undergraduates initiated by the Massachusetts Institute of Technology (MIT) in the USA. Before the iGEM training workshop that was hosted by Tianjin University in the Spring of 2007, there were no research records and only two literature reviews on synthetic biology in Chinese scientific databases (Zhao & Wang, 2007). According to Chunting Zhang of Tianjin University—a leading figure in the promotion of synthetic biology in China—it was during these workshops that Chinese research institutions joined their efforts for the first time (Zhang, 2008). From the outset, the organization of the workshop had a national focus, while it engaged with international networks. Synthetic biologists, including Drew Endy from MIT and Christina Smolke from Stanford University, USA, were invited. Later that year, another training camp designed for iGEM tutors was organized in Tianjin and included delegates from Australia and Japan (Zhang, 2008).Through years of organizing iGEM-related conferences and workshops, Chinese universities have strengthened their presence at this international competition; in 2007, four teams from China participated. During the 2010 competition, 11 teams from nine universities in six provinces/municipalities took part. Meanwhile, recruiting, training and supervising iGEM teams has become an important institutional programme at an increasing number of universities.…training for iGEM has grown beyond winning the student awards and become a key component of exchanges between Chinese researchers and the international communityIt might be easy to interpret the enthusiasm for the iGEM as a passion for winning gold medals, as is conventionally the case with other international scientific competitions. This could be one motive for participating. Yet, training for iGEM has grown beyond winning the student awards and has become a key component of exchanges between Chinese researchers and the international community (Ding, 2010). Many of the Chinese scientists interviewed recounted the way in which their initial involvement in synthetic biology overlapped with their tutoring of iGEM teams. One associate professor at Tianjin University, who wrote the first undergraduate textbook on synthetic biology in China, half-jokingly said, “I mainly learnt [synthetic biology] through tutoring new iGEM teams every year.”Participation in such contests has not only helped to popularize synthetic biology in China, but has also influenced local research culture. One example of this is that the iGEM competition uses standard biological parts (BioBricks), and new BioBricks are submitted to an open registry for future sharing. A corresponding celebration of open-source can also be traced to within the Chinese synthetic-biology community. In contrast to the conventional perception that the Chinese scientific sector consists of a “very large number of ‘innovative islands''” (OECD, 2007; Zhang, 2010), communication between domestic teams is quite active. In addition to the formally organized national training camps and conferences, students themselves organize a nationwide, student-only workshop at which to informally test their ideas.More interestingly, when the author asked one team whether there are any plans to set up a ‘national bank'' for hosting designs from Chinese iGEM teams, in order to benefit domestic teams, both the tutor and team members thought this proposal a bit “strange”. The team leader responded, “But why? There is no need. With BioBricks, we can get any parts we want quite easily. Plus, it directly connects us with all the data produced by iGEM teams around the world, let alone in China. A national bank would just be a small-scale duplicate.”From the beginning, interest in the development of synthetic biology in China has been focused on collective efforts within and across national borders. In contrast to conventional critiques on the Chinese scientific community''s “inclination toward competition and secrecy, rather than openness” (Solo & Pressberg, 2007; OECD, 2007; Zhang, 2010), there seems to be a new outlook emerging from the participation of Chinese universities in the iGEM contest. Of course, that is not to say that the BioBricks model is without problems (Rai & Boyle, 2007), or to exclude inputs from other institutional channels. Yet, continuous grass-roots exchanges, such as the undergraduate-level competition, might be as instrumental as formal protocols in shaping research culture. The indifference of Chinese scientists to a ‘national bank'' seems to suggest that the distinction between the ‘national'' and ‘international'' scientific communities has become blurred, if not insignificant.However, frequent cross-institutional exchanges and the domestic organization of iGEM workshops seem to have nurtured the development of a national synthetic-biology community in China, in which grass-roots scientists are comfortable relying on institutions with a cosmopolitan character—such as the BioBricks Foundation—to facilitate local research. To some extent, one could argue that in the eyes of Chinese scientists, national and international resources are one accessible global pool. This grass-roots interest in incorporating local and global advantages is not limited to student training and education, but also exhibited in evolving funding and regulatory debates.In the development of research funding for synthetic biology, a similar bottom-up consolidation of national and global resources can also be observed. As noted earlier, synthetic-biology research in China is in its infancy. A popular view is that China has the potential to lead this field, as it has strong support from related disciplines. In terms of genome sequencing, DNA synthesis, genetic engineering, systems biology and bioinformatics, China is “almost at the same level as developed countries” (Pan, 2008), but synthetic-biology research has only been carried out “sporadically” (Pan, 2008; Huang, 2009). There are few nationally funded projects and there is no discernible industrial involvement (Yang, 2010). Most existing synthetic-biology research is led by universities or institutions that are affiliated with the Chinese Academy of Science (CAS). As one CAS academic commented, “there are many Chinese scientists who are keen on conducting synthetic-biology research. But no substantial research has been launched nor has long-term investment been committed.”The initial undertaking of academic research on synthetic biology in China has therefore benefited from transnational initiatives. The first synthetic-biology project in China, launched in October 2006, was part of the ‘Programmable Bacteria Catalyzing Research'' (PROBACTYS) project, funded by the Sixth Framework Programme of the European Union (Yang, 2010). A year later, another cross-border collaborative effort led to the establishment of the first synthetic-biology centre in China: the Edinburgh University–Tianjing University Joint Research Centre for Systems Biology and Synthetic Biology (Zhang, 2008).There is also a comparable commitment to national research coordination. A year after China''s first participation in iGEM, the 2008 Xiangshan conference focused on domestic progress. From 2007 to 2009, only five projects in China received national funding, all of which came from the National Natural Science Foundation of China (NSFC). This funding totalled ¥1,330,000 (approximately £133,000; www.nsfc.org), which is low in comparison to the £891,000 funding that was given in the UK for seven Networks in Synthetic Biology in 2007 alone (www.bbsrc.ac.uk).One of the primary challenges in obtaining funding identified by the interviewees is that, as an emerging science, synthetic biology is not yet appreciated by Chinese funding agencies. After the Xiangshan conference, the CAS invited scientists to a series of conferences in late 2009. According to the interviewees, one of the main outcomes was the founding of a ‘China Synthetic Biology Coordination Group''; an informal association of around 30 conference delegates from various research institutions. This group formulated a ‘regulatory suggestion'' that they submitted to MOST, which stated the necessity and implications of supporting synthetic-biology research. In addition, leading scientists such as Chunting Zhang and Huanming Yang—President of the Beijing Genomic Institute (BGI), who co-chaired the Beijing Institutes of Life Science (BILS) conferences—have been active in communicating with government institutions. The initial results of this can be seen in the MOST 2010 Application Guidelines for the National Basic Research Program, in which synthetic biology was included for the first time, among ‘key supporting areas'' (MOST, 2010). Meanwhile, in 2010, NSFC allocated ¥1,500,000 (approximately £150,000) to synthetic-biology research, which is more than the total funding the area had received in the past three years.The search for funding further demonstrates the dynamics between national and transnational resources. Chinese R&D initiatives have to deal with the fact that scientific venture-capital and non-governmental research charities are underdeveloped in China. In contrast to the EU or the USA, government institutions in China, such as the NSFC and MOST, are the main and sometimes only domestic sources of funding. Yet, transnational funding opportunities facilitate the development of synthetic biology by alleviating local structural and financial constraints, and further integrate the Chinese scientific community into international research.This is not a linear ‘going-global'' process; it is important for Chinese scientists to secure and promote national and regional support. In addition, this alignment of national funding schemes with global research progress is similar to the iGEM experience, as it is being initiated through informal bottom-up associations between scientists, rather than by top-down institutional channels.As more institutions have joined iGEM training camps and participated in related conferences, a shared interest among the Chinese scientific community in developing synthetic biology has become visible. In late 2009, at the conference that founded the informal ‘coordination group'', the proposition of integrating national expertise through a big-question approach emerged. According to one professor in Beijing—who was a key participant in the discussion at the time—this proposition of a nationwide synergy was not so much about ‘national pride'' or an aim to develop a ‘Chinese'' synthetic biology, it was about research practicality. She explained, “synthetic biology is at the convergence of many disciplines, computer modelling, nano-technology, bioengineering, genomic research etc. Individual researchers like me can only operate on part of the production chain. But I myself would like to see where my findings would fit in a bigger picture as well. It just makes sense for a country the size of China to set up some collective and coordinated framework so as to seek scientific breakthrough.”From the first participation in the iGEM contest to the later exploration of funding opportunities and collective research plans, scientists have been keen to invite and incorporate domestic and international resources, to keep up with global research. Yet, there are still regulatory challenges to be met.…with little social discontent and no imminent public threat, synthetic biology in China could be carried out in a ‘research-as-usual'' mannerThe reputation of “the ‘wild East'' of biology” (Dennis, 2002) is associated with China'' previous inattention to ethical concerns about the life sciences, especially in embryonic-stem-cell research. Similarly, synthetic biology creates few social concerns in China. Public debate is minimal and most media coverage has been positive. Synthetic biology is depicted as “a core in the fourth wave of scientific development” (Pan, 2008) or “another scientific revolution” (Huang, 2009). Whilst recognizing its possible risks, mainstream media believe that “more people would be attracted to doing good while making a profit than doing evil” (Fang & He, 2010). In addition, biosecurity and biosafety training in China are at an early stage, with few mandatory courses for students (Barr & Zhang, 2010). The four leading synthetic-biology teams I visited regarded the general biosafety regulations that apply to microbiology laboratories as sufficient for synthetic biology. In short, with little social discontent and no imminent public threat, synthetic biology in China could be carried out in a ‘research-as-usual'' manner.Yet, fieldwork suggests that, in contrast to this previous insensitivity to global ethical concerns, the synthetic-biology community in China has taken a more proactive approach to engaging with international debates. It is important to note that there are still no synthetic-biology-specific administrative guidelines or professional codes of conduct in China. However, Chinese stakeholders participate in building a ‘mutual inclusiveness'' between global and domestic discussions.One of the most recent examples of this is a national conference about the ethical and biosafety implications of synthetic biology, which was jointly hosted by the China Association for Science and Technology, the Chinese Society of Biotechnology and the Beijing Institutes of Life Science CAS, in Suzhou in June 2010. The discussion was open to the mainstream media. The debate was not simply a recapitulation of Western worries, such as playing god, potential dual-use or ecological containment. It also focused on the particular concerns of developing countries about how to avoid further widening the developmental gap with advanced countries (Liu, 2010).In addition to general discussions, there are also sustained transnational communications. For example, one of the first three projects funded by the NSFC was a three-year collaboration on biosafety and risk-assessment frameworks between the Institute of Botany at CAS and the Austrian Organization for International Dialogue and Conflict Management (IDC).Chinese scientists are also keen to increase their involvement in the formulation of international regulations. The CAS and the Chinese Academy of Engineering are engaged with their peer institutions in the UK and the USA to “design more robust frameworks for oversight, intellectual property and international cooperation” (Royal Society, 2009). It is too early to tell what influence China will achieve in this field. Yet, the changing image of the country from an unconcerned wild East to a partner in lively discussions signals a new dynamic in the global development of synthetic biology.Student contests, funding programmes, joint research centres and coordination groups are only a few of the means by which scientists can drive synthetic biology forward in ChinaFrom self-organized participation in iGEM to bottom-up funding and governance initiatives, two features are repeatedly exhibited in the emergence of synthetic biology in China: global resources and international perspectives complement national interests; and the national and cosmopolitan research strengths are mostly instigated at the grass-roots level. During the process of introducing, developing and reflecting on synthetic biology, many formal or informal, provisional or long-term alliances have been established from the bottom up. Student contests, funding programmes, joint research centres and coordination groups are only a few of the means by which scientists can drive synthetic biology forward in China.However, the inputs of different social actors has not led to disintegration of the field into an array of individualized pursuits, but has transformed it into collective synergies, or the big-question approach. Underlying the diverse efforts of Chinese scientists is a sense of ‘inclusiveness'', or the idea of bringing together previously detached research expertise. Thus, the big-question strategy cannot be interpreted as just another nationally organized agenda in response to global scientific advancements. Instead, it represents a more intricate development path corresponding to how contemporary research evolves on the ground.In comparison to the increasingly visible grass-roots efforts, the role of the Chinese government seems relatively small at this stageIn comparison to the increasingly visible grass-roots efforts, the role of the Chinese government seems relatively small at this stage. Government input—such as the potential stewardship of the MOST in directing a big-question approach or long-term funding—remain important; the scientists who were interviewed expend a great deal of effort to attract governmental participation. Yet, China'' experience highlights that the key to comprehending regional scientific capacity lies not so much in what the government can do, but rather in what is taking place in laboratories. It is important to remember that Chinese iGEM victories, collaborative synthetic-biology projects and ethical discussions all took place before the government became involved. Thus, to appreciate fully the dynamics of an emerging science, it might be necessary to focus on what is formulated from the bottom up.The experience of China in synthetic biology demonstrates the power of grass-roots, cross-border engagement to promote contemporary researchThe experience of China in synthetic biology demonstrates the power of grass-roots, cross-border engagement to promote contemporary research. More specifically, it is a result of the commitment of Chinese scientists to incorporating national and international resources, actors and social concerns. For practical reasons, the national organization of research, such as through the big-question approach, might still have an important role. However, synthetic biology might be not only a mosaic of national agendas, but also shaped by transnational activities and scientific resources. What Chinese scientists will collectively achieve remains to be seen. Yet, the emergence of synthetic biology in China might be indicative of a new paradigm for how research practices can be introduced, normalized and regulated.  相似文献   

4.
Biopedagogy     
The world is changing fast and teachers are struggling to familiarize children and young people with the norms and values of society. Biopedagogy—the biology behind pedagogy approaches—might provide some insights and guidance.Humans are the only animals that are subject to cumulative cultural evolution. Biological evolution would be too slow to allow for the invention and continuous improvement of complex artefacts, both material and abstract. One of the mechanisms that makes it possible for humans to pass on social, cultural and technological advances over generations is teaching. Some scholars believe that teaching itself might be an exclusive human trait; others argue that teaching is more prevalent in nature, if it is defined as cooperative behaviour that promotes learning, independent of the mental states and cognitive intentions [1].The science, art and profession of teaching is known as ‘pedagogy'' and its biological basis might well be dubbed ‘biopedagogy''. Pedagogy embeds the learner into a particular culture by exposing the developing mind to cultural values and practices. Human teaching represents two disparate but closely linked activities—education and instruction. To ‘educate'' means to unfold the latent potential of the learner and to cultivate human nature by promoting impulses that conform to a culture and inhibiting those that contradict it. ‘Instruction'' is the provision of knowledge and skills.According to the ethologist Konrad Lorenz, human nature, which is a result of biological evolution, functions as the “inborn schoolmaster” [2] by both allowing and constraining the learning of all the products of cultural evolution. Lorenz received the Nobel Prize for his discovery of ‘imprinting'': the irreversible, life-long fixation of a response to a situation encountered by an organism during development. Imprinting is not specific to humans, but humans have evolved, along with the formation of the central nervous system, more sophisticated mental organs that we call the social brain, the group mind and the darwinian soul. As the physical development of an organism proceeds in stages, so does the mental development. Some of these stages represent critical periods that are particularly sensitive to imprinting. The rapid acquisition of the mother tongue for instance apparently represents such a specific period in human development.According to Lorenz, during and shortly after puberty, humans are prone to a specific kind of imprinting from a culture and its abstract norms and values, driven by a need to become members of a reference group striving for a common ideal [2]. We might call this developmental stage of humans the ‘second (or ideational) imprinting''. This imprinting presupposes a stable society with firmly established norms and values and, in turn, it serves to ensure that stability. The British neuroscientist Sarah-Jayne Blakemore [3] corroborated that the human brain undergoes protracted development and demonstrated that adolescence, in particular, represents a period during which the neuronal basis of the social brain reorganizes. This provides opportunities, but also imposes great responsibility to high-school and university teachers.Jan Amos Comenius, a seventeenth century Moravian educator, already suggested that the mastery of teaching consists in recognizing stages of mental development in which a student is prepared and eager to learn stage-specific knowledge spontaneously. In his view, a teacher is more similar to a gardener, who gives plants care and nutrients to allow them to develop, grow and flourish. Comenius also anticipated the crucial role of positive emotions in pedagogy. His commandment Scola ludus—The School of Play—expresses the fact that teaching and learning can, and should be, associated with pleasure and joy by both teachers and students.Human nature evolved during the Pleistocene about 1.8 million to 10,000 years ago to cope with a hunter–gatherer lifestyle. Yet, modern humans live in and adapt to vastly different environments created by cultural evolution. Apparently, the human genetic outfit is highly versatile and encompasses abstract ‘cultural loci''. Such a cultural locus is an ‘empty slot'' that functions only when it is filled with a meme from the cultural environment; memes would be akin to ‘alleles'' that are specific to a particular cultural locus. This cross-talk of biology and culture makes humans symbolic animals—our social brain allows us to behave altruistically towards ‘symbolic kin'' with whom we share no genetic relationship; and our group mind embraces not only our relatives and friends, but also our tribe or nation and possibly humanity as a whole. Education, and in particular imprinting, essentially determines the extent, quality and scope of this deployment.A developing child has to pass all crucial periods of learning successfully to become a mature human—and humane—being. As Lorenz noted, once a sensitive period has elapsed and the opportunity to learn has been missed, the ability to catch up is considerably reduced or irreversibly lost. We live in a time when cultural values and norms are rapidly changing, often within less than a generation. The ideational imprinting of developing adolescents becomes a problem when the traditional role of family and school is displaced by new social forces such as the internet, Facebook, Twitter and the blogosphere. How to preclude that young people do not develop as persons with stunted social brains, with narrow group minds attuned to fleeting reference groups and with fragmented darwinian souls, and how to promote the development of strong personalities is a challenge for the education of the twenty-first century.  相似文献   

5.
Paul van Helden 《EMBO reports》2012,13(11):942-942
We tend to think in black and white terms of good versus bad alleles and their meaning for disease. However, in doing so, we ignore the potential importance of heterozygous alleles.The structure and function of any protein is determined by its amino acid sequence. Thus, the substitution of one amino acid for another can alter the activity of a protein or its function. Mutations—or rather, polymorphism, once they become fixed in the population—can be deleterious, such that the altered protein is no longer able to fulfil its role with potentially devastating effects on the cell. Rarely, they can improve protein function and cell performance. In either case, any changes in the amino acid sequence, whether they affect only one amino acid or larger parts of the protein, are encoded by polymorphisms in the nucleotide sequence of that protein''s gene. For any given polymorphism, diploid organisms with two sets of chromosomes can therefore exist in either a heterozygous state or one of two homozygous states. When the polymorphism is rare, most individuals are homozygous for the ‘wild-type'' state, some individuals are heterozygous and a few are homozygous for the rare polymorphic variant. Conversely, if the polymorphism occurs in 50% of the alleles, the heterozygous state is common.At first glance, the deleterious homozygous state seems to be something that organisms try to avoid: close relatives usually do not breed, probably to prevent the homozygous accumulation of deleterious alleles. Thus, human cultural norms, founded in our biology, actively select for heterozygosity as many civilizations and societies regard incest as a social taboo. The fields of animal husbandry and conservation biology are littered with information about the significant positive correlation between genetic diversity, evolutionary advantage and fitness [1]. In sexually reproducing organisms, heterozygosity is generally regarded as ‘better'' in terms of adaptability and evolutionary advantage.Why then do we seldom, if ever, regard allelic heterozygosity as an advantage when it comes to genes linked with health and disease? Perhaps it is because we tend to distinguish between the ‘good'' allele, the ‘bad'' allele and the ‘ugly'' heterozygote—since it is burdened with one ‘bad'' allele. Maybe this attitude is a remnant of the outdated ‘one gene, one disease'' model, or of the early studies on inheritable diseases that focused on monogenic or autosomal-dominant genetic disorders. Even modern genetics almost always assigns ‘risk'' to an allele that is associated with a health condition or disadvantaged phenotype; clearly, then, the one homozygous state must have an advantage—sometimes referred to as wild-type—but the heterozygote is often ignored altogether.Maybe we also shun heterozygosity because it is hard to prove, beyond a few examples, that it might offer advantage. A 2010 paper published in Cell claimed that heterozygosity of the lth4A locus conveys protection against tuberculosis [2]. There is a mechanistic basis for the claim: lth4A encodes leukotriene A4 hydrolase, which is the final catalyst to synthesize leukotriene B4, an efficient pro-inflammatory eicosanoid. However, an extensive case–control study could not confirm the association between heterozygosity and protection against tuberculosis [3]. Therefore, many in the field dismiss the prior claim to protection conferred by the heterozygous state.Yet, we know that most biochemical and physiological processes are highly complex systems that involve multiple, interlinked steps with extensive control and feedback mechanisms. Heterozygosity might be one strategy by which an organism maintains flexibility, as it provides more than one allele to fall back on, should conditions change. We may therefore hypothesize that heterozygosity can be either a risk or an advantage, depending on the penetrance or dominance of the alleles. Indeed, there are a few cases in which heterozygosity confers some advantage. For example, individuals who are homozygous for the CCR5 deletion polymorphism (D32/D32) are protected against HIV1 infection, whereas CCR5/D32 heterozygotes have a slower progression to acquired immunodeficiency syndrome (AIDS). In sickle-cell anaemia, heterozygotes have a protective advantage against malaria, whereas the homozygotes either lack protection or suffer health consequences. Thus, although heterozygosity might not create a general fitness advantage, it is advantageous under certain specific conditions, namely the presence of the malaria parasite.In most aspects of life, there are few absolutes and many shades of grey. The ‘normal'' range of parameters in medicine is a clear example of this: optimal functioning of the relevant physiological processes depends on levels that are ‘just right''. As molecular and genetic research tackles the causes and risk factors of complex diseases, we may perhaps find more examples of how heterozygosity at the genetic level conveys health advantages in humans. As the above example regarding tuberculosis indicates, it is difficult to demonstrate any advantage of the heterozygous state. We simply need to be receptive to such possibilities, and improve and reconcile our understanding of allelic diversity and heterozygosity. Researchers working on human disease could benefit from the insights of evolutionary biologists and breeders, who are more appreciative of the heterozygous state.  相似文献   

6.
Cockell CS 《EMBO reports》2011,12(3):181-181
Our ability to disrupt habitats and manipulate living organisms requires a discussion of the ethics of microbiology, even if we argue that microbes themselves have no rights.Synthetic biology and the increasing complexity of molecular biology have brought us to the stage at which we can synthesize new microorganisms. This has generated pressing questions about whether these new organisms have any place in our system of ethics and how we should treat them.The idea that microbes might have some moral claims on us beyond their practical uses or instrumental value is not a new question. Microbiologist Bernard Dixon (1976) presciently asked whether it was ethical to take the smallpox virus to extinction at the height of the attempts of the World Health Organization in the 1970s to eradicate it. There is no unambiguous answer. Today, we might still ask this question, but we might extend it to ask whether the destruction or extinction of a synthetic microbe that was made by humans is also ethically questionable or is such an entity—in that it is designed—more like a machine, which we have no compunction in terminating? Would two lethal pathogens, one of them synthetic and one of them natural, but otherwise identical, command the same moral claims?In a colloquial way, we might ask whether microbes have rights. In previous papers (Cockell, 2004) I have discussed the ‘rights'' of microbes and further explored some issues about the ethics we apply to them (Cockell, 2008). Julian Davies, in a recent opinion article in EMBO reports (Davies, 2010) described my assertion that they should have constitutional rights as ‘ridiculous''. Although I did suggest that environmental law could be changed to recognize the protection of microbial ecosystems—which would imply statutory rights or protection—nowhere have I claimed that microbes should have ‘constitutional'' rights. Nevertheless, this misattribution provides a useful demonstration of the confusion that exists about exactly how we should treat microbes.Few people are in any doubt that microbes should be conserved for their direct uses to humans, for example, in food and drug production, and their indirect uses such as the crucial role they have in the health of ecosystems. Indeed, these motivations can be used to prioritize microbial conservation and protection efforts (Cockell & Jones, 2009). The crucial question is whether microbes have ‘intrinsic value'' beyond their practical uses. If the answer is ‘no'', then we should have no guilt about deliberately driving microbes to extinction for our benefit. However, there are people who feel uneasy with this conclusion, a feeling that calls forth more complex ethical questions.The question is whether microbes have some sort of ‘interests'' that make demands on our treatment of them that go beyond a mere utilitarian calculation. These arguments themselves question what we define as ‘interests'' and whether interests make demands on us. A microbe has no future plans or thought processes; the sorts of interests that are accepted as being of sufficient scope to place demands on our treatment of other human beings, for instance. However, microbes do have biological interests. A halophilic microbe might eventually die if it is dropped into freshwater. Does our knowledge of what is in the biological interests of a microbe mean that we must show it any consideration beyond practical uses? The answer is not obviously negative (Taylor, 1981), but even if we decide that it is, this does not let us off the hook quite yet.There are other intrinsic value arguments that are more obscure, particularly those around the notion of ‘respect''; the idea that we should show empathy towards the trajectory, however deterministic, of other life forms. These unquantifiable and controversial arguments might, nevertheless, partly explain any unease that we have in watching a group of people smash up and destroy some exquisite microbial mats, just because they were bored.Clearly, human instrumental needs do trump microbes at some level. If they did not, we could not use bleach in our houses, an absurd end-point raised in a 1970s science fiction story that explored the futuristic ramifications of full microbial rights, in which household bleaches and deodorants are banned (Patrouch, 1977).However, we should not be so quick to ridicule ideas about microbial ethics and rights. Although it might be true that phages kill a large percentage of the bacterial population of the world every few days, as Julian Davies points out, human society has achieved an unprecedented capacity for destruction and creation. Our ability to poison and disrupt habitats has been unquantified, with respect to the loss of microbial species. Both synthetic biology and bioterrorism raise the spectre of creating new organisms, including pathogens, which we might need to control or deliberately pursue to extinction. Dixon''s dilemma about the smallpox virus, raised more than 30 years ago, has become an urgent point of discussion in the ethics of molecular biology and microbiology.  相似文献   

7.
Denis Duboule 《EMBO reports》2010,11(7):489-489
Where is ‘evo-devo'' going and how will it get there? Denis Duboule analyses the fields of evolution and development and argues that their current marriage is likely a transitory affair.In his inspiring book Ontogeny and Phylogeny (1977), the late Stephen J. Gould explained why developmental biology and evolution, two essential domains of the life sciences, had diverged during the course of the twentieth century; both disciplines had to reach independently a platform of mutual understanding, a theoretical framework wherein concepts are understood and accepted by both parties. A step towards this goal was achieved in the 1980s with the discovery that animals not only share similar ‘developmental genes'', but also more integrated structural and functional aspects of their ontogenies (McGinnis et al, 1984; Akam, 1989). While these advances opened the door towards a molecular understanding of development, the analyses of gene expression in various species also allowed for the establishment of correlations between genetic activities and evolving forms.This comparative approach triggered the emergence of novel animal models and generated a portfolio of concepts, which now-a-days form the basis of a discipline sometimes referred to as ‘evo-devo'' (Wallace, 2002; Carroll, 2005; de Robertis, 2008). The frontiers of this field, however, are not clearly defined. Evo-devo research extends from simply ‘PCRing'' a trendy gene from a weird animal, up to the most sophisticated molecular genetic approaches dealing with the evolution of gene function and regulation. Yet the experiments are always within the general context of homology, as understood by using either morphological, functional or regulatory criteria, indiscriminately.With our improved knowledge of the mechanisms underlying animal development, we can now address the question of natural variation; we have learnt, for instance, that rather limited sets of genes and signalling pathways are used over and over again, hence the development of most organs or structures relies on comparable rules. This, in turn, implies that developing systems have highly constrained roadmaps, the modifications of which lead generally to pleiotropic effects (Duboule & Wilkins 1998; Kirschner & Gerhart, 2005). This natural parsimony in the use of genetic tools makes it sometimes difficult to infer a conservation in function from the mere conservation of gene expression patterns—for instance, between two evolutionary distant animals—and thus calls for a deeper level of conservation to ascertain such phylogenetic relationships.This issue can be addressed either by a thorough understanding of those regulations at work, assuming that a conservation of regulatory circuits demonstrates a common phylogenetic history, or by a large survey of various species, should we accept that a robust association between gene expression and a particular trait bears an evolutionary meaning. The latter point raises the paradox of model systems: that is, whether general conclusions can be extracted from given biological items, which themselves were often chosen for study owing to particularly well-adapted features, rather than for their elusive paradigmatic value. In other words, will we ever understand the full set of core principles by working exclusively with adaptive traits that intrinsically tend to distract from these rules? While this issue is somewhat theoretical, the popular idea that some species can display advantages over others, in terms of experimental benefit, indicates clearly that such questions have not yet been discussed in sufficient depth.The lack of a clear definition of what evo-devo covers as a discipline is echoed by the difficulty to elaborate a commonly accepted set of guidelines, mostly owing to the conflicting ménage between developmental geneticists on the one hand, with their mindset inherited from T. Morgan and H. Spemann, and population (evolutionary) geneticists on the other hand, the direct descents of the new synthesis. In fact, we face a modern version of the classical dichotomy between variation—the ‘how'' question—and selection—the ‘why'' question—and we may wonder how long this productive relationship will last. While it might consolidate itself and lead to an integrated theory of evolution that includes the emerging mechanistic side, it could well split again into divergent trajectories, like a comet that returns closer to a planet every hundred years to fill itself with concepts and energy before leaving again for yet another journey.Evo-devo is arguably a transitory discipline. We are witnessing the emergence of a new developmental biology, relying on high-throughput approaches, systems analyses and modelling to use gene (information) clusters, or even full genomes, as we currently use single genes. The accompanying shift in the required competencies—for example, bioinformatics, physics and maths—although of great interest mechanistically speaking, does not necessarily strengthen the link with the genetic framework of evolution. Also, we should remember that evolution and development are disciplines built on different epistemological grounds, which bring to their fusion an unstable equilibrium. Development is a science of recurrence; based on the assumption that the same process will happen again, in each generation, leading to results that we can predict. As such, it has a fixed timeframe. Evolution relies on the exact opposite premises; it is by definition a linear process, wherefrom recurrence is impossible. It has no clear timeframe and (so far) no predictable result. The former discipline explains how things happen, the latter how things most likely happened.This theoretical antagonism might nevertheless become obsolete once the mechanisms of development are fully understood and once the computation of various ontogenetic roadmaps will discriminate the possible from the impossible, thus telling us which form could evolve out of a given species. This will primarily concern macro-evolution, as micro-evolutionary phenomena are probably less constrained and, as such, more difficult to anticipate. If this were true, one should be able to predict with some accuracy the few alternative solutions offered to one particular species for the next million years, especially if environmental conditions can also be predicted. In such a scenario, the next rendezvous with the comet will turn evolution into a predictive science. This may indeed take another century.  相似文献   

8.
9.
10.
The differentiation of pluripotent stem cells into various progeny is perplexing. In vivo, nature imposes strict fate constraints. In vitro, PSCs differentiate into almost any phenotype. Might the concept of ‘cellular promiscuity'' explain these surprising behaviours?John Gurdon''s [1] and Shinya Yamanaka''s [2] Nobel Prize involves discoveries that vex fundamental concepts about the stability of cellular identity [3,4], ageing as a rectified path and the differences between germ cells and somatic cells. The differentiation of pluripotent stem cells (PSCs) into progeny, including spermatids [5] and oocytes [6], is perplexing. In vivo, nature imposes strict fate constraints. Yet in vitro, reprogrammed PSCs liberated from the body government freely differentiate into any phenotype—except placenta—violating even somatic cell against germ cell segregations. Albeit that it is anthropomorphic, might the concept of ‘cellular promiscuity'' explain these surprising behaviours?Fidelity to one''s differentiated state is nearly universal in vivo—even cancers retain some allegiance. Appreciating the mechanisms in vitro that liberate reprogrammed cells from the numerous constraints governing development in vivo might provide new insights. Similarly to highway guiderails, a range of constraints preclude progeny cells within embryos and organisms from travelling too far away from the trajectory set by their ancestors. Restrictions are imposed externally—basement membranes and intercellular adhesions; internally—chromatin, cytoskeleton, endomembranes and mitochondria; and temporally by ageing.‘Cellular promiscuity'' was glimpsed previously during cloning; it was seen when somatic cells successfully ‘fertilized'' enucleated oocytes in amphibians [1] and later with ‘Dolly'' [7]. Embryonic stem cells (ESCs) corroborate this. The inner cell mass of the blastocyst cells develops faithfully, but liberation from the trophoectoderm generates pluripotent ESCs in vitro, which are freed from fate and polarity restrictions. These freedom-seeking ESCs still abide by three-dimensional rules as they conform to chimaera body patterning when injected into blastocysts. Yet if transplanted elsewhere, this results in chaotic teratomas or helter-skelter in vitro differentiation—that is, pluripotency.August Weismann''s germ plasm theory, 130 years ago, recognized that gametes produce somatic cells, never the reverse. Primordial germ cell migrations into fetal gonads, and parent-of-origin imprints, explain how germ cells are sequestered, retaining genomic and epigenomic purity. Left uncontaminated, these future gametes are held in pristine form to parent the next generation. However, the cracks separating germ and somatic lineages in vitro are widening [5,6]. Perhaps, they are restrained within gonads not for their purity but to prevent wild, uncontrolled misbehaviours resulting in germ cell tumours.The ‘cellular promiscuity'' concept regarding PSCs in vitro might explain why cells of nearly any desired lineage can be detected using monospecific markers. Are assays so sensitive that rare cells can be detected in heterogeneous cultures? Certainly population heterogeneity is considered for transplantable cells—dopaminergic neurons and islet cells—compared with applications needing few cells—sperm and oocytes. This dilemma of maintaining cellular identity in vitro after reprogramming is significant. If not addressed, the value of unrestrained induced PSCs (iPSCs) as reliable models for ‘diseases in a dish'', let alone for subsequent therapeutic transplantations, might be diminished. X-chromosome re-inactivation variants in differentiating human PSCs, epigenetic imprint errors and copy number variations are all indicators of in vitro infidelity. PSCs, which are held to be undifferentiated cells, are artefacts after all, as they undergo their programmed development in vivo.If correct, the hypothesis accounts for concerns raised about the inherent genomic and epigenomic unreliability of iPSCs; they are likely to be unfaithful to their in vivo differentiation trajectories due to both the freedom from in vivo developmental programmes, as well as poorly characterized modifications in culture conditions. ‘Memory'' of the PSC''s identity in vivo might need to be improved by using approaches that might not fully erase imprints. Regulatory authorities, including the Food & Drug Administration, require evidence that cultured PSCs do retain their original cellular identity. Notwithstanding fidelity lapses at the organismal level, the recognition that our cells have intrinsic freedom-loving tendencies in vitro might generate better approaches for only partly releasing somatic cells into probation, rather than full emancipation.  相似文献   

11.
The authors of “The anglerfish deception” respond to the criticism of their article.EMBO reports (2012) advanced online publication; doi: 10.1038/embor.2012.70EMBO reports (2012) 13 2, 100–105; doi: 10.1038/embor.2011.254Our respondents, eight current or former members of the EFSA GMO panel, focus on defending the EFSA''s environmental risk assessment (ERA) procedures. In our article for EMBO reports, we actually focused on the proposed EU GMO legislative reform, especially the European Commission (EC) proposal''s false political inflation of science, which denies the normative commitments inevitable in risk assessment (RA). Unfortunately the respondents do not address this problem. Indeed, by insisting that Member States enjoy freedom over risk management (RM) decisions despite the EFSA''s central control over RA, they entirely miss the relevant point. This is the unacknowledged policy—normative commitments being made before, and during, not only after, scientific ERA. They therefore only highlight, and extend, the problem we identified.The respondents complain that we misunderstood the distinction between RA and RM. We did not. We challenged it as misconceived and fundamentally misleading—as though only objective science defined RA, with normative choices cleanly confined to RM. Our point was that (i) the processes of scientific RA are inevitably shaped by normative commitments, which (ii) as a matter of institutional, policy and scientific integrity must be acknowledged and inclusively deliberated. They seem unaware that many authorities [1,2,3,4] have recognized such normative choices as prior matters, of RA policy, which should be established in a broadly deliberative manner “in advance of risk assessment to ensure that [RA] is systematic, complete, unbiased and transparent” [1]. This was neither recognized nor permitted in the proposed EC reform—a central point that our respondents fail to recognize.In dismissing our criticism that comparative safety assessment appears as a ‘first step'' in defining ERA, according to the new EFSA ERA guidelines, which we correctly referred to in our text but incorrectly referenced in the bibliography [5], our respondents again ignore this widely accepted ‘framing'' or ‘problem formulation'' point for science. The choice of comparator has normative implications as it immediately commits to a definition of what is normal and, implicitly, acceptable. Therefore the specific form and purpose of the comparison(s) is part of the validity question. Their claim that we are against comparison as a scientific step is incorrect—of course comparison is necessary. This simply acts as a shield behind which to avoid our and others'' [6] challenge to their self-appointed discretion to define—or worse, allow applicants to define—what counts in the comparative frame. Denying these realities and their difficult but inevitable implications, our respondents instead try to justify their own particular choices as ‘science''. First, they deny the first-step status of comparative safety assessment, despite its clear appearance in their own ERA Guidance Document [5]—in both the representational figure (p.11) and the text “the outcome of the comparative safety assessment allows the determination of those ‘identified'' characteristics that need to be assessed [...] and will further structure the ERA” (p.13). Second, despite their claims to the contrary, ‘comparative safety assessment'', effectively a resurrection of substantial equivalence, is a concept taken from consumer health RA, controversially applied to the more open-ended processes of ERA, and one that has in fact been long-discredited if used as a bottleneck or endpoint for rigorous RA processes [7,8,9,10]. The key point is that normative commitments are being embodied, yet not acknowledged, in RA science. This occurs through a range of similar unaccountable RA steps introduced into the ERA Guidance, such as judgement of ‘biological relevance'', ‘ecological relevance'', or ‘familiarity''. We cannot address these here, but our basic point is that such endless ‘methodological'' elaborations of the kind that our EFSA colleagues perform, only obscure the institutional changes needed to properly address the normative questions for policy-engaged science.Our respondents deny our claim concerning the singular form of science the EC is attempting to impose on GM policy and debate, by citing formal EFSA procedures for consultations with Member States and non-governmental organizations. However, they directly refute themselves by emphasizing that all Member State GM cultivation bans, permitted only on scientific grounds, have been deemed invalid by EFSA. They cannot have it both ways. We have addressed the importance of unacknowledged normativity in quality assessments of science for policy in Europe elsewhere [11]. However, it is the ‘one door, one key'' policy framework for science, deriving from the Single Market logic, which forces such singularity. While this might be legitimate policy, it is not scientific. It is political economy.Our respondents conclude by saying that the paramount concern of the EFSA GMO panel is the quality of its science. We share this concern. However, they avoid our main point that the EC-proposed legislative reform would only exacerbate their problem. Ignoring the normative dimensions of regulatory science and siphoning-off scientific debate and its normative issues to a select expert panel—which despite claiming independence faces an EU Ombudsman challenge [12] and European Parliament refusal to discharge their 2010 budget, because of continuing questions over conflicts of interests [13,14]—will not achieve quality science. What is required are effective institutional mechanisms and cultural norms that identify, and deliberatively address, otherwise unnoticed normative choices shaping risk science and its interpretive judgements. It is not the EFSA''s sole responsibility to achieve this, but it does need to recognize and press the point, against resistance, to develop better EU science and policy.  相似文献   

12.
13.
-Adrenergic receptor activation regulates cardiac myocyte function through the stimulation of cAMP production and subsequent activation of protein kinase A (PKA). Furthermore, muscarinic receptor activation inhibits as well as facilitates these cAMP-dependent effects. However, it has not always been possible to correlate the muscarinic responses with the direct measurement of changes in cellular cAMP activity. Genetically encoded biosensors have recently been developed, making it possible to monitor real-time changes in cAMP and PKA activity at the single cell level. One such biosensor consists of the regulatory and catalytic subunits of PKA labeled with cyan and yellow fluorescent proteins, respectively. Changes in cAMP activity affecting the association of these labeled PKA subunits can be detected as changes in fluorescence resonance energy transfer. In the present study, an adenovirus-based approach was developed to express this recombinant protein complex in adult cardiac myocytes and use it to monitor changes in cAMP activity produced by -adrenergic and muscarinic receptor activation. The biosensor expressed with the use of this system is able to detect changes in cAMP activity produced by physiologically relevant levels of -adrenergic receptor activation without disrupting normal functional responses. It was also possible to directly demonstrate the complex temporal pattern of inhibitory and stimulatory changes in cAMP activity produced by muscarinic receptor activation in these cells. The adenovirus-based approach we have developed should facilitate the use of this biosensor in studying cAMP and PKA-dependent signaling mechanisms in a wide variety of cell types. adenovirus; protein kinase A; phosphodiesterase; L-type Ca2+ channel  相似文献   

14.
Greener M 《EMBO reports》2008,9(11):1067-1069
A consensus definition of life remains elusiveIn July this year, the Phoenix Lander robot—launched by NASA in 2007 as part of the Phoenix mission to Mars—provided the first irrefutable proof that water exists on the Red Planet. “We''ve seen evidence for this water ice before in observations by the Mars Odyssey orbiter and in disappearing chunks observed by Phoenix […], but this is the first time Martian water has been touched and tasted,” commented lead scientist William Boynton from the University of Arizona, USA (NASA, 2008). The robot''s discovery of water in a scooped-up soil sample increases the probability that there is, or was, life on Mars.Meanwhile, the Darwin project, under development by the European Space Agency (ESA; Paris, France; www.esa.int/science/darwin), envisages a flotilla of four or five free-flying spacecraft to search for the chemical signatures of life in 25 to 50 planetary systems. Yet, in the vastness of space, to paraphrase the British astrophysicist Arthur Eddington (1822–1944), life might be not only stranger than we imagine, but also stranger than we can imagine. The limits of our current definitions of life raise the possibility that we would not be able to recognize an extra-terrestrial organism.Back on Earth, molecular biologists—whether deliberately or not—are empirically tackling the question of what is life. Researchers at the J Craig Venter Institute (Rockville, MD, USA), for example, have synthesized an artificial bacterial genome (Gibson et al, 2008). Others have worked on ‘minimal cells'' with the aim of synthesizing a ‘bioreactor'' that contains the minimum of components necessary to be self-sustaining, reproduce and evolve. Some biologists regard these features as the hallmarks of life (Luisi, 2007). However, to decide who is first in the ‘race to create life'' requires a consensus definition of life itself. “A definition of the precise boundary between complex chemistry and life will be critical in deciding which group has succeeded in what might be regarded by the public as the world''s first theology practical,” commented Jamie Davies, Professor of Experimental Anatomy at the University of Edinburgh, UK.For most biologists, defining life is a fascinating, fundamental, but largely academic question. It is, however, crucial for exobiologists looking for extra-terrestrial life on Mars, Jupiter''s moon Europa, Saturn''s moon Titan and on planets outside our solar system.In their search for life, exobiologists base their working hypothesis on the only example to hand: life on Earth. “At the moment, we can only assume that life elsewhere is based on the same principles as on Earth,” said Malcolm Fridlund, Secretary for the Exo-Planet Roadmap Advisory Team at the ESA''s European Space Research and Technology Centre (Noordwijk, The Netherlands). “We should, however, always remember that the universe is a peculiar place and try to interpret unexpected results in terms of new physics and chemistry.”The ESA''s Darwin mission will, therefore, search for life-related gases such as carbon dioxide, water, methane and ozone in the atmospheres of other planets. On Earth, the emergence of life altered the balance of atmospheric gases: living organisms produced all of the Earth'' oxygen, which now accounts for one-fifth of the atmosphere. “If all life on Earth was extinguished, the oxygen in our atmosphere would disappear in less than 4 million years, which is a very short time as planets go—the Earth is 4.5 billion years old,” Fridlund said. He added that organisms present in the early phases of life on Earth produced methane, which alters atmospheric composition compared with a planet devoid of life.Although the Darwin project will use a pragmatic and specific definition of life, biologists, philosophers and science-fiction authors have devised numerous other definitions—none of which are entirely satisfactory. Some are based on basic physiological characteristics: a living organism must feed, grow, metabolize, respond to stimuli and reproduce. Others invoke metabolic definitions that define a living organism as having a distinct boundary—such as a membrane—which facilitates interaction with the environment and transfers the raw materials needed to maintain its structure (Wharton, 2002). The minimal cell project, for example, defines cellular life as “the capability to display a concert of three main properties: self-maintenance (metabolism), reproduction and evolution. When these three properties are simultaneously present, we will have a full fledged cellular life” (Luisi, 2007). These concepts regard life as an emergent phenomenon arising from the interaction of non-living chemical components.Cryptobiosis—hidden life, also known as anabiosis—and bacterial endospores challenge the physiological and metabolic elements of these definitions (Wharton, 2002). When the environment changes, certain organisms are able to undergo cryptobiosis—a state in which their metabolic activity either ceases reversibly or is barely discernible. Cryptobiosis allows the larvae of the African fly Polypedilum vanderplanki to survive desiccation for up to 17 years and temperatures ranging from −270 °C (liquid helium) to 106 °C (Watanabe et al, 2002). It also allows the cysts of the brine shrimp Artemia to survive desiccation, ultraviolet radiation, extremes of temperature (Wharton, 2002) and even toyshops, which sell the cysts as ‘sea monkeys''. Organisms in a cryptobiotic state show characteristics that vary markedly from what we normally consider to be life, although they are certainly not dead. “[C]ryptobiosis is a unique state of biological organization”, commented James Clegg, from the Bodega Marine Laboratory at the University of California (Davies, CA, USA), in an article in 2001 (Clegg, 2001). Bacterial endospores, which are the “hardiest known form of life on Earth” (Nicholson et al, 2000), are able to withstand almost any environment—perhaps even interplanetary space. Microbiologists isolated endospores of strict thermophiles from cold lake sediments and revived spores from samples some 100,000 years old (Nicholson et al, 2000).…life might be not only stranger than we imagine, but also stranger than we can imagineAnother problem with the definitions of life is that these can expand beyond biology. The minimal cell project, for example, in common with most modern definitions of life, encompass the ability to undergo Darwinian evolution (Wharton, 2002). “To be considered alive, the organism needs to be able to undergo extensive genetic modification through natural selection,” said Professor Paul Freemont from Imperial College London, UK, whose research interests encompass synthetic biology. But the virtual ‘organisms'' in computer simulations such as the Game of Life (www.bitstorm.org/gameoflife) and Tierra (http://life.ou.edu/tierra) also exhibit life-like characteristics, including growth, death and evolution—similar to robots and other artifical systems that attempt to mimic life (Guruprasad & Sekar, 2006). “At the moment, we have some problems differentiating these approaches from something biologists consider [to be] alive,” Fridlund commented.…to decide who is first in the ‘race to create life'' requires a consensus definition of lifeBoth the genetic code and all computer-programming languages are means of communicating large quantities of codified information, which adds another element to a comprehensive definition of life. Guenther Witzany, an Austrian philosopher, has developed a “theory of communicative nature” that, he claims, differentiates biotic and abiotic life. “Life is distinguished from non-living matter by language and communication,” Witzany said. According to his theory, RNA and DNA use a ‘molecular syntax'' to make sense of the genetic code in a manner similar to language. This paragraph, for example, could contain the same words in a random order; it would be meaningless without syntactic and semantic rules. “The RNA/DNA language follows syntactic, semantic and pragmatic rules which are absent in [a] random-like mixture of nucleic acids,” Witzany explained.Yet, successful communication requires both a speaker using the rules and a listener who is aware of and can understand the syntax and semantics. For example, cells, tissues, organs and organisms communicate with each other to coordinate and organize their activities; in other words, they exchange signals that contain meaning. Noradrenaline binding to a β-adrenergic receptor in the bronchi communicates a signal that says ‘dilate''. “If communication processes are deformed, destroyed or otherwise incorrectly mediated, both coordination and organisation of cellular life is damaged or disturbed, which can lead to disease,” Witzany added. “Cellular life also interprets abiotic environmental circumstances—such as the availability of nutrients, temperature and so on—to generate appropriate behaviour.”Nonetheless, even definitions of life that include all the elements mentioned so far might still be incomplete. “One can make a very complex definition that covers life on the Earth, but what if we find life elsewhere and it is different? My opinion, shared by many, is that we don''t have a clue of how life arose on Earth, even if there are some hypotheses,” Fridlund said. “This underlies many of our problems defining life. Since we do not have a good minimum definition of life, it is hard or impossible to find out how life arose without observing the process. Nevertheless, I''m an optimist who believes the universe is understandable with some hard work and I think we will understand these issues one day.”Both synthetic biology and research on organisms that live in extreme conditions allow biologists to explore biological boundaries, which might help them to reach a consensual minimum definition of life, and understand how it arose and evolved. Life is certainly able to flourish in some remarkably hostile environments. Thermus aquaticus, for example, is metabolically optimal in the springs of Yellowstone National Park at temperatures between 75 °C and 80 °C. Another extremophile, Deinococcus radiodurans, has evolved a highly efficient biphasic system to repair radiation-induced DNA breaks (Misra et al, 2006) and, as Fridlund noted, “is remarkably resistant to gamma radiation and even lives in the cooling ponds of nuclear reactors.”In turn, synthetic biology allows for a detailed examination of the elements that define life, including the minimum set of genes required to create a living organism. Researchers at the J Craig Venter Institute, for example, have synthesized a 582,970-base-pair Mycoplasma genitalium genome containing all the genes of the wild-type bacteria, except one that they disrupted to block pathogenicity and allow for selection. ‘Watermarks'' at intergenic sites that tolerate transposon insertions identify the synthetic genome, which would otherwise be indistinguishable from the wild type (Gibson et al, 2008).Yet, as Pier Luigi Luisi from the University of Roma in Italy remarked, even M. genitalium is relatively complex. “The question is whether such complexity is necessary for cellular life, or whether, instead, cellular life could, in principle, also be possible with a much lower number of molecular components”, he said. After all, life probably did not start with cells that already contained thousands of genes (Luisi, 2007).…researchers will continue their attempts to create life in the test tube—it is, after all, one of the greatest scientific challengesTo investigate further the minimum number of genes required for life, researchers are using minimal cell models: synthetic genomes that can be included in liposomes, which themselves show some life-like characteristics. Certain lipid vesicles are able to grow, divide and grow again, and can include polymerase enzymes to synthesize RNA from external substrates as well as functional translation apparatuses, including ribosomes (Deamer, 2005).However, the requirement that an organism be subject to natural selection to be considered alive could prove to be a major hurdle for current attempts to create life. As Freemont commented: “Synthetic biologists could include the components that go into a cell and create an organism [that is] indistinguishable from one that evolved naturally and that can replicate […] We are beginning to get to grips with what makes the cell work. Including an element that undergoes natural selection is proving more intractable.”John Dupré, Professor of Philosophy of Science and Director of the Economic and Social Research Council (ESRC) Centre for Genomics in Society at the University of Exeter, UK, commented that synthetic biologists still approach the construction of a minimal organism with certain preconceptions. “All synthetic biology research assumes certain things about life and what it is, and any claims to have ‘confirmed'' certain intuitions—such as life is not a vital principle—aren''t really adding empirical evidence for those intuitions. Anyone with the opposite intuition may simply refuse to admit that the objects in question are living,” he said. “To the extent that synthetic biology is able to draw a clear line between life and non-life, this is only possible in relation to defining concepts brought to the research. For example, synthetic biologists may be able to determine the number of genes required for minimal function. Nevertheless, ‘what counts as life'' is unaffected by minimal genomics.”Partly because of these preconceptions, Dan Nicholson, a former molecular biologist now working at the ESRC Centre, commented that synthetic biology adds little to the understanding of life already gained from molecular biology and biochemistry. Nevertheless, he said, synthetic biology might allow us to go boldly into the realms of biological possibility where evolution has not gone before.An engineered synthetic organism could, for example, express novel amino acids, proteins, nucleic acids or vesicular forms. A synthetic organism could use pyranosyl-RNA, which produces a stronger and more selective pairing system than the natural existent furanosyl-RNA (Bolli et al, 1997). Furthermore, the synthesis of proteins that do not exist in nature—so-called never-born proteins—could help scientists to understand why evolutionary pressures only selected certain structures.As Luisi remarked, the ratio between the number of theoretically possible proteins containing 100 amino acids and the real number present in nature is close to the ratio between the space of the universe and the space of a single hydrogen atom, or the ratio between all the sand in the Sahara Desert and a single grain. Exploring never-born proteins could, therefore, allow synthetic biologists to determine whether particular physical, structural, catalytic, thermodynamic and other properties maximized the evolutionary fitness of natural proteins, or whether the current protein repertoire is predominately the result of chance (Luisi, 2007).In the final analysis, as with all science, deep understanding is more important than labelling with words.“Synthetic biology also could conceivably help overcome the ‘n = 1 problem''—namely, that we base biological theorising on terrestrial life only,” Nicholson said. “In this way, synthetic biology could contribute to the development of a more general, broader understanding of what life is and how it might be defined.”No matter the uncertainties, researchers will continue their attempts to create life in the test tube—it is, after all, one of the greatest scientific challenges. Whether or not they succeed will depend partly on the definition of life that they use, though in any case, the research should yield numerous insights that are beneficial to biologists generally. “The process of creating a living system from chemical components will undoubtedly offer many rich insights into biology,” Davies concluded. “However, the definition will, I fear, reflect politics more than biology. Any definition will, therefore, be subject to a lot of inter-lab political pressure. Definitions are also important for bioethical legislation and, as a result, reflect larger politics more than biology. In the final analysis, as with all science, deep understanding is more important than labelling with words.”  相似文献   

15.
16.
17.
18.
Rinaldi A 《EMBO reports》2012,13(1):24-27
Does the spin of an electron allow birds to see the Earth''s magnetic field? Andrea Rinaldi investigates the influence of quantum events in the biological world.The subatomic world is nothing like the world that biologists study. Physicists have struggled for almost a century to understand the wave–particle duality of matter and energy, but many questions remain unanswered. That biological systems ultimately obey the rules of quantum mechanics might be self-evident, but the idea that those rules are the very basis of certain biological functions has needed 80 years of thought, research and development for evidence to begin to emerge (Sidebar A).

Sidebar A | Putting things in their place

Although Erwin Schrödinger (1887–1961) is often credited as the ‘father'' of quantum biology, owing to the publication of his famous 1944 book, What is Life?, the full picture is more complex. While other researchers were already moving towards these concepts in the 1920s, the German theoretical physicist Pascual Jordan (1902–1980) was actually one of the first to attempt to reconcile biological phenomena with the quantum revolution that Jordan himself, working with Max Born and Werner Heisenberg, largely ignited. “Pascual Jordan was one of many scientists at the time who were exploring biophysics in innovative ways. In some cases, his ideas have proven to be speculative or even fantastical. In others, however, his ideas have proven to be really ahead of their time,” explained Richard Beyler, a science historian at Portland State University, USA, who analysed Jordan''s contribution to the rise of quantum biology (Beyler, 1996). “I think this applies to Jordan''s work in quantum biology as well.”Beyler also remarked that some of the well-known figures of molecular biology''s past—Max Delbrück is a notable example—entered into their studies at least in part as a response or rejoinder to Jordan''s work. “Schrödinger''s book can also be read, on some level, as an indirect response to Jordan,” Beyler said.Jordan was certainly a complex personality and his case is rendered more complicated by the fact that he explicitly hitched his already speculative scientific theories to various right-wing political philosophies. “During the Nazi regime, for example, he promoted the notion that quantum biology served as evidence for the naturalness of dictatorship and the prospective death of liberal democracy,” Beyler commented. “After 1945, Jordan became a staunch Cold Warrior and saw in quantum biology a challenge to philosophical and political materialism. Needless to say, not all of his scientific colleagues appreciated these propagandistic endeavors.”Pascual Jordan [pictured above] and the dawn of quantum biology. From 1932, Jordan started to outline the new field''s background in a series of essays that were published in journals such as Naturwissenschaften. An exposition of quantum biology is also encountered in his book Die Physik und das Geheimnis des organischen Lebens, published in 1941. Photo courtesy of Luca Turin.Until very recently, it was not even possible to investigate whether quantum phenomena such as coherence and entanglement could play a significant role in the function of living organisms. As such, researchers were largely limited to computer simulations and theoretical experiments to explain their observations (see A quantum leap in biology, www.emboreports.org). Recently, however, quantum biologists have been making inroads into developing methodology to measure the degree of quantum entanglement in light-harvesting systems. Their breakthrough has turned once ephemeral theories into solid evidence, and has sparked the beginning of an entirely new discipline.How widespread is the direct relevance of quantum effects in nature is hard to say and many scientists suspect that there are only a few cases in which quantum mechanics have a crucial role. However, interest in the field is growing and researchers are looking for more examples of quantum-dependent biological systems. In a way, quantum biology can be viewed as a natural evolution of biophysics, moving from the classical to the quantum, from the atomic to the subatomic. Yet the discipline might prove to be an even more intimate and further-reaching marriage that could provide a deeper understanding of things such as protein energetics and dynamics, and all biological processes where electrons flow.Recently […] quantum biologists have been making inroads into developing methodology to measure the degree of quantum entanglement in light-harvesting systemsAmong the biological systems in which quantum effects are believed to have a crucial role is magnetoreception, although the nature of the receptors and the underlying biophysical mechanisms remain unknown. The possibility that organisms use a ferromagnetic material (magnetite) in some cases has received some confirmation, but support is growing for the explanation lying in a chemical detection mechanism with quantum mechanical properties. This explanation posits a chemical compass based on the light-triggered production of a radical pair—a pair of molecules each with an unpaired electron—the spins of which are entangled. If the products of the radical pair system are spin-dependent, then a magnetic field—like the geomagnetic one—that affects the direction of spin will alter the reaction products. The idea is that these reaction products affect the sensitivity of light sensors in the eye, thus allowing organisms to ‘see'' magnetic fields.The research comes from a team led by Thorsten Ritz at the University of California Irvine, USA, and other groups, who have suggested that the radical pair reaction takes place in the molecule cryptochrome. Cryptochromes are flavoprotein photoreceptors first identified in the model plant Arabidopsis thaliana, in which they play key roles in growth and development. More recently, cryptochromes have been found to have a role in the circadian clock of fruit flies (Ritz et al, 2010) and are known to be present in migratory birds. Intriguingly, magnetic fields have been shown to have an effect on both Arabidopsis seedlings, which respond as though they have been exposed to higher levels of blue light, and Drosophila, in which the period length of the clock is lengthened, mimicking the effect of increased blue light signal intensity on cryptochromes (Ahmad et al, 2007; Yoshii et al, 2009).“The study of quantum effects in biological systems is a rapidly broadening field of research in which intriguing phenomena are yet to be uncovered and understood”Direct evidence that cryptochrome is the avian magnetic compass is currently lacking, but the molecule does have some features that make its candidacy possible. In a recent review (Ritz et al, 2010), Ritz and colleagues discussed the mechanism by which cryptochrome might form radical pairs. They argued that “Cryptochromes are bound to a light-absorbing flavin cofactor (FAD) which can exist in three interconvertable [sic] redox forms: (FAD, FADH, FADH),” and that the redox state of FAD is light-dependent. As such, both the oxidation and reduction of the flavin have radical species as intermediates. “Therefore both forward and reverse reactions may involve the formation of radical pairs” (Ritz et al, 2010). Although speculative, the idea is that a magnetic field could alter the spin of the free electrons in the radical pairs resulting in altered photoreceptor responses that could be perceived by the organism. “Given the relatively short time from the first suggestion of cryptochrome as a magnetoreceptor in 2000, the amount of studies from different fields supporting the photo-magnetoreceptor and cryptochrome hypotheses […] is promising,” the authors concluded. “It suggests that we may be only one step away from a true smoking gun revealing the long-sought after molecular nature of receptors underlying the 6th sense and thus the solution of a great outstanding riddle of sensory biology.”Research into quantum effects in biology took off in 2007 with groundbreaking experiments from Graham Fleming''s group at the University of California, Berkeley, USA. Fleming''s team were able to develop tools that allowed them to excite the photosynthetic apparatus of the green sulphur bacterium Chlorobium tepidum with short laser pulses to demonstrate that wave-like energy transfer takes place through quantum coherence (Engel et al, 2007). Shortly after, Martin Plenio''s group at Ulm University in Germany and Alán Aspuru-Guzik''s team at Harvard University in the USA simultaneously provided evidence that it is a subtle interplay between quantum coherence and environmental noise that optimizes the performance of biological systems such as the photosynthetic machinery, adding further interest to the field (Plenio & Huelga, 2008; Rebentrost et al, 2009). “The recent Quantum Effects in Biological Systems (QuEBS) 2011 meeting in Ulm saw an increasing number of biological systems added to the group of biological processes in which quantum effects are suspected to play a crucial role,” commented Plenio, one of the workshop organizers; he mentioned the examples of avian magnetoreception and the role of phonon-assisted tunnelling to explain the function of the sense of smell (see below). “The study of quantum effects in biological systems is a rapidly broadening field of research in which intriguing phenomena are yet to be uncovered and understood,” he concluded.“The area of quantum effects in biology is very exciting because it is pushing the limits of quantum physics to a new scale,” Yasser Omar from the Technical University of Lisbon, Portugal commented. ”[W]e are finding that quantum coherence plays a significant role in the function of systems that we previously thought would be too large, too hot—working at physiological temperatures—and too complex to depend on quantum effects.”Another growing focus of quantum biologists is the sense of smell and odorant recognition. Mainstream researchers have always favoured a ‘lock-and-key'' mechanism to explain how organisms detect and distinguish different smells. In this case, the identification of odorant molecules relies on their specific shape to activate receptors on the surface of sensory neurons in the nasal epithelium. However, a small group of ‘heretics'' think that the smell of a molecule is actually determined by intramolecular vibrations, rather than by its shape. This, they say, explains why the shape theory has so far failed to explain why different molecules can have similar odours, while similar molecules can have dissimilar odours. It also goes some way to explaining how humans can manage with fewer than 400 smell receptors.…determining whether quantum effects have a role in odorant recognition has involved assessing the physical violations of such a mechanism […] and finding that, given certain biological parameters, there are noneA recent study in Proceedings of the National Academy of Sciences USA has now provided new grist for the mill for ‘vibrationists''. Researchers from the Biomedical Sciences Research Center “Alexander Fleming”, Vari, Greece—where the experiments were performed—and the Massachusetts Institute of Technology (MIT), USA, collaborated to replace hydrogen with deuterium in odorants such as acetophenone and 1-octanol, and asked whether Drosophila flies could distinguish the two isotopes, which are identically shaped but vibrate differently (Franco et al, 2011). Not only were the flies able to discriminate between the isotopic odorants, but when trained to discriminate against the normal or deuterated isotopes of a compound, they could also selectively avoid the corresponding isotope of a different odorant. The findings are inconsistent with a shape-only model for smell, the authors concluded, and suggest that flies can ‘smell molecular vibrations''.“The ability to detect heavy isotopes in a molecule by smell is a good test of shape and vibration theories: shape says it should be impossible, vibration says it should be doable,” explained Luca Turin from MIT, one of the study''s authors. Turin is a major proponent of the vibration theory and suggests that the transduction of molecular vibrations into receptor activation could be mediated by inelastic electron tunnelling (Fig 1; see also The scent of life, www.emboreports.org). “The results so far had been inconclusive and complicated by possible contamination of the test odorants with impurities,” Turin said. “Our work deals with impurities in a novel way, by asking flies whether the presence of deuterium isotope confers a common smell character to odorants, much in the way that the presence of -SH in a molecule makes it smell ‘sulphuraceous'', regardless of impurities. The flies'' answer seems to be ‘yes''.”Open in a separate windowFigure 1Diagram of a vibration-sensing receptor using an inelastic electron tunnelling mechanism. An odorant—here benzaldehyde—is depicted bound to a protein receptor that includes an electron donor site at the top left to which an electron—blue sphere—is bound. The electron can tunnel to an acceptor site at the bottom right while losing energy (vertical arrow) by exciting one or more vibrational modes of the benzaldehyde. When the electron reaches the acceptor, the signal is transduced via a G-protein mechanism, and the olfactory stimulus is triggered. Credit: Luca Turin.One of the study''s Greek co-authors, Efthimios Skoulakis, suggested that flies are better suited than humans at doing this experiment for a couple of reasons. “[The flies] seem to have better acuity than humans and they cannot anticipate the task they will be required to complete (as humans would), thus reducing bias in the outcome,” he said. “Drosophila does not need to detect deuterium per se to survive and be reproductively successful, so it is likely that detection of the vibrational difference between such a compound and its normal counterpart reflects a general property of olfactory systems.”The question of whether quantum mechanics really plays a non-trivial role in biology is still hotly debated by physicists and biologists alikeJennifer Brookes, a physicist at University College London, UK, explained that recent advances in determining whether quantum effects have a role in odorant recognition has involved assessing the physical violations of such a mechanism in the first instance, and finding that, given certain biological parameters, there are none. “The point being that if nature uses something like the quantized vibrations of molecules to ‘measure'' a smell then the idea is not—mathematically, physically and biologically—as eccentric as it at first seems,” she said. Moreover, there is the possibility that quantum mechanics could play a much broader role in biology than simply underpinning the sense of smell. “Odorants are not the only small molecules that interact unpredictably with large proteins; steroid hormones, anaesthetics and neurotransmitters, to name a few, are examples of ligands that interact specifically with special receptors to produce important biological processes,” Brookes wrote in a recent essay (Brookes, 2010).The question of whether quantum mechanics really plays a non-trivial role in biology is still hotly debated by physicists and biologists alike. “[A] non-trivial quantum effect in biology is one that would convince a biologist that they needed to take an advanced quantum mechanics course and learn about Hilbert space and operators etc., so that they could understand the effect,” argued theoretical quantum physicists Howard Wiseman and Jens Eisert in their contribution to the book Quantum Aspects of Life (Wiseman & Eisert, 2008). In their rational challenge to the general enthusiasm for a quantum revolution in biology, Wiseman and Eisert point out that a number of “exotic” and “implausible” quantum effects—including a quantum life principle, quantum computing in the brain, quantum computing in genetics, and quantum consciousness—have been suggested and warn researchers to be cautious of “ideas that are more appealing at first sight than they are realistic” (Wiseman & Eisert, 2008).“One could easily expect many more new exciting ideas and discoveries to emerge from the intersection of two major areas such as quantum physics and biology”Keeping this warning in mind, the view of life from a quantum perspective can still provide a deeper insight into the mechanisms that allow living organisms to thrive without succumbing to the increasing entropy of their environment. But does quantum biology have practical applications? “The investigation of the role of quantum physics in biology is fascinating because it could help explain why evolution has favoured some biological designs, as well as inspire us to develop more efficient artificial devices,” Omar said. The most often quoted examples of such devices are solar collectors that would use efficient energy transport mechanisms inspired by the quantum proficiency of natural light-harvesting systems, and quantum computing. But there is much more ahead. In 2010, the Pentagon''s cutting-edge research branch, DARPA (Defense Advanced Research Projects Agency, USA), launched a solicitation for innovative proposals in the area of quantum effects in a biological environment. “Proposed research should establish beyond any doubt that manifestly quantum effects occur in biology, and demonstrate through simulation proof-of-concept experiments that devices that exploit these effects could be developed into biomimetic sensors,” states the synopsis (DARPA, 2010). This programme will thus look explicitly at photosynthesis, magnetic field sensing and odour detection to lay the foundations for novel sensor technologies for military applications.Clearly a number of civil needs could also be fulfilled by quantum-based biosensors. Take, for example, the much sought-after ‘electronic nose'' that could replace the use of dogs to find drugs or explosives, or could assess food quality and safety. Such a device could even be used to detect cancer, as suggested by a recent publication from a Swedish team of researchers who reported that ovarian carcinomas emit a different array of volatile signals to normal tissue (Horvath et al, 2010). “Our goal is to be able to screen blood samples from apparently healthy women and so detect ovarian cancer at an early stage when it can still be cured,” said the study''s leading author György Horvath in a press release (University of Gothenburg, 2010).Despite its already long incubation time, quantum biology is still in its infancy but with an intriguing adolescence ahead. “A new wave of scientists are finding that quantum physics has the appropriate language and methods to solve many problems in biology, observing phenomena from a different point of view and developing new concepts. The next important steps are experimental verification/falsification,” Brookes said. “One could easily expect many more new exciting ideas and discoveries to emerge from the intersection of two major areas such as quantum physics and biology,” Omar concluded.  相似文献   

19.
20.
While Europe is locked in the debate about basic versus applied research, Louis Pasteur solved the problem more than 100 years ago. Antoine Danchin comments on Pasteur''s notion of ‘motivated research'' and how it leads both to new discoveries and to new applications.Three years ago, a senior politician attended his country''s Annual Congress for the Advancement of Science to give the introductory lecture. He asked the attending scientists to make science and research more attractive to young students and the general public, and asked his countrymen to support scientists to address the urgent challenges of global climate change, energy needs and dwindling water resources. It was neither a European nor a US politician, but the Indian Prime Minister Manmohan Singh who made this speech about the relationship between research and its practical applications. This is such an important topic that one might think it deserves appropriate attention in Europe, yet we fail to address it properly. Instead, we just discuss how science should serve society or contribute to the ‘knowledge-based economy'', or how ‘basic'' or ‘fundamental'' research is opposed to ‘applied'' or ‘industrial'' research and how funding for ‘big science'' comes at the expense of ‘little academic'' research.This dichotomy between the research to generate knowledge and the application of that knowledge to benefit humankind seems to be a recent development. In fact, more than 100 years ago Louis Pasteur avoided this debate altogether: one of his major, yet forgotten, contributions to science was the insight that research and its applications are not opposed, but orthogonal to each other (Stokes, 1997). If Niels Bohr ‘invented'' basic academic research—which was nevertheless the basis for many technological inventions and industrial applications—Pasteur developed what we might call ‘motivated'' research.How is research motivated and by what? By definition, scientists are citizens and members of the general public and, like the public, they are motivated by two forces: on the one hand, in Rudyard Kipling''s words, “man''s insatiable curiosity”; on the other hand, a desire for maintaining and improving their well-being. These are not contradictory to one another; curiosity nourishes dreams of a brighter future and leads to discoveries that contribute to well-being.Pasteur understood that it is essential to take account of society''s demands and desires; that science must be motivated by what people want. Still, there are severe misgivings about the nature of research. These stem from the mistaken but popular assumption that the scientists'' main task is to find solutions to current problems or to fulfil our desires. Problems and desires, however, are not enough, because finding solutions also requires creativity and discovery, which, by their very nature, are unpredictable. Often we do not even know what we need or desire and it is only through curiosity and more knowledge that we find new ways to improve our well-being. Motivation by itself is, therefore, not enough to lead to discovery. Motivation simply helps us choose between many different goals and an infinite number of paths to gain novel knowledge. Subsequently, each path, once chosen, must be explored using the scientific method, which is the only way to new discoveries.Motivation helps us to ask relevant questions. For example, why do wine and beer go sour without any apparent reason? Pasteur set out to design experiments that showed that fermentation is caused by microorganisms. A few years later, silkworms were suddenly dying of a terrible disease in the silk factories of southern France. The French government called on Pasteur for help, who eventually found that a parasite had infected silkworm eggs and proposed solutions to eradicate the disease. The original question therefore led to germ theory and bacteriology, helped to develop solutions to infectious diseases, and eventually created the whole field of microbiology.Motivation leads to conceptual and experimental research, which generates discoveries and new technologies. Discoveries, in turn, are the basic resource for the creation of general knowledge and the development of new products, services and other goods that fulfil public demands and generate jobs. The study of the ‘diseases'' of beer and wine also led to the development of fermentation processes that are still in use today. The same motivation that drove Pasteur in the nineteenth century now enables us to tackle current problems, such as pollution, by studying microbial communities that make compost or thrive in garbage dumps. Motivated research therefore reconciles our curiosity with the creation of knowledge and enables us to address pressing needs for humanity.Because it is strongly inspired by—even rooted in—society''s demands and desires, motivated research also raises accompanying ethical, legal, social and safety issues that should be compelling for all research. As mentioned above, scientists are members of the public who share the same concerns and demands as their fellow citizens and therefore participate with a general, public intelligence that, too often, is absent from academic research. This absence of ‘common sense'' or societal expectations generates the misunderstandings concerning research in biology and the development of biotechnology. These misconceptions—whether about the purported risks of genetically modified organisms or the exaggerated expectations for cancer therapies—can create real suffering in society and inefficient allocation of limited resources. It is therefore advisable for researchers to listen more to the public at large in order to find the motivation for their work.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号